paper_id
stringlengths
9
16
version
stringclasses
26 values
yymm
stringclasses
311 values
created
timestamp[s]
title
stringlengths
6
335
secondary_subfield
sequencelengths
1
8
abstract
stringlengths
25
3.93k
primary_subfield
stringclasses
124 values
field
stringclasses
20 values
fulltext
stringlengths
0
2.84M
1105.6094
1
1105
2011-05-30T20:00:01
Hybrid methods in planetesimal dynamics (I) : Description of a new composite algorithm
[ "astro-ph.EP", "astro-ph.GA" ]
The formation and evolution of protoplanetary systems, the breeding grounds of planet formation, is a complex dynamical problem that involves many orders of magnitudes. To serve this purpose, we present a new hybrid algorithm that combines a Fokker-Planck approach with the advantages of a pure direct-summation N-body scheme, with a very accurate integration of close encounters for the orbital evolution of the larger bodies with a statistical model, envisaged to simulate the very large number of smaller planetesimals in the disc. Direct-summation techniques have been historically developped for the study of dense stellar systems such as open and globular clusters and, within some limits imposed by the number of stars, of galactic nuclei. The number of modifications to adapt direct-summation N-body techniques to planetary dynamics is not undemanding and requires modifications. These include the way close encounters are treated, as well as the selection process for the "neighbour radius" of the particles and the extended Hermite scheme, used for the very first time in this work, as well as the implementation of a central potential, drag forces and the adjustment of the regularisation treatment. For the statistical description of the planetesimal disc we employ a Fokker-Planck approach. We include dynamical friction, high- and low-speed encounters, the role of distant encounters as well as gas and collisional damping and then generalise the model to inhomogenous discs. We then describe the combination of the two techniques to address the whole problem of planetesimal dynamics in a realistic way via a transition mass to integrate the evolution of the particles according to their masses.
astro-ph.EP
astro-ph
Mon. Not. R. Astron. Soc. 000, 1–42 (2011) Printed 6 July 2018 (MN LATEX style file v2.2) Hybrid methods in planetesimal dynamics (I) : Description of a new composite algorithm P. Glaschke1 , P. Amaro-Seoane(cid:63)2, 3, 4 & R. Spurzem3, 1, 5 1Astronomisches Rechen-Institut, Monchhofstrasse 12-14, Zentrum fur Astronomie, Universitat Heidelberg, Germany 2Max Planck Institut fur Gravitationsphysik (Albert-Einstein-Institut), D-14476 Potsdam, Germany 3National Astronomical Observatories of China, Chinese Academy of Sciences, 20A Datun Lu, Chaoyang District, 100012, Beijing, China 4Institut de Ci`encies de l'Espai (CSIC-IEEC), Campus UAB, Torre C-5, parells, 2na planta, ES-08193, Bellaterra, Barcelona, Spain 5Kavli Institute for Astronomy and Astrophysics, Peking University, China draft 6 July 2018 ABSTRACT The formation and evolution of protoplanetary systems, the breeding grounds of planet formation, is a complex dynamical problem that involves many orders of magnitudes. To serve this purpose, we present a new hybrid algorithm that combines a Fokker- Planck approach with the advantages of a pure direct-summation N−body scheme, with a very accurate integration of close encounters for the orbital evolution of the larger bodies with a statistical model, envisaged to simulate the very large number of smaller planetesimals in the disc. Direct-summation techniques have been historically developped for the study of dense stellar systems such as open and globular clus- ters and, within some limits imposed by the number of stars, of galactic nuclei. The number of modifications to adapt direct-summation N−body techniques to planetary dynamics is not undemanding and requires modifications. These include the way close encounters are treated, as well as the selection process for the "neighbour radius" of the particles and the extended Hermite scheme, used for the very first time in this work, as well as the implementation of a central potential, drag forces and the adjustment of the regularisation treatment. For the statistical description of the planetesimal disc we employ a Fokker-Planck approach. We include dynamical friction, high- and low-speed encounters, the role of distant encounters as well as gas and collisional damping and then generalise the model to inhomogenous discs. We then describe the combination of the two techniques to address the whole problem of planetesimal dynamics in a realistic way via a transition mass to integrate the evolution of the particles according to their masses. Key words: protoplanetary discs, planets and satellites: dynamical evolution and stability, methods: numerical, methods: N-body, methods: statistical 1 1 0 2 y a M 0 3 . ] P E h p - o r t s a [ 1 v 4 9 0 6 . 5 0 1 1 : v i X r a 1 INTRODUCTION The formation of a planetary system is closely related to the formation of the host star itself. Cool molecular clouds collapse and fragment into smaller substructures which are the seeds for subsequent star formation. Angular momen- tum conservation in the forming clumps forces the infalling matter into a disc-like structure. The subsequent viscous evolution of the disc leads to a transport of angular momen- tum which channels gas to the protostar in the centre. These protoplanetary discs are the birth place of planets (for a de- (cid:63) E-mail: [email protected] [email protected] [email protected] (RS) (PAS, (PG); Pau.Amaro- author); corresponding c(cid:13) 2011 RAS tailed review see Armitage 2010). Embedded dust grains are the seed for the enormous growth to bodies of planetary size. The first hint to the structure of protoplanetary discs has been provided by our own solar system. Through "smearing out" all planets and adding the missing fraction of volatile elements, one can estimate the structure and mass of the protoplanetary disc. Since the efficiency of planet formation is unknown, this yields only a lower limit – the minimum mass solar nebula (Hayashi 1981). The inferred surface den- sity decreases with the distance from the sun as ∝ r−3/2. Further insight has been obtained by the detection of an infrared excess of some stars, i. e. additional infrared radia- tion that does not originate from the star but an unresolved disc. Advancements in observation led in the mid-90s to the direct imaging of nearby star forming regions which opened a 2 P. Glaschke, P. Amaro-Seoane & R. Spurzem new flourishing field in astronomy (see O'dell et al. 1993,for an example with Hubble Space Telescope images)). Since then a big amount of observations of protoplanetary discs across the electromagnetic range has been obtained and in- terpreted, both space-based (ISO, Spitzer and now Herschel) and in the ground (VLT, VLT-I, Subaru, Keck). We refer the reader to "The Star Formation Newsletter" URL1 for an overview of the recent papers in star- and planet formation and the book review "Protostars and Planets V" (Reipurth et al. 2007). Protoplanetary discs masses cover a range from 10−3 to 0.1M(cid:12) with a peak around 0.01M(cid:12) (data from Tau- rus/Ophiuchus Beckwith 1996), in accordance with mass es- timates deduced from the minimum mass solar nebula. Since the disc lifetime can not be measured directly, it is derived indirectly from the age of young, naked (i. e. discless) stars which sets an upper limit. Pre-main sequence evolutionary tracks are used to gauge the stellar ages, providing lifetimes of a few 106 years. The subsequent evolution of the disc proceeds in several stages. Two different scenarios have been proposed to explain the formation of kilometre-sized planetesimals. (i) One process is the gravitational instability of the dust component in a protoplanetary disc that leads to the direct formation of larger bodies. Goldreich & Ward (1973) pro- pose that an initial growth phase of dust grains leads to a thin dust disc that undergoes a gravitational collapse. As the dense dust layer decouples from the gas, it rotates with the local Keplerian velocity, whereas the gas component ro- tates slower as it is partially pressure supported. This gives rise to a velocity shear at the boundary, which may excite turbulence through the Kelvin-Helmholtz-instability. Since the motion of small dust grains in the boundary layer is cou- pled to the gas, the turbulent velocity field could suppress the formation of a stratified dust layer, which is a necessary prerequisite for the gravitational instability (Weidenschilling 1977). (ii) The collisional agglomeration of dust particles is an opposed formation mechanism. Relative velocities are dom- inated by the Brownian motion in the early phases of the growth process. This mechanism becomes increasingly in- efficient with growing mass, but successively the particles decouple from the gas and settle to the midplane – a pro- cess that yields even larger velocities with increasing mass. The sedimentation initiates a growth mode that is similar to the processes in rain clouds: Larger grains drop faster, thus accreting smaller grains on their way to the midplane. Turbulence may modify this basic growth scenario by forc- ing the dust grains in a convection-like motion. Dust grains still grow during the settling process, but the turbulent ve- locity field could mix up dust from the midplane, and a new cycle begins. Each cycle adds a new layer to the dust grains – a mechanism that also operates in hail clouds – until the grains are large enough to decouple from the turbulent mo- tion. Again, turbulence plays an important role in determin- ing the growth mode and the relative velocities. While the relative velocities are high enough to allow for a fast growth, it is not clear a priory that collisions are sticky enough to 1 http://www.ifa.hawaii.edu/users/reipurth/newsletter. htm allow for a net growth. High speed encounters lead to frag- mentation, which counteracts agglomeration (e.g. Blum & Wurm 2000,and references therein). An important bottle- neck in the agglomeration process is the fast orbital decay of 1 m sized boulders. Their orbital lifetime is as short as 100 years, and a quick increase in size – at least over one order of magnitude – is needed to reduce the radial drift significantly. To overcome the difficulties associated with each of these scenarios, modifications have been proposed. Magneto- hydrodynamic simulations include electro-magnetic interac- tions in hydrodynamical calculations. See the reviews of Bal- bus & Hawley 1998 and Balbus. 2003. The MHD simulations by Johansen et al. (2006) show that trapping of larger parti- cles in turbulent vortices helps in increasing the orbital life- time, but could also trigger local instabilities that may lead to the direct formation of planetesimals (Inaba et al. 2005). Johansen et al. (2007) describe a gravoturbulence mecha- nism as a feasible pathway to planetesimal formation in ac- creting circumstellar discs. The details of agglomeration have drawn a lot of at- tention and are still under question (see Kempf et al. 1999; Paszun & Dominik 2009; Wada et al. 2009), but the succes- sive agglomeration of planetesimals is commonly accepted. 1.1 Formation of protoplanets The further growth of planetesimals proceeds through mu- tual collisions, where the initial phase involves a large num- ber of particles and is well described by a coagulation equa- tion (Safronov 1969). While earlier works (e.g. Nakagawa et al. 1983) focussed mainly on the evolution of the size dis- tribution, subsequent refinements of the evolution of the ran- dom velocities showed that it is important to evolve the size distribution and the velocity dispersion in a consistent way. A fixed velocity dispersion is an oversimplification, which changes the growth mode and increases the growth timescale as well (Wetherill 1989). The initial growth is quite democratic. All planetesimals grow roughly at the same rate and the maximum of the size distribution is shifted gently towards larger sizes. As soon as gravitational focusing and dynamical friction become im- portant, the growth mode changes to a qualitatively differ- ent mechanism. Efficient gravitational focusing leads to a growth timescale (which we denote as M/ M ) that decreases with mass. Hence larger bodies grow faster than the smaller planetesimals, a trend that is further supported by energy equipartition due to planetesimal–planetesimal encounters. This dynamical friction keeps the largest bodies on nearly circular orbit, thus the relative velocities are small and gravi- tational focusing remains efficient. Smaller planetesimals are stirred up into eccentricity orbits, which slows down their growth rate compared to the largest bodies. This acceler- ated growth, denoted as runaway growth of planetesimals (see e.g Greenberg et al. 1978; Wetherill 1989, 1993), short- ens the growth timescale to a few 105 years. The term run- away growth stresses that the growth timescale of a particle decreases with mass, hence the largest body "runs away" to the high mass end of the distribution (see Kokubo 1995). While energy equipartition increases the velocity dis- persion with decreasing mass of the planetesimals, addi- c(cid:13) 2011 RAS, MNRAS 000, 1–42 tional damping due to the gaseous disc leads to a turn-over at smaller sizes. However, higher relative velocities may lead nevertheless to destructive encounters, but these fragmen- tation events could even speed up the growth (Chambers 2006; Bromley & Kenyon 2006, 2010). Since smaller bodies are more subjected to gas drag, their velocity dispersion is smaller which allows an efficient accretion by the runaway bodies. Moreover, smaller particles damp the velocity dis- persion of the largest bodies more efficiently. As runaway proceeds, the system becomes more and more dominated by few big bodies – the protoplanets . Due to the dominance of few, very large bodies one can not use statistical methods anymore to simulate the problem. Kobayashi et al. (2010) derived analytical expressions for the final masses of these planetary "embryos", as they call them, including the role of planetesimal depletion due to collisional disruption. They conclude that the final mass in the minimum-mass solar nebula at several AU can achieve ∼ 0.1 Earth mass within 107 years. 1.2 Oligarchic growth The runaway growth of large bodies (i. e. protoplanets) ceases to be efficient as soon as the protoplanets start to control the velocity dispersion of the remaining planetesi- mals in their vicinity. Gravitational focusing becomes less effective, therefore the growth timescale increases with size and the growth mode changes to oligarchic growth. The pro- toplanets still grow faster than the field planetesimals2, but the masses of the protoplanets remain comparable. A combination of dynamical friction due to the field planetesimals and perturbations from the neighbouring pro- toplanets conserves a separation of five to ten Hill radii between neighbouring bodies. Therefore only planetesimals from a limited area, the feeding zone, are accreted by a given protoplanet. If this zone is emptied, they have reached their final isolation mass (Kokubo et al. 1998; Kokubo & Ida 2000, 2002). As the damping of the remaining field planetesimals is weak enough, further growth is dominated by mutual per- turbations among the protoplanets, which leads to giant im- pacts. Protoplanets beyond the "snow line" (or ice conden- sation point) can grow larger than 5–15 earth masses and initiate the formation of giant planets. If the protoplanetary disc is very massive in the inner planetary system, this may lead to an in-situ formation of hot Jupiters (see Bodenheimer et al. 2000). Kokubo et al. (2006) used the oligarchic growth model of protoplanets to statistically quantify the giant impact stage (i.e. collisions of protoplanets to form planets) with N−body simulations. They found that for steeper surface density profiles, large planets usually form closer to the star. 1.3 Migration The proposed three-stage scenario of planet formation cov- ers the dominant growth processes, but a major mechanism 2 The term "field planetesimals" denotes in the following the smooth component of smaller planetesimals. c(cid:13) 2011 RAS, MNRAS 000, 1–42 Hybrid methods: A new composite algorithm 3 is still missing – the migration of bodies in the system. Mi- gration is a generic term that summarises a set of different mechanisms that lead to secular radial drift of bodies (see e.g. the review of Papaloizou 2006; Armitage 2010): (i) The dissipation due to the remaining gas disc leads to an orbital decay of the planetesimals. While this poses a severe problem for 1 m–sized objects, larger bodies drift very slowly inward. One denotes this process as type 0 migration. (ii) Planets which are embedded in a gaseous disc launch spiral density waves at the inner and outer Lindblad reso- nances, which leads to an exchange of angular momentum with the resonant excited waves. This type I migration leads to a robust inward migration independent of the density pro- file. The perturbation from the planet is small, hence linear perturbation theory is in principle applicable (Ward 1986), but recent work proves that non-linear and radiative effects can be quite important (Paardekooper et al. 2011). (iii) If the protoplanet is massive enough, it opens a gap in the gaseous disc and excites waves through tidal interac- tion with the gaseous disc. An imbalance of the exchange of angular momentum with the inner and outer part of the disc leads to type II migration. The strong interaction be- tween planet and disc leaves the linear regime and requires a numerical solution of the hydrodynamic equations (Lin & Papaloizou 1979). (iv) Even with an opened gap, the planet still channels gas between the inner and outer part of the disc. While this corotational flow is to some extent already present during type II migration, it dominates the angular momentum bal- ance in the case of type III migration due to an asymmetry in the leading and trailing part of the flow. An imbalance allows for an efficient exchange of angular momentum with this corotational gas stream, which gives rise to a remark- ably fast migration (Masset & Papaloizou 2003). Type III migration is subjected to a positive feedback: A faster mi- gration increases the asymmetry in the corotational flow, which speeds up the migration. Both inward and outward migration are possible. The four migration types modify the three-stage scenario in different ways. Type 0 migration is most efficient for small planetesimals (i. e. 1 km or smaller). It poses a severe problem during the early stages of planet formation, as it may induce a signifi- cant loss of solid material, accompanied by a global change in the initial surface density (Kornet et al. 2001). The im- portance of this process diminishes as planetesimal growth proceeds, but it still leads to the loss of collisional fragments during the final disc clearing. Type I is most important for protoplanets (i. e. 0.1 earth masses or larger). It leads to an orbital decay of protoplan- ets, but this does not only imply a loss of protoplanets, but also breaks the conditions for isolation. Migrating bodies can accrete along their way through the disc and are thus not constrained to a fixed feeding zone, which may increase the isolation mass. Type II and type III mainly influences giant planets (larger than 10 earth masses) causing an inward or outward migra- tion, depending on the angular momentum exchange bal- ance. It can explain the large number of giant planets close to their host star (hot Jupiters) found in extrasolar systems. An important issue is the timescale of the migration process. 4 P. Glaschke, P. Amaro-Seoane & R. Spurzem If the migration is too fast, virtually all planets spiral inward and leave an empty system behind. Migration is a powerful process with the capability to reshape an entire planetary system, at least for gas giants, since terrestrial planets are likely not affected by migration in gaseous discs (Bromley & Kenyon 2011). However, it also requires some "parking mechanism" which terminates mi- gration before all planets (or protoplanets) are lost. Inho- mogeneities in the gaseous disc may change the crucial mo- mentum balance of the inner and outer part of the disc, thus stalling or even reversing the drift of a planet. The migration processes end after the dissipation of the gaseous disc due to photoevaporation or star-star encounters (i. e. after a few 106 to 107 years). The formation of a planetary system is a vital pro- cess that is driven by the interplay between the different growth phases and the migration of planets and protoplan- etary cores (i. e. the precursors of giant planets). While the preceding sections only summarised the main evolutionary processes, even more processes could influence the forma- tion of planetary systems. A fast accretion of giant planets in the outer parts of a planetary system could introduce fur- ther perturbations on the inner part and may even trigger the formation of terrestrial planets. Moreover, the stellar en- vironment in dense star clusters and multiple stellar systems also perturbs planet formation, which would require an even broader view on the problem. Last but not least, the exci- tation due to gravitational coupling to gas surface density fluctuations plays an important role (see e.g. Ida et al. 2008). Any approach to planet formation can hardly include this wealth of different phenomena, thus it is important to focus on a well-defined subproblem. In this work we focus on the formation of protoplanets for the following reasons: (i) The size, growth timescale and spacing of the proto- planets is a key element in the planet formation process. (ii) The protoplanet growth is well-defined by different growth modes. It starts with the already formed planetes- imals (≈ 1–5 km) and ends, when orbital crossing of the protoplanets initiates the final growth phase. (iii) The planetesimals are large enough to treat the re- maining gaseous disc as a small perturbation. (iv) The protoplanets are small enough to neglect tidal interaction with the disc in the inner planetary system. Col- lective planetesimal–protoplanet interaction are also negli- gible. Though the protoplanet formation is a well-posed subprob- lem, our approach has to incorporate various mechanisms and techniques to cover the full size range of the problem (see the review of Armitage 2010,for a extensive review of the problem and references therein). However, it is still accessi- ble to theoretical calculations to some extent which provide a safe ground for the analysis of the results. In next section 2 we briefly summarise the analytical model that we use for our initial models. Then we intro- duce the problem that we will address later, as well as the timescales and quantities of interest for the general prob- lem of planetesimal formation: In sections 3, 4, 5 we give a description for a test particle moving around a body with a central mass M , the Hill's problem and the protoplanet growth from a theoretical point of view, respectively. Subsequently, in section 6 and its corresponding subsec- tions, we ascertain the direct-summation part of the general algorithm, as well as the required modifications and addi- tions to tailor it to the specific planetesimal problem. In section 9 we introduce our collisional treatment, including fragmentation, collisional cascades, migration and coagula- tion. The statistical part of the algorithm is described in detail in section 16. Finally, in section 17 we explain how to bring together both schemes into a single numerical tool and in section 18 we give our conclusions and compare our scheme to other numerical approaches. 2 INITIAL MODELS The basis for all planet formation models is the structure of the protoplanetary disc. We summarise the pioneering work of Hayashi (1981) to have a robust initial model at hand. Subsequent evolution of the disc may change this simple approach, but it is still a valuable guideline. A basic estimate of the minimum surface density of solid material in the disc can be deduced from the mass and lo- cation of the present planets in the solar system: Σsolid(r) = 7.1 (r/1AU) 30.0 (r/1AU) −3/2 g/cm2 −3/2 g/cm2 0.35 (cid:54) r (cid:54) 2.7 2.70 (cid:54) r (cid:54) 36.0 (cid:40) The discontinuity at 2.7 AU stems from the location of the ice condensation point (or snow line) that allows the forma- tion of icy grains in the outer solar system. Furthermore, the total surface density is estimated through the chemical composition of the disc, which gives the ratio of gas to solids. A fiducial value is 1:0.017 (see Cameron 1973). The surface density of the gas component is therefore: Σgas(r) = 1700 (1) (cid:16) r (cid:17)−3/2 g 1AU cm2 Since the dust content is rather low, the gaseous component is transparent to the visible solar radiation. Thus the gas temperature follows from the radiation balance: (cid:16) r (cid:17)−1/2(cid:18) L (cid:19)1/4 1AU L(cid:12) T = T0 T0 = 280 K (2) L is the solar luminosity during the early stages, normalised by the present value L(cid:12). The three-dimensional density structure is given by an isothermal profile − z2 2h2 ρgas(r, z) = Σgas√ 2πh (cid:18) (cid:19) exp (3) with a radially changing scale height h: (cid:19)1/2 (cid:18) kBT (cid:19)1/8 (cid:17)5/4(cid:18) L µmH h = = cs Ω c(0) s Ω0 cs = (cid:16) r 1AU L(cid:12) µ = 2.34 c(0) s = 993.56 m s (4) cs is the sound velocity of an ideal gas with a mean molec- ular weight µ in units of the hydrogen mass mH. Since the density profile is related to a radially varying pressure, the gas velocity deviates from the local Keplerian velocity. The balance of forces relates the angular velocity Ωg to the pres- sure gradient: rΩ2 g = rΩ2 + 1 ρ dP dr (5) c(cid:13) 2011 RAS, MNRAS 000, 1–42 Hybrid methods: A new composite algorithm 5 Thus the angular velocity Ωg of the gas is (see e.g. Adachi et al. 1976): (cid:18) cs (cid:19)2 Ωg = Ω(cid:112)1 − 2ηg(r) ηg = − 1 2 d ln(ρgasc2 s) d ln(r) vK (6) e sin(φE) = (9) e cos(φE) = L = r × v rv2 GM − 1 r · v√ GM a It is more appropriate to formulate the rotation of the gaseous disc in terms of a velocity lag ∆vg normalised to the local Keplerian velocity vK : ∆vg = r(Ωg − Ω) ≈ −ηgvK (7) A typical value of ∆vg for the minimum mass solar nebula at 1 AU is ∆vg = −60 m/s. This simple model provides a brief description of the initial disc. However, the subsequent evolution further mod- ifies the structure of the protoplanetary disc. Since embed- ded dust grains are coupled to the gas, it is likely that a global migration of solid material changes the surface den- sity. Moreover, the dust grains are chemically processed, de- pending on the local temperature and composition which in- troduces additional spatial inhomogeneities. When the grow- ing particles pass the critical size of ∼ 1 metre, the strong onset of radial migration may lead to a final reshaping of the distribution of solid material. While these restrictions weaken the validity of this approach as the "true" initial model, it is still a robust guideline to choose reasonable sur- face densities for the solid and the gaseous component after the formation of planetesimals. 3 KEPLER ORBITS Planetesimals in a protoplanetary disc are subjected to var- ious perturbations: Close encounters change their orbits, a small but steady gas drag gives rise to a radial drift and accretion changes the mass of the planetesimals. While all theses processes drive the disc evolution on a timescale of at least a few thousand years, each planetesimal moves most of the time on an orbit close to an unperturbed Kepler ellipse. Though the protoplanetary disc introduces additional per- turbations, the central potential dominates for typical disc masses around 0.01 M(cid:12). Therefore the classical orbital ele- ments still provide a proper framework to study planetesimal dynamics. The orbital elements of a test particle moving around a mass M are: v2 2 − GM r E = a = − GM 2E 1 − L2 GM a (cid:114) e = cos(i) = Lz L c(cid:13) 2011 RAS, MNRAS 000, 1–42 a is the semimajor axis, e is the eccentricity and i is the inclination of the orbit. As long as no dominant body is structuring the protoplanetary disc, it is justified to assume axisymmetry. Hence the argument of the perihelion ω, the longitude of the ascending node Ω and the eccentric anomaly φE are omitted in the statistical description. The deviation of planetesimal orbits from a circle is quite small. Thus it is appropriate to expand the above set of equations. A planetesimal at a distance r0, in a distance z above the midplane and with a velocity (vr, vφ, vz) with respect to the local circular velocity vK has orbital elements (leading order only): a ≈ r0 + 2 e2 ≈ v2 r0vφ vK r + 4v2 φ v2 K i2 ≈ z2 r2 0 + v2 z v2 K (10) (11) (12) These expressions allow us later a convenient transforma- tion between the statistical representation through orbital elements and the utilisation of a velocity distribution func- tion. 4 HILL'S PROBLEM When two planetesimals pass close by each other, they ex- change energy and angular momentum and separate with modified orbital elements. Successive encounters transfer en- ergy between planetesimals with different masses, driving an evolution of the overall velocity distribution. It seems that an encounter is a two-body problem, as there are only two planetesimals involved, but the central mass has also a major influence turning the problem into a three-body encounter 3 The complexity of the problem is considerably simplified by reducing it to Hill's problem (Hill 1878). Consider two masses m1 and m2 that orbit a much larger mass Mc, where both masses are small compared to the central mass Mc. The mass ratio m1 : m2 could be arbi- trary. This special type of a three body problem is denoted as Hill's problem, originally devised to calculate the orbit of the moon. It provides a convenient framework to exam- ine planetesimal encounters in the potential of a star. The equations of motion4 of the two planetesimals including the 3 The term three-body encounter does not imply a close passage of all involved bodies, but emphasises the strong influence of a third one. 4 The following derivation is quite common to the literature (see e.g. Ida 1990; H´enon 1986). The later work uses a slightly different scaling. (8) 6 P. Glaschke, P. Amaro-Seoane & R. Spurzem central potential and their mutual interaction are: r1 = −r1 r2 = −r2 GMc r3 1 GMc r3 2 − (r1 − r2) − (r2 − r1) Gm2 r3 12 Gm1 r3 12 (13) We now introduce the relative vector r and the centre-of- mass R: r = r2 − r1 R = m2r2 + m1r1 m1 + m2 (14) Furthermore, the equations of motion are transformed to a corotating set of coordinates which are scaled by the mutual Hill radius rHill and the local Kepler frequency Ω x = y = z = r(cid:48) − a0 rHill a0(φ − Ωt) rHill z(cid:48) rHill (cid:114) m1 + m2 3 3Mc rHill = a0 (15) where (r(cid:48), φ, z(cid:48)) are heliocentric cylindrical coordinates. a0 is the radius of a properly chosen reference orbit, while the mutual Hill radius rHill is an intrinsic length scale of the problem : Since the two orbiting planetesimals are small compared to the central star, the Hill radius is much smaller than the size a0 of the reference orbit. Hence it is possible to expand the central potential about the reference orbit. This yields approximate equations for the relative motion and the centre-of-mass: x = 2 y + 3x − 3x/r3 y = −2 x − 3y/r3 z = −z − 3z/r3 X = 2 Y + 3X Y = −2 X Z = −Z (16) (17) The equations 16 and 17 have some interesting proper- ties: Firstly, the centre-of-mass motion separates from the interaction of the two bodies. Secondly, the scaled relative motion is independent of the masses m1 and m2, imply- ing a fundamental similarity of planetesimal encounters5. As Eq. 17 is a simple linear differential equation, one read- ily obtains the solution of the centre-of-mass motion X = b − e cos(t − τ ) Y = − 3 2 Z = i sin(t − ω) bt + ψ + 2e sin(t − τ ) (18) which is equivalent to a first-order expansion of a Kepler ellipse. ω and τ are the longitudes of the ascending node and the pericentre, while e and i are the eccentricity and 5 section 16 makes extensive use of this property. Figure 1. Equipotential lines for the effective potential U at z = 0 (see Eq. 20). U = U∗ refers to the largest allowed volume, which is enclosed by the Hill sphere and the two Lagrange points L1 and L2. inclination scaled by the reduced (i. e. dimensionless) Hill radius rHill/a0. The value of b depends on the choice of the reference orbit, but it is natural to set b = 0 which implies that the centre-of-mass defines the reference orbit. While the nonlinear nature of the relative motion (see Eq. 16) prevents any general analytical solution, Eq. 18 pro- vides at least an asymptotic solution for a large separation of the planetesimals, where b can be interpreted as an impact parameter. Nevertheless, small b do not necessarily imply close encounters, as opposed to the standard definition of the impact parameter. However, the expression b = a2 − a1 provides a measure of the distance of the two colliding bod- ies without invoking the complicated encounter geometry. A special solution to Eq. 16 are the unstable equilibrium points L1 and L2 at (x, y, z) = (±1, 0, 0), denoted as La- grange points 6. In addition, an inspection of Eq. 16 reveals that the Jacobi energy EJ is conserved: (cid:0) x2 + y2 + z2 + z2 − 3x2(cid:1) − 3 EJ = 1 2 (19) r Since the kinetic energy is always a positive quantity, the following inequality holds: EJ (cid:62) U = (cid:0)z2 − 3x2(cid:1) − 3 (20) 1 2 r Thus the allowed domain of the particle motion is enclosed by the equipotential surfaces of the effective potential U . A subset of these equipotential surfaces restricts the allowed domain to the vicinity of the origin (see Fig. 1). The largest of these surfaces passes through the Lagrange points L1 and L2. Hence we identify the Hill radius (or Hill sphere) as the maximum separation which allows the bound motion of two planetesimals7. 6 The additional Lagrange points L3–L5 are missing due to the Hill approximation. 7 The same argument applies to the tidal boundary in cluster c(cid:13) 2011 RAS, MNRAS 000, 1–42 -2-1.5-1-0.5 0 0.5 1 1.5 2 2.5-3-2-1 0 1 2 3xyL1L2U*Hill sphere which provide a convenient description of the initial relative orbit and the modified orbit after the encounter. M m > 3√ Beside the numerical solution of the equations of mo- tion, it is useful to define osculating (or instantaneous) or- bital elements b = 4x + 2 y i2 = z2 + z2 e2 = x2 + (3x + 2 y)2 3 2 ψ = −2 x + y + bt ω = t − arctan(z, z) τ = t − arctan( x, 3x + 2 y) EJ = 1 2 (e2 + i2) − 3 8 b2 − 3 r (21) (22) (23) 5 PROTOPLANET GROWTH Our work follows the evolution of a planetesimal disc into few protoplanets, including the full set of interaction pro- cesses. Hence we summarise the main aspects of protoplanet growth first to provide a robust framework. Although the sizes of the planetesimal cover a wide range, they virtually form two main groups: The smaller field planetesimals and the embedded protoplanets (or their precursors). This two-group approximation (e.g. Wetherill 1989; Ida 1993) allows one to have a clearer insight into the growth process. During the initial phase all planetesimals share the same velocity dispersion independently of their mass. The initial random velocities are low enough for an efficient gravita- tional focusing. Hence, the growth rate of a protoplanet with mass M radius R can be estimated as (e.g. Ida et al. 1993): (cid:18) (cid:19) M ≈ vrel Σ H πR2 1 + 2GM Rv2 rel (24) Eq. 24 is the two-body accretion rate, which should be mod- ified due to the gravity of the central star. Nevertheless this approximation gives an appropriate description to discuss the basic properties of the growth mode. The scale height H (Eq. 215) and the relative velocity vrel are related to the mean eccentricity em =(cid:112)(cid:104)e2(cid:105) of the field planetesimals: H ≈ vrel/Ω vrel ≈ emaΩ (25) Thus the accretion rate in the limit of strong gravitational focusing (2GM/R (cid:29) v2 rel) is: M ≈ 2πR ∝ M 4/3 ΣGM a2Ωe2 m (26) (27) If the protoplanets are massive enough, they start to control the velocity dispersion of the planetesimals in their vicinity. dynamics or the Roche lobe in stellar dynamics, which are equiv- alent to the Hill sphere. c(cid:13) 2011 RAS, MNRAS 000, 1–42 Hybrid methods: A new composite algorithm 7 (cid:114) The width ∆a of this sphere of influence, the heating zone, is related to the Hill radius RHill of the protoplanet (Ida 1993): ∆a = ∆aRHill = 4RHill (cid:114) M 3Mc h = 3 4 3 (e2 m + i2 m) + 12 (28) em and i are eccentricity and inclination of the field plan- etesimals, scaled by the reduced Hill radius h of the pro- toplanet. Mc is the mass of the central star. The condition that the protoplanet controls the velocity dispersion of the field planetesimals reads (Ida et al. 1993): 2M 2 2πa∆a > Σm (29) This condition is equivalent to a lower limit of the proto- planetary mass: (cid:18) π∆a (cid:19)3/5(cid:18) Σa2 (cid:19)3/5(cid:18) m (cid:19)3/5 (30) 3 Mc Mc M/m depends on several parameters, but reasonable values yield M/m ≈ 50–100. The velocity dispersion in the heated region is roughly v ≈ RHillΩ (31) which gives an interesting relation to the condition that leads to gap formation. A protoplanet can open a gap in the planetesimal component if it is larger than a critical mass Mgap (Rafikov 2001)  Σa2 Mc Σa2 Mc (cid:16) m (cid:16) m Mc (cid:17)1/3 (cid:17)1/3(cid:16) ΩrHill Mc v Mgap Mc ≈ (cid:17)2 v (cid:46) ΩrHill v (cid:29) ΩrHill if if (32) where rHill is the Hill radius of the field planetesimals. If the velocity dispersion v is controlled by the protoplanet, Eq. 32 together with Eq. 31 demonstrate that the condition for gap formation is equivalent to Eq. 30, i. e. the efficient heating of the field planetesimals implies gap formation and vice versa. The higher velocity dispersion of the field planetesimals (see Eq. 31) reduces the growth rate given by equation 26 to M ≈ 6πΣΩ ∝ M 2/3 RRHill e2 m (33) (34) Different mass accretion rates imply different growth mode. If two protoplanets have different masses M1 and M2, their mass ratio evolves as: (cid:18) M2 M2 (cid:19) d dt M2 M1 = M2 M1 − M1 M1 (35) When the growth timescale M/ M decreases with mass, a small mass difference increases with time. This is the case for Eq. 26, which gives rise to runaway accretion. As soon as the protoplanets control the velocity dispersion of the field planetesimals, the growth timescale increases with mass and therefore the protoplanet masses become more similar as they grow oligarchically. The field planetesimals also damp the excitation due 8 P. Glaschke, P. Amaro-Seoane & R. Spurzem 2 Σ[g/cm2] Miso/M(cid:12) 3.91 × 10−8 4.33 × 10−7 1.37 × 10−5 100 10 Miso/M⊕ 0.013 0.144 4.548 Table 1. Isolation mass for different surface densities at r = 1 AU and Mc = 1 M(cid:12). to protoplanet–protoplanet interactions and keep them on nearly circular orbits. The balance between these scatterings and the dynamical friction due to smaller bodies establishes a roughly constant orbital separation b (Kokubo 1997): (cid:114) 5 b = RHill b = b/RHill 7e2 mM 2πΣaR (36) (37) R is the radius of the protoplanet, M is its mass and e is the reduced eccentricity of the field planetesimals. The stabilised spacing prevents collisions between protoplanets, but it also restricts the feeding zone – the area from which a protoplanet accretes. If all matter in the feeding zone is ac- creted by the protoplanet, it reaches its final isolation mass (Kokubo & Ida 2000): Miso = 2πbaΣ (38) Inserting Eq. 36 yields the isolation mass in units of the mass of the host star Mc: (cid:18) 1 (cid:19)3/2(cid:18) a3ρ (cid:19)3/2(cid:18) a3ρ Mc 3 (cid:19)1/8(cid:0)e2 (cid:19)5/8(cid:18) 4π (cid:19)1/8 ≈ 19.67 ×(cid:0)e2 (cid:19)1/8 3 m m (cid:1)3/8 (cid:1)3/8 (cid:18) a2Σ (cid:18) a2Σ Mc Mc Mc (39) Miso/Mc = (112π4)3/8 There is a weak dependence on the density ρ of the pro- toplanet, but the most important parameter is the surface density Σ. If we take the minimum mass solar nebula as an example, the radial dependence of the surface density im- plies an isolation mass that grows with increasing distance to the host start. Hence protoplanets beyond some critical radial distance are massive enough (larger than ≈ 15 M⊕ Bodenheimer 1986) to initiate gas accretion from the proto- planetary disc. As the protoplanets approach the isolation mass, inter- actions with the gaseous disc and neighbouring protoplanets become increasingly important. We estimate the onset of or- bit crossing by a comparison of the perturbation timescale τpert of protoplanet–protoplanet interactions with the damp- ing timescale τdamp due to planetesimal–protoplanet scatter- ings. Since the protoplanets are well separated (b ≈ 5 . . . 10), it is possible to apply perturbation theory (Petit 1986,see e.g.): τpert ≈ b5 7hΩ (40) We anticipate section 16 (see Eq. 221) to derive the damping timescale τdamp ≈ 1 2 √ 2πG2 ln(Λ)(M + m)n0m T 3/2 r (41) where Tr and Tz are the radial and vertical velocity disper- sion of the field planetesimals. Hence the criterion for the onset of orbital crossing is: τpert < τdamp (42) As the protoplanets control the velocity dispersion of the field planetesimals (see Eq. 31), this condition reduces to: (cid:18) b (cid:19)4 ΣM > Σm ln(Λ) > Σm × f 72 7π e (43) Thus orbital crossing sets in when the mean surface den- sity ΣM of the protoplanets exceeds some fraction f of the field planetesimal density Σm. While the factor f depends strongly on the separation b of the protoplanets, a fiducial value is f ≈ 1, in agreement with the estimates of Goldreich et al. (2004). The onset of migration and the resonant interaction of protoplanets with the disc and other protoplanets ter- minates the local nature of the protoplanet accretion pro- cess and requires a global evolution of the planetary system. While the final stage deserves a careful analysis, further re- search is beyond the scope of this work. 6 DIRECT-SUMMATION TECHNIQUES FROM THE STANDPOINT OF PLANETARY DYNAMICS: FIRST STEPS The protoplanet formation is essentially an N –body prob- lem. Although we seek for a more elaborated solution to this problem which benefits from statistical methods, the pure N –body approach is a logical starting point. Direct calcu- lations with a few thousand bodies have provided us with a valuable insight into the growth mode (Ida 1992; Kokubo 1996,see e.g.), but they are also powerful guidelines that help developing other techniques. Statistical calculations rely on a number of approximations and "exact" N –body calcula- tions provide the necessary, unbiased validation of the de- rived formula. The choice of the integrator is a key element in the numerical solution of the equations of motion. Our require- ments are the stable long-term integration of a few ten thou- sand planetesimals with the capability of treating close en- counters, collisions and the perspective to evolve it into an improved hybrid code. Approximative methods like the Fast Multipole Method or Tree codes have a scaling of the com- putational time close to N , but the accuracy in this regime is too poor to guarantee the stable integration of Keple- rian orbits (compare the discussion in Hernquist et al. 1993; Spurzem 1999). The class of exact methods (all scale asymptotically with N 2) roughly divides in two parts: (i) Symplectic methods (see e.g. Wisdom 1991 or the Symba code, Duncan et al. 1998) rely on a careful expan- sion of the Hamiltonian which guarantees that the numerical integration follows a perturbed Hamiltonian. While there is c(cid:13) 2011 RAS, MNRAS 000, 1–42 still an integration error, all properties of a Hamiltonian sys- tem like conservation of phase-space volume are conserved by the numerical integration. The drawback of these very el- egant methods is that the symplecticity is immediately bro- ken by adaptive time steps, collisions or complicated exter- nal forces if no special precaution is taken. Even the numeri- cal truncation error breaks the symplecticity to some extend (Skeel 1999). Moore & Quillen (2010) recently developed a parallel integrator for graphics processing unit (GPU) that uses symplectic and Hermite algorithms according to the resolution needed. The algorithm is similar to Symba, but less accurate. (ii) The second group represents the "classical" methods that are based on Taylor expansions of the solution. They come in different flavours like implicit methods, predictor– corrector integrators or iterated schemes. Time symmetric methods stand out among these different approaches, as they show no secular drift in the energy error. These in- tegrators are so well-behaved that one may even call them "nearly symplectic". Taking all the requirements into account we have chosen Nbody6++8, an integrator which is the most recent descen- dant from the famous N –body code family from S. Aarseths' "factory". This version was parallelised by Spurzem (1999), which opened the use of current supercomputers. This parallel version, named Nbody6++, offers many versatile features that were included over the past years and more and more refined as time passed by. While all these elaborated tools deserve attention, we restrict ourselves for brevity to the components which contribute to the planetes- imal problem. The main components of the code are: (i) Individual time steps and a block time step scheme. (ii) Ahmad–Cohen neighbour scheme. (iii) Hermite scheme. (iv) KS–Regularisation for close two-body encounters. We will explain each of these features and highlight their advantages for the main goal of this work. This is impor- tant to understand the immediate first level of modifica- tions required to adapt the numerical scheme to planetary dynamics. We present later, in section 8, the next level of complexity in the modifications to be carried out to mold the numerical tool to our purposes. 6.1 Individual Time Steps The choice of the time step controls the accuracy as well as the efficiency of any given integrator. Too small time steps slow down the integration without necessity, whereas too large values increase the error. An efficient solution to this dilemma is an adaptive time step that is adjusted af- ter each integration step according to a specified accuracy limit. While the idea is quite clear, there is no unique receipt how to choose the proper time step. A common approach for N –body systems is to use the parameters at hand (like par- ticle velocity, force, etc.) to derive a timescale of the particle motion. This procedure leaves enough space for a wealth of 8 Aarseth (1999) gives a nice review on the remarkable history of the Nbody-codes, more details are given in Aarseth (2003). c(cid:13) 2011 RAS, MNRAS 000, 1–42 Hybrid methods: A new composite algorithm 9 different time step criteria. Nbody6++ uses the standard Aarseth expression (Aarseth 1985) (cid:115) ∆t = η F F (2) + (F (1))2 F (1)F (3) + (F (2))2 (44) which makes use of the force and the time derivatives up to third order. The time step choice is not unique in multi-particle sys- tems. One solution is to take the minimum of all these values as a shared (or global) time step, but this is not recom- mended unless all individual steps are of the same size. The second option is to evolve each particle track with its own, individual time step. This method abandons the convenient synchronisation provided by a global step, there- fore each integration of a particle demands a synchronisation of all particles through predictions. Since the prediction of all particles is O(N ), it is counter-balanced by the saving of force computations. Nevertheless the overhead is still sig- nificant, so a further optimisation might be desired. The basic idea of the block time step method is to force parti- cles in groups that are integrated together, which reduces the number of necessary predictions by a factor comparable to the mean group size. These groups are enforced through two constraints. The first condition is a discretisation of the steps in powers of two: −k ∆t = 2 k (cid:62) 0 (45) This condition increases the chances that two different par- ticles share the same timelevel, but it also reduces roundoff errors since the time steps are now exactly representable numbers. The second condition locks the "phases" of parti- cles with the same time step Ti ≡ 0 mod ∆ti (46) i. e. the particle time Ti is an integer multiple of the actual time step ∆ti so that all particles with the same step share the same block. Note that a step can not be increased at any time Ti, but only when the second condition (Eq. 46) for the larger step is met. 6.2 Ahmad–Cohen Neighbour Scheme The first N –body codes calculated always the full force (i. e. summation over all particles) to integrate a particle. But not all particles contribute with the same weight to the to- tal force. Distant particles provide a smooth, slowly vary- ing regular force, whereas the neighbouring particles form a rapidly changing environment which gives rise to an irregu- lar force. The Ahmad-Cohen neighbour scheme (Ahmad & Cohen 1973) takes advantage of this spatial hierarchy by di- viding the surrounding particles in the already mentioned two groups according to the neighbour sphere radius Rs. Both partial forces fluctuate on different timescales, which are calculated according to Eq. 44. The key to the efficiency of the method is the inequality ∆tirr (cid:28) ∆treg (47) Regular forces are extrapolated between two full force cal- culations Freg = F(0) reg + ∆t F(1) reg + 1 2 (∆t)2F(2) reg + 1 6 (∆t)3F(3) reg (48) 10 P. Glaschke, P. Amaro-Seoane & R. Spurzem Figure 2. All particles inside the neighbour sphere Rs are se- lected as neighbours (filled circles). As the neighbour list is fixed during regular steps (marked by black lines), a second shell in- cludes possible intruders. with a third order accurate expression, whereas the plain Nbody6++ uses only linear extrapolation (see the next sec- tions for a detailed discussion). 6.3 Hermite Scheme The Hermite scheme is a fourth-order accurate integrator that was applied first by Makino (1992) to the integra- tion of a planetesimal system. This scheme is used to in- tegrate single particles and CM-bodies9 in the main part of Nbody6++. It is accomplished by a predictor and a correc- tor step. The prediction is second order accurate: xp = x0 + v0∆t + vp = v0 + a0∆t + 1 2 1 2 a0(∆t)2 + 1 6 a0(∆t)3 a0(∆t)2 (49) Now the acceleration is evaluated at the predicted position to derive the second and third order time derivatives of the force: a(2) 0 = a(3) 0 = −4 a0 − 2 ap ∆t 6 a0 + 6 ap ∆t2 + −6a0 + 6ap + 12a0 − 12ap ∆t2 ∆t3 (50) (51) Using the additional derivatives one can improve the predic- tion: (cid:18) xc = xp + 1 a(2) 0 (∆t)4 + 24 − 3 = xp + 20 (∆t)3 + O(∆t6) 3 20 a0 + 1 (cid:19) 120 ap (cid:18) 7 0 (∆t)5 + O(∆t6) a(3) (∆t)2 − 1 30 a0 + 60 (cid:19) ap (52) The corrected positions are fourth order accurate. While the Hermite scheme is robust and stable, even in combination with the neighbour scheme, it is not accurate enough to integrate planetesimal orbits efficiently. 6.4 Hermite Iteration The Hermite scheme bears the potential for a powerful ex- tension, since it is a predictor–corrector scheme. An essential part of this scheme is the calculation of the new forces with the predicted positions, but it should improve the accuracy if they are recalculated using the corrected positions. The new forces are readily used to improve the corrected values, which closes the scheme to an iteration loop – the Hermite iteration (Kokubo et al. 1998). There are only few modifications necessary to obtain the iterated scheme. The particle prediction remains second- order accurate: xp = x0 + v0∆t + vp = v0 + a0∆t + 1 2 1 2 a0(∆t)2 + 1 6 a0(∆t)3 a0(∆t)2 (54) The force and its first time derivative are calculated to derive higher derivatives according to Eq. 50 and 51: ap = f (xp, vp) ap = f (xp, vp) (55) (56) The new corrector omits the highest order term in the posi- tion, making the velocity and the position to the same order accurate: (cid:18) 0 (∆t)4 + O(∆t5) 1 a(2) 24 − 1 (ap − a0)(∆t)2 + 1 6 4 (cid:18) (−a0 + ap)∆t + − 5 12 1 2 a0 − 1 12 ap a0 − 1 12 ap (cid:19) (cid:19) xc = xp + = xp + (∆t)3 + O(∆t5) vc = vp + (∆t)2 + O(∆t5) (57) (58) It seems unreasonable to drop one order in the position, but a reformulation of the predictor–corrector step reveals that this slight change yields a time symmetric scheme: ( ap − a0)(∆t)2 (ap − a0)(∆t)2 (ap + a0)∆t − 1 12 (vc + v0)∆t − 1 12 xc = x0 + vc = v0 + 1 2 1 2 (59) The iteration is achieved by returning to Eq. 55–56 with the predicted positions replaced by the more accurate values xc, vc. Although the integration does not need the second and third time derivative of the forces explicitly, it is useful to provide them at the end of the iteration for the calculation of the new time step: 1 a(2) 0 (∆t)3 + 24 (−a0 + ap)∆t + (cid:18) (cid:19) 0 (∆t)4 + O(∆t5) a(3) − 5 12 a0 − 1 12 ap 1 6 1 2 vc = vp + = vp + O(∆t5) (∆t)2+ (53) a(2)(t + ∆t) = a(3)(t + ∆t) = 2 a0 + 4 ap ∆t 6 a0 + 6 ap ∆t2 + + 6a0 − 6ap 12a0 − 12ap ∆t2 ∆t3 (60) (61) 9 CM denotes Regularisation for more details. centre–of–mass, see the section on KS- Numerical tests show that convergence is reached after one or two iterations, making this approach very efficient. It needs less force evaluations than the Hermite scheme for c(cid:13) 2011 RAS, MNRAS 000, 1–42 Rs the same accuracy. Moreover, its time symmetry suppresses a secular drift of the energy error. Finally, the derivatives are updated: Hybrid methods: A new composite algorithm 11 6.5 Extended Hermite Scheme The Hermite scheme is an integral part of Nbody6++ and proved its value in many applications. It would have been natural to improve the performance with an additional iteration, but our first tentative implementations showed rather negative results: The iterated scheme was more un- stable, slower and even less accurate than the plain Hermite scheme. An inspection of the code structure revealed that the Ahmad–Cohen neighbour scheme is the cause of this. Each particle integration is composed of two parts – frequent neighbour force evaluations and less frequent total force evaluations including derivative corrections. Every reg- ular correction leads to an additional change in the position of a particle, which introduces a spurious discontinuity in the neighbour force and its derivatives. The Hermite itera- tion reacts to this artificial jump in two ways: It increases the regular correction, and – what is more important – it am- plifies any spurious error during the iteration which leads to an extreme unstable behaviour. Since the Hermite iteration is a key element in the ef- ficient integration of planetesimal orbits, we sought for a modification of the Hermite scheme that circumvents the de- picted instability. The problem gives already an indication of a possible solution. A scheme with much smaller corrections would not suffer from the feedback of spurious errors. Nbody6++ stores already the second and third time derivative of the forces for the time step calculation. A man- ifest application of these derivatives at hand is the improve- ment of the predictions to fourth order: xp = x0 + v0∆t + a0(∆t)2 + a0(∆t)3 1 2 a(2) 0 (∆t)4 + 1 2 1 24 a(2) 0 (∆t)3 + + 1 24 + 1 6 vp = v0 + a0∆t + 1 6 a(3) 0 (∆t)5 1 120 a0(∆t)2 a(3) 0 (∆t)4 (62) (63) The prediction to fourth order was used in the iterative schemes of Kokubo et al. (1998); Mikkola & Aarseth (1998). Again, the new forces ap and ap are calculated to improve xp and vp – but with a modified corrector: a0(t + ∆t) = ap a0(t + ∆t) = ap (2)(t + ∆t) = a(2) (3)(t + ∆t) = a(3) n n + ∆t a(3) n a0 a0 (70) (71) (72) (73) The new scheme has an appealing property, which is related to the usage of the higher force derivatives. As the predictor is fourth-order accurate, it is equivalent to one full Hermite step. Since the corrector uses new forces to improve the two highest orders, it is equivalent to a first iteration step. Thus we obtained a one-fold iterated Hermite scheme at no ex- tra cost. This extended Hermite scheme reduces the energy error by three orders of magnitude, compared to the plain Nbody6++ with the same number of force evaluations. 6.6 KS–Regularisation Two bodies undergoing a close encounter are integrated in a special set of regular coordinates that separates the rela- tive motion from the motion of the centre-of-mass. A close encounter poses a numerical problem due to the singular- ity of the gravitational forces at zero separation. While the growing force amplifies roundoff errors as the two bodies approach each other closely, the collision is only an appar- ent singularity since the analytic solution stays well-defined. This opens the possibility of a proper coordinate transfor- mation which removes the singularity from the equations of motion. The Kustaanheimo–Stiefel regularisation takes ad- vantage of a four-dimensional set of variables to transform the Kepler problem into a harmonic oscillator (Kustaan- heimo & Stiefel 1965). Perturbations are readily included in the new set of equations of motion. The centre-of-mass is added as a pseudo-particle, the CM-body, which is integrated as a normal particle plus a perturbation force due to the deviation from a point mass. See Mikkola (1997) or Mikkola & Aarseth (1998) for more details. ap = f (xp, vp) ap = f (xp, vp) a(2) n (t) = a(3) n (t) = −4 a0 − 2 ap ∆t 6 a0 + 6 ap −6a0 + 6ap + 12a0 − 12ap ∆t2 + n − a0 (a(2) ∆t3 (2))(∆t)4+ ∆t2 1 24 xc = xp + n − a0 (a(3) 1 120 vc = vp + 1 6 (a(2) n − a0 (a(3) 1 24 (3))(∆t)5 n − a0 (3))(∆t)4 (2))(∆t)3+ c(cid:13) 2011 RAS, MNRAS 000, 1–42 (64) (65) (66) (67) (68) (69) 7 ADDITIONAL FORCES FOR PLANETESIMAL DISC DYNAMICS Nbody6++ only includes the gravitational interaction of all particles, therefore additional forces have to be added "per hand". A planetesimal disc requires two new forces: The presence of a central star introduces an additional central potential, while the gaseous component of the protoplane- tary disc is the source of a friction force. It is important that the new forces are properly included in the neighbour scheme to assure that regular steps remain larger than ir- regular steps. Since a dissipative force breaks the energy conservation, one has to integrate the energy loss as well to maintain a valid energy error control. In the next subsections we describe how we have done this. 12 P. Glaschke, P. Amaro-Seoane & R. Spurzem 7.1 Central Potential A star is much heavier than a planetesimal. Thus, the central star is introduced as a spatially fixed Keplerian potential: Φc = − GMc r F = − GmMc x x3 (74) Since the orbital motion of the planetesimals sets the domi- nant (and largest) dynamical timescale in the system, we in- cluded the central force as a component of the regular force. Moreover, the central potential also introduces a strong syn- chronisation, since planetesimals in a narrow ring share vir- tually the same regular block time step. 7.2 Drag Force As the whole planetesimal system is embedded in a dilute gaseous disc, each planetesimal is subjected to a small, but noticeable drag force. The drag regime10 depends on the gas density and the size of the planetesimals. Kilometer-sized planetesimals are subjected to the deceleration ρgasR2v − vg(v − vg) (75) = − πCD 2m dv dt CD = 0.5 (76) which is inversely proportional to the radius R(m) of the planetesimal in this drag regime. vg is the rotational velocity of the gaseous disc, which rotates slower than the planetes- imal system as it is partially pressure supported. The drag force leads to an orbital decay a of the semimajor axis of a planetesimals: a = − 3 4 CD ρgas ρBody (cid:104)(∆v)2(cid:105) R(m)Ω (77) Thus smaller particles migrate faster, with a maximum at R ≈ 1 m. Even smaller bodies couple to the gas, which reduces the effective drag force. The dissipation rate and its time derivative are: Wdrag = Fdrag · v Wdrag = Fdrag · v + Fdrag · Ftot (78) We integrate the dissipation rate Wdrag to maintain a valid energy error: (cid:90) t2 Wdrag dt ∆E = ∆t = t2 − t1 t1 (79) (80) ∆E = + 1 2 1 12 (Wdrag,1 + Wdrag,2) ∆t (cid:16) Wdrag,1 − Wdrag,2 (cid:17) ∆t2 + O(∆t5) (81) The expression is fourth order accurate in accordance with the order of the extended Hermite scheme. 10 The main drag regimes are Stokes (laminar flow), Epstein (mean free molecular path larger than object size) and Newton's drag law (turbulent flow). Weidenschilling (1977) provides a nice review on the different drag regimes. 7.3 Accurate integration of close encounters: Tidal perturbations of KS–Pairs and impact of the gaseous disc Both new forces also demand a modification of the regu- larisation treatment. They perturb the relative motion of a KS–pair and modify the orbit of the centre-of-mass. While the modification of the equations of motion is rather clear, the neighbour scheme requires some additional work. Let r1, r2 be the positions of the two regularised particles. The equations of motion read (G = 1) r = r2 − r1 r1 = −Mc r1 r3 1 r2 = −Mc r2 r3 2 + m2 − m1 r r3 + Fdrag,1 r r3 + Fdrag,2 (82) where the perturbations by other particles have been omit- ted for clarity. Centre-of-mass motion and the orbital motion are separated: r = −M R = −Mc r r3 − Mc m1 M r1 r3 1 r2 r3 2 − Mc + Mc m2 M r1 r3 1 r2 r3 2 + Fdrag,2 − Fdrag,1 m1 M Fdrag,1+ + m2 Fdrag,2 M r = r2 − r1 M = m1 + m2 R = 1 M (m1r1 + m2r2) (83) (84) (85) (86) Two new contributions show up due to the external forces: The KS–pair is tidally perturbed by the central star and influenced by the gaseous disc. While the aerodynamic properties of a single particle are well understood, two bod- ies revolving about each other may induce complex gas flows in their vicinity, which could invalidate the linear combina- tion of the drag forces on each component. Therefore we drop the drag force term to avoid spurious dissipation. Since the dynamic environment allows virtually no stable binaries11 in a planetesimal disc, the influence of the drag force on the encounter dynamics is negligible. We further decompose the additional acceleration of the centre-of-mass motion, since the neighbour scheme benefits from a clear separation of the timescales. Therefore, the tidal perturbation is split into a smooth mean force and a pertur- bation force: R = Fmean + Fpert Fmean = −Mc R Fpert = Mc R R3 R3 − Mc = 0 + O(r2) m1 M r1 r3 1 − Mc m2 M r2 r3 2 (87) (88) The mean forces varies on the orbital timescale and is hence 11 Tidal capturing of moons starts in the late stages of planet formation, but is limited to the planets or their precursors. How- ever, the quiescent conditions in an early Kuiper belt lead to a more prominent role of binaries. See the summary of Astakhov et al. (2005). c(cid:13) 2011 RAS, MNRAS 000, 1–42 Hybrid methods: A new composite algorithm 13 however, that the relation of regular and irregular costs is more complicated with GPU technology) (cid:115) ∆t = η F F (2) + (F (1))2 F (1)F (3) + (F (2))2 (89) where F (i) are the force and its time derivatives. It is applied to the calculation of the regular step using the regular force and accordingly to the irregular step based on the irregular force. The regular time step of a particle orbiting the central mass Mc at a distance r0 is: (cid:115) √ ηr ∆tr = 1 Ω Ω = GMc r3 0 (cid:112)ηirr/ηreg. The free parameters of the problem are the For simplicity, we introduce the scaled timestep ratio γt = γt mean particle distance ¯r, the velocity dispersion σv (addi- tional to the Keplerian shear), the particle mass mi and the neighbour number Nnb. We employ Hill's approximation for the central potential and obtain: (90) (91) (92) (93) (94) f is a yet unknown function. Dimensional analysis leads to √ ηi Ω 3Mc (cid:114) 2mi (cid:18) σv (cid:18) σv f √ ηi Ω ∆ti ≈ f (Ω, rHill, ¯r, σv, Nnb) rHill = r0 3 (cid:19) ∆ti ≈ γt ≈ f , ¯r rHill , Nnb (cid:19) , Nnb rHillΩ ¯r , rHillΩ rHill which shows that the time step ratio is essentially con- trolled by the interparticle distance and the velocity disper- sion. We generated different random realisations of planetes- imals discs with different densities and velocity dispersions to cover the range of possible values. The neighbour number is fixed to Nnb = 100 to reduce the noise due to small num- ber statistics, but γt converges to a value independently of the neighbour number already for Nnb > 10. Fig. 3 shows the numerical calculation of the time step ratio for various values of ¯r and σv. A good approximation to the calculated values of γt is: γt ≈ 1.79 × 1 + 1.03 σ2 v ¯r2Ω2 + 0.94 r3 Hill ¯r3 (95) Planetesimal discs have usually a small velocity dispersion (compared to the orbital velocity) and a low density in terms of the Hill radius, which leaves a major influence to the Ke- plerian shear. Since the shear motion is directly linked to the local Keplerian frequency, this synchronisation reduces γt to values smaller than ten. The numerical calculations show larger time step ratios with increasing velocity disper- sion and for high densities12, but planetesimal discs are far from these extreme parameter values. (cid:114) Figure 3. Time step ratio for Nnb = 100. Curves are plotted for different values of σv/(rHillΩ). The dotted line is approximation Eq. 95. included as a regular force component, while the perturba- tion is treated as an irregular force as it changes with the internal orbital period of the pair. 8 FURTHER TUNING THE DIAL: OPTIMISING N−BODY TO THE DISC An astrophysical simulation is a tool to analyse problems and predict dynamical systems which are not accessible to experiments. The design of a new simulation tool does not only require the careful implementation of the invoked physics, but also an analysis of the code performance to make best use of the available hardware. We want to apply Nbody6++ for the first time to planet formation, a subject that is quite different to stellar clusters. The central star forces the planetesimals on regu- lar orbits which need higher accuracy than the motion of stars in a cluster. In addition, the orbital motion also in- troduces a strong synchronisation among the planetesimals, thus allowing a more efficient integration. We examine the differences due to the integration of a disc system in the following sections. In particular, we will address in detail the role of the geometry of the problem and the neighbour scheme, the prediction of the number of neighbouring particles, the communication, the block size distribution and the optimal neighbour particle number for the direct-summation of the massive particles in the proto- planetary system 8.1 Disc Geometry and Neighbour Scheme The introduction of the neighbour scheme by Ahmad & Co- hen (1973) has provided us with a technique to save a con- siderable amount of computational time in star cluster simu- lations. Since the average ratio of the regular to the irregular time step γt is of the order of 10, the integration is speeded up by the same factor. One may expect a similar speedup for planetesimal systems, but in this case the time step ratio is roughly three. The time step is calculated with the stan- dard Aarseth time step criterion (it should be mentioned, c(cid:13) 2011 RAS, MNRAS 000, 1–42 1 10 100 1000 10000 0.001 0.01 0.1 1 10 100 1000γ~tr-/rHillTime Step Ratioσ/Ω rHill=0 σ/Ω rHill=2.2 σ/Ω rHill=10 Analytic 14 P. Glaschke, P. Amaro-Seoane & R. Spurzem a planetesimals system is smaller, as it is the case for the neighbour scheme. While a velocity dependent distance re- duces the number of necessary full force evaluations, it in- troduces a distance changing with time which destabilises the integration. The result is a much larger energy error compared to the achieved speedup. Therefore we only rec- ommend the mass modification of the apparent distance. 8.3 Neighbour Changes The rate at which the neighbours of a given particle change has a noticeable influence on the accuracy of the code. Dur- ing the course of an integration the second and third time derivative of the regular and irregular force are calculated from an interpolation formula (see Eq. 50 and Eq. 51). Whenever a particle leaves (or enters) the neighbour sphere, these derivatives are corrected by analytic expressions13. Hence many neighbour changes lead to a pronounced spuri- ous difference. We estimated the rate at which particles cross the neighbour sphere boundary to quantify this effect. Neigh- bour changes are due to the Keplerian shear and the super- imposed random velocities of the particles. The two effects lead to: N +/− N +/− = ∆tr Nnb Torb Shear = 3 2 ∆trNnb = ∆tr Nnb Torb σv Rnb 3πσv RnbΩ Dispersion (98) Rnb and Nnb are the neighbour sphere radius and the neighbour number, respectively. In practice, the neighbour changes due to the shear account for up to 80% of the total neighbour changes. The standard regular time step ∆tr = 2−5 and 50 neighbours yield a change of one par- ticle per regular step, which is fairly safe. 8.4 Neighbour Prediction Each integration step is preceded by the prediction of all neighbours of the particles that are due. A regular step re- quires the full prediction of all particles, so there is no possi- bility to save computing time. In contrast, an irregular step calculates only neighbour forces, which requires the predic- tion of less particles. Thus the prediction of all particles to prepare an irregular step is a simple, but, depending on the block size, computational costly solution. It seems to be more efficient to predict only the required particles, but ran- dom access to the particle data and the complete check of all neighbour list entries introduces an additional overhead. Therefore large block sizes should favour the first approach, whereas the second approach is more suitable for small block sizes. Both regimes are separated by a critical block size N∗ irr. If Nirr particles with (cid:104)Nnb(cid:105) neighbours are due, then only Nmerge particles need to be predicted: Figure 4. Regular steps per particle and per 1 N –body time in the inner core (Ri < 0.5) of a 5000 particle plummer model. Plotted are (1) different mass exponents with velocity exponent 0 and (2) different velocity exponents with mass exponent 1/6. 8.2 Optimal Neighbour Criterion The standard neighbour criterion uses the geometrical dis- tance: Particles are neighbours if their distance to the refer- ence particle is smaller than a limit Rs. This criterion is sim- ple and probably the best choice for an equal mass system. However, a multi-mass system may require a different crite- rion, since a massive particle outside the neighbour sphere could have a stronger influence than lighter particles inside the neighbour sphere. Also, relative effects are smaller at large distances. A more appropriate selection should rely on some "perturbation strength" of a particle. It turned out that a better criterion is the magnitude of the fourth time derivative of the pairwise force F (4) ij , i. e. those particles are selected as neighbours which pro- duce the largest integration error in accordance with the Hermite scheme. F (4) is a complicated expression (compare ij Appendix A), but the leading term can be estimated via dimensional analysis: ij ij ∝ mjv4 F (4) r6 ij (cid:18) mi (cid:19)1/6(cid:18) vs mj vij (cid:19)2/3 (96) (97) We use this expression to define a new apparent distance between the integrated particle i and a neighbour j: rapp = rij vs is an arbitrary scaling velocity to obtain a distance with dimension length. This new distance definition moves mas- sive or fast neighbours to an apparently smaller distance, thus enforcing that these particles are preferentially included in the neighbour list. In addition, the modified distance is readily included in the conventional neighbour scheme. We tested different mass and velocity exponents to verify that Eq. 97 is the optimal choice. Figure 4 shows that these ex- ponents are indeed the optimal choice for a Plummer model (Plummer 1911) with mass spectrum. The new scheme saves 25% of the force evaluations in the core, but the impact on 12 ¯r/rHill < 1 corresponds to an unstable self-gravitating disc. 13 Appendix A gives a complete set of the force derivatives up to third order. c(cid:13) 2011 RAS, MNRAS 000, 1–42 4.2 4.4 4.6 4.8 5 5.2 5.4 5.6 5.8 6 0 0.2 0.4 0.6 0.8 1Nreg/∆TExponent(1) V-Exponent=0(2) M-Exponent=1/6 Hybrid methods: A new composite algorithm 15 Process Send to 0 1 Receive from 7 1 2 0 2 3 1 3 4 2 4 5 3 5 6 4 6 7 5 7 0 6 Table 2. Ring Communication. Communication partners are fixed, while the exchanged data varies. np − 1 cycles are needed. Block Irregular Regular Irregular Regular np 10 10 20 20 α[µs] τl[µs] A B 0.35 0.22 0.35 0.22 51 145 4.5 113 308 512 40 877 8.8 368 1668 75 Cycle Process 1 2 3 Exchange with Exchange with Exchange with 0 1 2 4 1 0 3 5 2 3 0 6 3 2 1 7 4 5 6 0 5 4 7 1 6 7 4 2 7 6 5 3 Table 3. Hierarchical Communication. Communication partners change after every cycle. The exchanged data amount doubles with every new cycle, hence only ln2(np) cycles are needed. (cid:18) 1 − exp Nmerge ≈ Ntot Nirr(cid:104)Nnb(cid:105) (cid:18) − Nirr(cid:104)Nnb(cid:105) Ntot (cid:19)(cid:19) (cid:54) (99) The size Nmerge of the merged neighbour lists is smaller than the total number of neighbour list entries, since some particles are by chance members of more than one neighbour list. Performance measurements show that the prediction of the merged neighbour lists is 10 % more costly (per particle) than the full prediction, mainly due to additional sorting and a random memory access. Thus N∗ irr satisfies: Ntot = 1.1 × Nmerge Inserting Eq. 99 yields the critical block size: irr ≈ 2.4 ∗ N Ntot (cid:104)Nnb(cid:105) (100) (101) The prediction mode is chosen according to the actual block size. 8.5 Communication Scheme Nbody6++ is parallelised using a copy algorithm. A com- plete copy of the particle data is located on each node, so the integration step of one particle does not need any com- munication. Therefore a block of Nbl particles is divided in np parts (np is the processor number), which are inte- grated by different processors in parallel. The integration step is completed by an all–to–all communication of the different subblocks to synchronise the particle data on all nodes. Hence the amount of communicated data is propor- tional to Nbl × np. A communication in a ring-like fashion (see table 2) needs np − 1 communication cycles, but a hi- erarchical scheme (see table 3) sends the same amount of data with only ln2(np) communication cycles. The differ- ence between the two approaches remains small, as long as the communication is bandwidth limited, i. e. the blocks are large. Small block sizes shift the bottleneck to the latency, which is significantly reduced by the second scheme – espe- cially if the code runs on many processors. c(cid:13) 2011 RAS, MNRAS 000, 1–42 Table 4. Timings on a Beowulf cluster (Hydra, see table 5). See text for an explanation of the variables. Timings are obtained for a maximal neighbour number LMAX=64. In practice, B is twice as large due to storage rearrangements in Nbody6++. See Appendix 5 for details on the computers. Block np α[µs] τl[µs] A Irregular Regular Irregular Regular Irregular Regular 8 8 16 16 64 64 0.29 0.60 0.28 0.60 0.27 0.46 B 1.7 4.1 1.7 6.7 255 837 981 1763 700 561 188 306 241 401 887 7.7 871 21.7 Table 6. Timings on the IBM. More than 32 processors require more than one node. A hierarchical scheme reduces the latency, but never- theless it is possible that the parallel integration is actually slower than a single CPU integration. We estimated both the runtime on one CPU and on a parallel machine to ex- plore the transition between these two regimes. The latency time τl per communication is included in the wallclock time expressions for one regular/irregular step: τl = αA tsingle = αNblNnb (cid:16) NblNnb (cid:124) (cid:123)(cid:122) (cid:125) np Arithmetic tpar = α + A ln2(np) + (cid:124) (cid:123)(cid:122) (cid:125) (cid:124)(cid:123)(cid:122)(cid:125) BNbl Latency Communication (cid:17) (102) (103) If tsingle (runtime on a single CPU) is equal to tpar (parallel computation), one can deduce the critical block size Nmin which gives the minimal block size for efficient parallelisa- tion: tsingle = tpar Nmin = A ln2(np)np Nnb(np − 1) − Bnp (104) The hierarchical communication gives a minimal block size that increases logarithmically with the processor number. Eq. 103 gives immediately the speedup S and the optimal processor number for a certain block size Nbl: S = np,opt(Nbl) = np ln2(np) NblNnb 1 + Anp ln(2)Nbl × Nnb A + B np Nnb − Hierarchical (105) A comparison to the optimal processor number for a ring 16 P. Glaschke, P. Amaro-Seoane & R. Spurzem Beowulf (Hydra) Titan JUMP ARI/ZAH ARI/ZAH Forschungszentrum Julich Name Institute Location Processors Speed Heidelberg Heidelberg 20 64 2.2GHz 3.2GHz Julich 1248 1.7GHz 32 Processors/Node 2 2 Network Myrinet Infiniband Gigabit–Ethernet Bandwidth 2Gbit/sec 20 Gbit/sec 10 Gbit/sec Table 5. Specs of the different supercomputers used in running the algorithm communication np,opt(Nbl) = (cid:114) (cid:16) NblNnb Nbl × Nnb A np − Ring (cid:17) (106) tpar = α + Anp + BNbl stresses the efficiency of the hierarchical communication, since it allows a much larger processor number for a given problem size. Equation 102 and 103 are also useful to derive the total wallclock time, since the total runtime scales with the number of regular and irregular blocks: Nreg ≈ T Nirr ≈ T N 1/3 √ ηreg N 1/3 nb N 1/3√ ηirr (107) (108) These equations are only approximate expressions, but they give the right order of magnitude without detailed calcula- tions that need a precise knowledge of the N –body model. Table 4 and table 6 summarise the timing parameters drawn from our experience with the Hydra and JUMP parallel su- percomputers. 8.6 Block Size Distribution The preceding section showed that the block size is closely related to the efficiency of the parallelisation. Small blocks are dominated by the latency and the parallelisation could be even slower than a single CPU calculation. Therefore we derive the block size distribution for the block time step scheme to asses its influence on the efficiency. Suppose that the time steps14 h of all N particles in the model are distributed according to some known function f : Figure 5. Timestep distribution f = dp/dh. The short-dashed line on the left indicates approximation Eq. 110, whereas the dashed line on the right defines a reasonable upper limit hmax. restricts the major fraction of the time steps to a finite in- terval. Thus it is possible to capture the main features of the time step distribution with an expansion around h = 0 (Fig. 5 sketches this approximation): f ≈ C(N )ha h (cid:54) hmax (110) a is the lowest non-vanishing order of the expansion. Now we consider a block level with the largest possible time step hk. The number of particles Nbl in this block is: (cid:90) hk 0 dp = f (N, h)dh (109) Nbl = N f dh ≈ C(N ) a + 1 (hk)a+1 (111) f is in most cases a complicated function. It involves spa- tial averaging and integration over the velocity distribution, which could be quite complicated even for simple time step formulae. Nevertheless there is a constraint on the time step distribution, simply because every particle has a neighbour within a finite distance: There is some upper limit hmax that According to the block time step scheme the number of blocks per time with the largest possible time step hk is pro- portional to (hk)−1. Therefore the probability that a block size is in the range [Nbl, Nbl + dNbl] is (cid:19) 1 hk Nbl − C(N ) a + 1 ha+1 k dNbl (112) dp ∝(cid:88) δ (cid:18) k 14 We use h instead of ∆t in this section to avoid unclear nota- tion. where δ is Dirac's delta function. The sum over the loga- rithmically equidistant time steps hk is approximated by an c(cid:13) 2011 RAS, MNRAS 000, 1–42 hdpdhfhmax Hybrid methods: A new composite algorithm 17 Figure 7. Optimal neighbour number as function of particle number N . The plot includes the numerical solution of Eq. 124 and the two asymptotic solutions. Timing constants are taken from a Beowulf cluster. as all the neighbour lists have to be broadcast to synchro- nise the different nodes. The best choice balances these two extremes, thus maximising the speed. Before we derive the optimal neighbour number on a parallel machine, we briefly summarise the known solution for a single CPU run (Makino 1988) for an extensive deriva- tion. The computational effort of the irregular steps is pro- portional to the neighbour number, while the number of force evaluations for the regular steps is proportional to the total number of particles, reduced by the time step ratio γt: γt := ∆treg ∆tirr (cid:18) TCPU = f (N ) γt(Nnb) ≈ N 1/3 nb (cid:19) Nnb + N γt(Nnb) (116) (117) f (N ) collects all factors depending only on the total num- ber of particles. Optimisation with respect to the neighbour number Nnb yields the well known result: d 0 = dNnb Nnb,opt ∝ N 3/4 TCPU (118) The calculation of the elapsed time for Nbody6++ on a PC cluster includes more terms. For clearness, we restrict ourselves to a rather simple model that involves only the dominant terms to show how parallelisation influences the optimal neighbour number. We make the following approx- imations: (i) We only take the force calculation and communication into account. Figure 6. Cumulative irregular block size distribution for a N = 2500 particle Plummer model. (cid:18) integral: dp ∝ (cid:90) ∞ δ 0 ≈ 1 a + 1 Nbl − C(N ) a + 1 −(a+2)/(a+1) bl N dNbl (cid:19) d ln(h) h ha+1 dNbl (113) Thus the average block size and the median of the block size distribution are: (cid:104)Nbl(cid:105) ≈ 1 a N a/(a+1) median(Nbl) ≈ 2a+1 (114) Special expressions for the average block size were already derived by Makino (1988), but the general relation of the time step distribution to the block size distribution is a new result. The median is surprisingly independent of the parti- cle number, i. e. 50 % of all blocks are always smaller than a fixed value. It seems that this is a threat to the efficiency of the method, but the median of the wallclock time median(N 2 bl) ≈ N 2(a+1)/a (115) demonstrates that these small blocks account only for a small fraction of the total CPU time. We confirmed the de- rived block size distribution (Eq. 113) by numerical calcula- tions (see Fig. 6). The order parameter a is roughly two in (at least locally) homogenous systems, while an additional Keplerian potential reduces the order to a = 1. A planetesi- mal disc – or more precisely, a narrow ring of planetesimals – has a very narrow distribution of time steps since all par- ticles share nearly the same orbital period. Thus the regular block size is always equal to the total particle number mak- ing the parallelisation very efficient. 8.7 Optimal Neighbour Number (ii) We use the same time constants for regular and irreg- We treated the mean neighbour number Nnb so far as some fixed value. But it is also a mean to optimise the speed of the integration. Large neighbour spheres reduce fluctuations in the regular forces allowing larger regular steps, which reduces the total number of force evaluations. But larger neighbour lists also imply a larger communication overhead, ular expressions. (iii) We neglect all numerical factors that are comparable to unity. The total CPU time is an extension of Eq. 106, which is applied to the regular and the irregular step. A new constant Bn includes the neighbour list communication separately, c(cid:13) 2011 RAS, MNRAS 000, 1–42 1e-04 0.001 0.01 0.1 1 1e-04 0.001 0.01 0.1 1p(>N)N/NtotPlummer Model N=2500Theory a=20.11101001000100001000001101001000100001000001e+061e+071e+08NnbNNnb,optN1/5N3/5 18 P. Glaschke, P. Amaro-Seoane & R. Spurzem while all factors depending on N are represented by f (N ): Nbl ≈ N 2/3 γt ≈ N 1/3 nb Tirr = f (N ) Treg = 1 γt f (N ) Ttot = Tirr + Treg (cid:18) NblNnb (cid:18) NblN p p (cid:19) + Ap + BNbl + Ap + (B + BnNnb)Nbl (cid:19) (119) Optimisation with respect to the processor number p leads to: 0 = popt = Ttot ∂ ∂p (cid:118)(cid:117)(cid:117)(cid:116) N 2/3N 4/3 nb + N 5/3 nb + 1) A(N 1/3 √ ≈ N 5/6 AN 1/6 nb (120) Further optimisation with respect to the neighbour number gives the expression: ∂ 0 = Ttot ∂Nnb nb − 1 0 = N 4/3 3 N − 1 3 2 3 NnbBnp AN −2/3p2 − 1 3 Bp + (121) (122) For a fixed p or Bn = 0 (very fast neighbour list communi- cation), we recover for large N : Nnb,opt ∝ N 3/4 (123) In general, one can not neglect the neighbour list commu- nication. Therefore we seek for the optimal choice of p and Nnb, thus combining Eq. 120 and Eq. 122: of the neighbour list communication favours much smaller neighbour numbers. Nnb increases so slowly with the particle number that a neighbour number around 100 is a safe choice. 9 COLLISIONAL AND FRAGMENTATION MODEL The growth of planetesimals proceeds through collisions among planetesimals which form (at least in a sufficient frac- tion of incidents) larger bodies with a net gain of accreted matter. But some collisions are mere destructive events that shatter and disperse the colliding planetesimals. Small bod- ies are more susceptible to destruction, but they are also driven to high relative velocities due to the global energy equipartition making them even more vulnerable. A model that attempts to cover the full size range from objects rang- ing between a kilometer and the size of Mars needs a realistic collision algorithm that covers both fragmentation and ac- cretion. Some examples in the literature are Cazenave et al. (1982); Beauge & Aarseth (1990). In our algorithm we use the approach of Glaschke (2003), which was applied to as- teroid families. The following section describes the approach to fragmentation that is implemented in the code. 9.1 Handy quantities for quantifying the models Two colliding bodies are equal in the sense that their intrin- sic properties are not different. Only the comparison of two bodies defines the larger body – usually denoted as target – and the smaller one denoted as projectile. The two terms stem from laboratory experiments where they indicate much more than different sizes. A small projectile is shot on a target at rest to study the various parameters related to fragmentation. In the following, projectile and target only indicate the relative size of the two bodies. The collision of two bodies initiates a sequence of com- plex phenomena. Shock waves run through the material, flaws start to grow rapidly breaking the bodies apart in many pieces. Some kinetic energy is transferred to the frag- ments, which leads to the ejection of fragments at differ- ent velocities in various directions. If the fragment cloud is massive enough, some of the larger fragments may capture debris. This post-collisional accretion is denoted as reaccu- mulation. Although the depicted scenario is quite complex, there are a few measures that describe the most important as- pects: (i) Mass of the largest fragment ML, or dimensionless fl = ML/M where M is the combined mass of the two colliding bodies. (ii) fl < 1 2 refers to fragmentation, whereas fl > 1 2 is denoted as cratering. (iii) Energy per volume S that yields fl = 1 2 is denoted as impact strength. (iv) fKE := 2Efrag that is converted into kinetic energy of the fragments. kin /Ekin: Fraction of the impact energy Different fragment sizes and velocities are summarised by appropriate distribution functions. mi, Di and vi are mass, diameter and modulus of the velocity of a given fragment, respectively. c(cid:13) 2011 RAS, MNRAS 000, 1–42 (cid:18) 2 (cid:19) AN 5/3 nb + Nnb − B 3Bn 3 1 3 (N 1/3 nb + 1)N A Nnb,opt ≈ N 1/5 (cid:19)3/5 (cid:18) A (cid:19)3/5 (cid:18) N 4B2 n √ AN 1/6 nb N 5/6 = Bn (124) N (cid:29) (cid:19)3/2 (cid:18) 3A (cid:19)3/2 (cid:18) 3A 4B2 n (125) (126) Since this equation has no closed solution, we identify the dominant terms in Eq. 124 to calculate the asymptotic solution for large N : For small N we get the approximated solution: Nnb,opt ≈ 3 N < 4B2 n Fig. 7 compares the approximate expressions with the nu- merical solution of equation 124. In spite of the complicated structure of Eq. 124, both approximate expressions are re- liable solutions. The example uses timing constants derived form our local Beowulf cluster: A ≈ 200 B ≈ 5 Bn ≈ 0.5 (127) If we compare the new optimal neighbour number to the single CPU expression (Eq. 118), we find that the influence Hybrid methods: A new composite algorithm 19 Figure 9. Impact strength according to Eq. 128 and Eq. 130. The right axis gives the corresponding impact velocity according to S = 1/2ρv2 with ρ = 2.7 g/cm3. (i) Fragment size distribution: S = S0 Figure 8. Section for fl = 0.04 and n = 3. The largest fragment is coloured in dark-grey. In this calculation 60× 60× 60 grid cells are used. Note the decomposition in grid cells and the Voronoy polyhedra which form the fragments. m, (a) Nm(m) : Number of all fragments with a mass mi (cid:62) (b) M (m) : Mass of all fragments with a mass mi (cid:62) m, (c) ND(D) : Number of all fragments with a diameter Di (cid:62) D. The distribution functions are related to each other: Nm(m) = ND(D(m)) (cid:90) ∞ (cid:90) ∞ m m (cid:12)(cid:12)(cid:12)(cid:12) dx (cid:12)(cid:12)(cid:12)(cid:12) dNm(x) (cid:12)(cid:12)(cid:12)(cid:12) dx (cid:12)(cid:12)(cid:12)(cid:12) dM (x) dx dx x 1 x M (m) = Nm(m) = D(m) is the size–mass relation. (ii) Velocity distribution: (a) ¯v(m): mean velocity as a function of mass. 9.2 Prediction of collisional outcomes: Derivation from a Voronoy tessellation Any theoretical or empirical prescription of a collision has to relate the afore mentioned parameters, namely the impact energy, to the sizes and velocities of the produced fragments. The central quantity is the impact strength, which is a mea- sure for the overall stability of a body. Objects smaller than 1 m are accessible to laboratory experiments, while collisions of larger bodies up to asteroid size have to be analysed by complex computer simulations. Asteroid families, which are remnants of giant collisions in the asteroid belt, provide in- dependent insight, although the data is difficult to interpret. c(cid:13) 2011 RAS, MNRAS 000, 1–42 We selected two different impact strength models as reference for our work. The first was obtained by Housen (1990) through the combination of asteroid family data and laboratory experiments via scaling laws: (cid:19)−0.24(cid:34) (cid:18) R 1 m 1 + 1.6612 × 10 −7 (cid:19)1.89(cid:35) (cid:18) R 1 m fKE = 0.1 S0 = 1.726 × 106 Jm −3 = 1.726 × 107erg cm −3 (128) (129) Later, Benz (1999) obtained another result through SPH simulations (for basalt, v =3 km/s): 1 + 6.989 × 10 −5 (cid:18) R (cid:19)−0.38(cid:34) 1 m S = S0 fKE ≈ 0.01 S0 = 6.082 × 105 Jm ρ = 2.7 g cm3 −3 (cid:19)1.74(cid:35) (cid:18) R 1 m (130) (131) (132) fKE is a measure of the kinetic energy that is transferred to the fragments: Efrag kin = fKE 2 Ekin (133) We introduce a dimensionless measure γ of the relative im- portance of gravity for the result of a collision. It is defined as the ratio of the energy per volume SG that is necessary to disperse the fragments to the impact strength S0: 4 − 2 3√ 5fKE 2 GR2ρ2 SG = 2π γ := SG/S (134) The first step towards the prediction of a collisional out- come is to relate the impact energy and the impact strength to ascertain the size of the largest fragment fl. Laboratory 100000 1e+06 1e+07 1e+08 1e+09 1e+10 0.01 0.1 1 10 100 1000 10 100 1000S[J/m3] v[m/s]R[km]Housen 1990Benz 1999 20 P. Glaschke, P. Amaro-Seoane & R. Spurzem experiments and simulations indicate the functional form (cid:26) 2(1 − fl) (2fl)− 1 for fl > 1 2 K otherwise. (fl) = The distribution function is evolved by the coagulation equa- tion. We modified the equation given by Tanaka et al. (1996) by introducing a new function Mred to arrive at a more con- cise expression:  = Ekinρ 2SM (135) 0 = mn(t, m) + ∂ ∂m Fm(t, m) (136) The mass flux Fm is given by: Fm = − n(t, m1)n(t, m2)ξdm1dm2 ξ ≡ σ(m1, m2)vrelMred(m, m1, m2) ∂ ∂t (cid:90)(cid:90) (cid:90) which is both valid in the fragmentation regime and the cra- tering limit. The size of the largest fragment is used to derive the full size distribution. To accomplish the decomposition "seed fragments" are distributed inside the target accord- ing to the largest desired fragment. The full set of fragment is derived from a Voronoy tessellation 15 using these seed points. Fig. 8 depicts the result of such a decomposition. The fragment velocities are calculated from the total kinetic energy after the collision to initiate a post-collisional N – body calculation to treat reaccumulation. We conducted a large set of such calculations to cover a sufficient range in f i l (i. e. impact energy) and γ (i. e. body size). Table 7 summarises the derived values of the largest and second largest fragment including reaccumulation. 10 COLLISIONAL CASCADES A first well-defined application of the fragmentation model is a collisional cascade. The term cascade denotes that frag- ments of one collision in a many-body system may hit other bodies, whose fragments further shatter even more bodies. Thus the particle number increases exponentially with every subsequent collision. Although the formation of planets requires a net growth due to collisions, this destructive process plays a role in the formation of larger bodies as the overall size distribution controls the accretion rate of the protoplanets. Therefore it is worth to have a closer look into this mechanism. 10.1 Self–similar collisions A system of colliding bodies is usually embedded in a broader context, like stars moving in a galaxy or asteroids orbiting in our own solar system. First, we simplify this dy- namical background as well as some aspects of the collisions to make the problem tractable. The first step is to decompose an inhomogeneous sys- tem into smaller subvolumes which are locally homogenous. Furthermore, it is assumed that these subvolumes hardly interact with each other. Hence it is possible to apply the particle–in–a–box–method (Safronov 1969) to analyse colli- sions within the small subvolumes: (i) All particles are contained in a constant volume. (ii) The particle sizes are described by a distribution func- tion n(m), i. e. the particle number per volume and mass interval. (iii) For convenience, we assume a constant (or typical) relative velocity for a given pair of colliding bodies. 15 The Voronoy tessellation assigns every volume element to the closest seed point. First applications date back to the 17th cen- tury, but the Russian mathematician of Ukrainian origin Georgy Feodosevich Voronoy put it on a general base in 1908. (137) (138) (139) Mtot = n(t, m)mdm ∂ ∂t Mtot = Fm(mmin), where ξ is the coagulation kernel, n is the already intro- duced size distribution,vrel is the mean relative velocity, σ is the cross section for colliding bodies (m1, m2) and Mred is the newly introduced fragment redistribution function. Mred contains all information on the fragments arising from the breakup of body m1 due to the impact of body m2. Its defi- nition avoids double counting of collisions in the above inte- gral. The redistribution function is related to the differential number distribution function ncoll(m1, m2, m), i. e. the num- ber of fragments produced by a collision per mass interval. Since the target m1 formally disappears, it is included as a negative contribution: (cid:90) m Mred(m, m1, m2) := (cid:0)ncoll(m1, m2, m) − δ( m − m1)(cid:1) md m (140) 0 Mass conservation in each collision is reflected by Mred(0, m1, m2) = Mred(∞, m1, m2) = 0. The cross section σ depends on the velocities and radii Ri of the particles. A simple approach is the geometric cross section: σ(m1, m2) = π(R1 + R2)2 (141) If gravity plays an important role during encounters, two colliding bodies move on hyperbolic orbits with a pericentre distance that is smaller than the impact parameter. This leads to an additional enlargement of the cross section, de- noted as gravitational focusing: σ(m1, m2) = π(R1 + R2)2 1 + 2G(m1 + m2) v2 rel(R1 + R2) (142) A special class of collisional models are self-similar collisions. Self-similarity implies an invariance of the collisional out- come with respect to the scale of the colliding bodies. If the target mass as well as the projectile mass are enlarged by a factor of two, then only the masses of all fragments doubles without further changes in the collisional outcome. They al- low us to introduce the convenient dimensionless fragment redistribution function fm: Mred(m, m1, m2) = mfm(m1/m, m2/m) (143) We follow Tanaka et al. (1996) and employ the substitu- tion16 m1 = mx1, m2 = mx2 to simplify Eq. 137: 16 A similar approach to the solution of the coagulation equation is the Zakharov transformation, see Connaughton et al. (2004). c(cid:13) 2011 RAS, MNRAS 000, 1–42 (cid:18) (cid:19) Hybrid methods: A new composite algorithm 21 Largest Fragment γ f i l 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.02552 0.10000 0.20000 0.30000 0.40000 0.50000 0.60000 0.70000 0.80000 0.19897 0.10000 0.20000 0.30000 0.40000 0.50000 0.60000 0.70000 0.80000 0.67985 0.10000 0.20000 0.30000 0.40000 0.57612 0.67315 0.82006 0.96073 1.14050 0.10000 0.31526 0.35708 0.61362 0.83511 0.92832 0.94380 0.97572 1.77057 0.10884 0.58883 0.75974 0.87922 0.92755 0.93662 0.97107 0.97924 2.26021 0.15895 0.68891 0.87217 0.89592 0.92965 0.96089 0.96701 0.98727 3.11626 0.30954 0.83774 0.90272 0.92682 0.95943 0.95565 0.97677 0.98791 Second largest Fragment γ f i l 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.02552 0.08171 0.10770 0.06471 0.08107 0.04976 0.03982 0.03171 0.02155 0.19897 0.09174 0.09533 0.08410 0.06967 0.04930 0.04825 0.03510 0.02077 0.67985 0.07713 0.08365 0.07791 0.08387 0.07026 0.06147 0.06082 0.00847 1.14050 0.08621 0.07256 0.10331 0.09640 0.02467 0.00675 0.00783 0.00265 1.77057 0.07909 0.05549 0.06961 0.02035 0.00329 0.00719 0.00273 0.00161 2.26021 0.06693 0.02288 0.00528 0.00882 0.00584 0.00268 0.00664 0.00126 3.11626 0.06940 0.00884 0.00384 0.00488 0.00064 0.01007 0.00225 0.00162 Table 7. Data compilation of the fragmentation calculations. (cid:90)(cid:90) Fm = − n(t, mx1)n(t, mx2)m11/3 σ(x1, x2)vrelfm(x1, x2)dx1dx2 (144) A simple solution is a steady-state cascade with Fm = const. The loss of bodies of a given size is balanced by the frag- ment supply from larger bodies, hence the system maintains a steady-state ∂ ∂t n(t, m) = 0. Eq. 144 inspires the ansatz n(m) ∝ m−k, which yields k = 11/6. This is the well known equilibrium slope in self-similar collisional cascades, which was already found by Dohnanyi (1969). Strong gravitational focusing changes the exponent17 to k = 13/6. Both steady- state solutions seem to be rather artificial, as they contain an infinite amount of mass and require a steady mass influx from infinity. However, they provide an appropriate descrip- tion for the relaxed fragment tail of a size distribution, as long as the largest bodies provide a sufficient flux of new fragments. Once the largest bodies start to decay, the finite amount of mass in the system leads to an overall decay of the collisional cascade. Thus we seek for a more general solution 17 Tanaka et al. (1996) state that k < 2 is a necessary condition for a finite mass flux. However, their analysis is not valid for all possible collisional models. c(cid:13) 2011 RAS, MNRAS 000, 1–42 to Eq. 136 using the ansatz n(t, m) = a(t)n0(m): a(t) = −Ca(t)2 ∂ ∂t mn0(m) = 1 C ∂ ∂m Fm (145) (146) C is determined by fixing n0 at an arbitrary value m∗. a(t) is independent of the collision model: 1 a(t) = C ∝ n(m ∗ 1 + Ct ) (147) (148) A power law solution is n0(m) ∝ −Cm−k+1 which is only valid for C < 0 (agglomeration dominates). To examine C > 0, we perturb the already known equilibrium solution: n0(m) = N0m (149) (cid:90)(cid:90) −2k+2 + O(C 2) −k − CN1m −2k+2 −k 1 x 2 σ(x1, x2) x = (2 − k) 1 N1 vrel (fm(x1, x2) + fm(x2, x1)) dx1dx2 (150) N1 is small if the integral on the right hand side is large. This is the case for a sufficiently large impact strength. Eq. 149 has the interesting property that n(m(cid:48)) = 0 for some mass m(cid:48), given that k < 2. This mass m(cid:48) represents the largest body in the system, e. g. the largest asteroid in a fictitious asteroid belt. 22 P. Glaschke, P. Amaro-Seoane & R. Spurzem 11 SIZE–DEPENDENT STRENGTH and the scaling exponent k(cid:48) of the total mass loss: Self-similarity is an enormous help in analysing the coagu- lation equation. It releases completely the need to know any specific details of the collisional process and provides valu- able insight at the same time. But self-similarity is also a strong limitation on the underlying collisional physics. A major component of a fragmentation model is the knowledge of the impact strength as a function of size. Sim- ulations as well as asteroid families establish that it is not some fixed value, but changes with size which immediately breaks the self-similarity. Larger bodies are weaker due to an increasing number of flaws (there are no big monocrys- tals), but then gravity leads to a turnover and increases the strength. We model the size dependent strength S with a power law to examine the influence on the equilibrium solution. The velocity dispersion v and the collisional cross section σ are also modelled with power laws to account for relaxation processes: (cid:18) m (cid:18) m (cid:18) m m0 m0 (cid:19)w (cid:19)s (cid:19)α m0 v = v0 σ = σ0 S = S0 (151) (152) (153) The subscript "0" denotes values for an arbitrarily chosen scaling mass. Since smaller bodies are more abundant than larger ones, we safely assume that most collisions involve a large mass ratio. In addition, we assume w < 0, since we expect energy equipartition to some degree in most cases. These restrictions lead to the following simplifications (m1 > m2): σ(m1, m2) ≈ σ(m1) vrel ≈ v(m2)  ≈ 1 2 m2ρv2 rel m1S1 (154) (155) (156) Therefore the smaller body m2 enters only through the spe- cific energy : (cid:90)(cid:90) Fm ≈ − n(t, m1)n(t, m2)σ(m1)vrel(m2)m1 fm(m1/m, )dm1dm2 (157) We introduce new dimensionless quantities with the help of Eq. 154–156 to simplify the integral: m1 = mx1 m2 = m0 (cid:18) m1 (cid:19) 1+α 1+2w (cid:18) 2S0 (cid:19) 1 1+2w m0 ρv2 0 1 1+2w  (158) Again we assume a power law for the density n ∝ m−k and change the integration parameters to (x1, ). Applying the constant–flux condition yields the equilibrium exponent (cid:48) ≈ s − w + 1 2 + α + 2w k Mtot ∝ − S −k(cid:48) ∂ ∂t S = 2S0 ρv2 0 (160) The exponent k(cid:48) in Eq. 160 is close to unity for realistic values of the free parameters. Thus the mass loss is roughly inversely proportional to the strength of the bodies. The gen- eral formula Eq. 159 contains the special solution of O'Brien (2003), who concentrated on the parameters s = 2/3, w = 0 and a special collisional model. In fact, the derivation applies to a much wider class of collisional models that we denote as scalable collisional models. Scalable indicates that the model is self-similar except a scaling of the impactor mass. 12 PERTURBATION OF EQUILIBRIUM The derived scaling relations provide insight into the over- all properties of a collisional cascade, which is in (or close to) equilibrium. However, they do not provide information on how the equilibrium is attained or how the system re- sponds to various external perturbations. A rigorous ap- proach would be the approximate solution of the coagulation equation18, which is by no means simple since it requires a careful analysis of the collision model. Hence we turn to perturbations of the equilibrium size distribution, as it is easier to asses the quality of the derived expression for a variety of collision models. In addition, all equations are linear in the perturbation, allowing the de- tailed analysis of the solution. If the equilibrium solution n(m) = n0(m/m0)−k is per- turbed with a small deviation ∆n(m), we get to first order: (cid:90)(cid:90) ∂ 0 = ∂t Fp = − m∆n(m) + ∂ ∂m Fp(t, m) (161) ∆n(m1)n(t, m2)σ(m1, m2)vrel × (Mred(m, m1, m2) + Mred(m, m2, m1)) dm1dm2 Despite of the expansion in ∆n, Eq. 161 is still a compli- cated integro-differential equation. Thus it is not possible to obtain a solution without further information about the problem. While there is no general solution, we restrict our attention to self-similar collisional processes. In virtue of this assumption it is possible to simplify Eq. 161, as we can see in Eq. 162 and 162. In those expressions σ0 and v0 are velocity and cross section of an arbitrarily chosen scaling mass m0. F (x1) contains all information about the collisional process. If collisions do not result in extreme outcomes, like cratering or a complete destruction of the target, most of the fragment mass is contained in bodies with similar size as the parent body. Hence we expect that F (x1) peaks around x1 ≈ 1 and drops to zero as x1 gets larger (or smaller). We introduce the dimensionless relative perturbation g(m): g(m) = ∆n(m) n(m) = ∆n(m)mk n0mk 0 (164) k ≈ s + 3 + α + w(2s + α + 5) 2 + α + 2w (159) 18 Appendix C highlights a possible approach. c(cid:13) 2011 RAS, MNRAS 000, 1–42 Hybrid methods: A new composite algorithm 23 (cid:90) ∂ ∂t (cid:90) 0 = F (x1) = m∆n(m) − n0m3 0σ0v0 ∂ ∂m ∆n(t, mx1)F (x1)(mx1/m0)kdx1 m2k−3 0 1 x−k x−k 2 σ(x1, x2) σ0 vrel v0 (fm(x1, x2) + fm(x2, x1))dx2 (162) (163) a fragmentation time τfrag(u) and transform Eq. 170 back to m: 0 = τfrag = = τfrag(m) g(t, m) − m eu(2−k) (m/m0)2−k ∂ ∂t τ0 G0 τ0 G0 ∂ ∂m g(t, m) (172) (173) Eq. 172 is a modified advection equation, which conserves the total mass. It is possible to derive equations similar to Eq. 172 for any collisional model. However, the general ap- proach is less fruitful, as it lacks a robust frame of a known equilibrium solution and reliable scaling relations. Therefore we provide only the extension to scalable collisional models in Appendix B. We readily obtain the general solution: (m/m0)(2−k) G0(2 − k) t + τ0 (cid:18) (m/m0)(2−k) G0(2 − k) (cid:19) (174) ∆n(t, m) = n(m)f t + τ0 (cid:18) (cid:19) Figure 10. Scaled fragmentation kernel G(u) for a simple fragmentation model (see Eq. 200) and different scaled impact strength S. g(t, m) = f Thus the new differential equation reads: (m/m0)1−k∆g(m)− 0 = ∂ ∂t n0m2 0σ0v0 ∂ ∂m (cid:90) ∆g(t, mx1)F (x1)dx1 (165) We change to logarithmic coordinates to arrive at a convo- lution integral: The function f is determined by the initial value g(0, m) of the perturbation. As the collisional cascade evolves, the initial perturbation function is shifted as a whole to smaller masses. This evolution becomes clearer if we attach labels M (0) to the initial perturbation function and follow the time evolution of these tags. The functions M (t) are the charac- teristics 20 of the differential equation 172: (cid:105)1/(2−k) (M (0)/m0)(2−k) − t/τ0G0(2 − k) (cid:104) u = ln(m/m0) u1 = ln(x1) (166) M (t) = m0 Furthermore we define a collisional timescale τ0 τ0 = (n0m0σ0v0) −1 (167) (175) The meaning of the fragmentation time τfrag becomes clear by the relation to obtain a more concise expression. The transformed equa- tion is: 0 = ∂ ∂t g(t, u)eu(2−k) − 1 τ0 ∂ ∂u G(u) =F (eu)eu g(t, u + u1)G(u1)du1 (168) (169) (cid:90) If g(u) is varying on a scale larger than the width of the kernel G(u) (compare Fig. 10), it is justified to expand g(u) under the integral. We retain the first two moments of G(u): ∂ ∂t (cid:90) 0 = Gk = g(t, u)eu(2−k) − G0 τ0 ∂ ∂u g(t, u) − G1 τ0 ∂2 ∂u2 g(t, u) ukG(u)du (170) (171) The first order moment G1, which introduces a diffusive term, is omitted in the following for clarity19. We introduce = −τfrag M M (176) which is the time until a body has lost a significant fraction of its mass due to destructive collisions. A comparison of the perturbation equation 172 with the scaling relations from the previous section gives the scaling of the zeroth order moment G0 with respect to the impact strength: (cid:48) G0 = G 0 −k(cid:48) S (177) G(cid:48) 0 should only depend on the fragmentation model (i. e. fragment size distribution as a function of the largest frag- ment fl) within the limits of this approximation. Fig. 10 shows that the scaling with the impact strength works quite well, except slight variations which are small compared to the covered range of impact strengths. Likewise, it is pos- sible to restate the total equilibrium flux Feq in terms of G(cid:48) 0: Feq(m) ≈ − G(cid:48) 0 2 n(m)2σ(m)m3vrel S −k(cid:48) (178) 19 The study of wave-like structures in the size distribution (Bagatin et al. 1994,see e.g.) requires even the second order mo- ment G2. 20 In general, characteristics of a partial differential equation are paths along which the solution is constant. c(cid:13) 2011 RAS, MNRAS 000, 1–42 0.001 0.01 0.1 1 10 100-15-10-5 0 5 10 15 20 25G(u)S~k'uS~=1000S~=10S~=0.1S~=0.001 24 P. Glaschke, P. Amaro-Seoane & R. Spurzem The fragmentation timescale τfrag(m) allows a more intuitive expression: Feq(m) ≈ − 1 2 n(m)m2 τfrag(m) (179) Our simple collisional model (see Fig. 10 and Eq. 200) refers to: Feq(m) = −(1 . . . 30) × n(m)2σ(m)m3vrel S −k(cid:48) (180) 13 MIGRATION AND COLLISIONS The local perturbation analysis is only applicable to a plan- etesimal disc, if the migration velocity of the planetesimals is negligible small. This assures that collisional cascades at different radial distances do not couple to each other, so that the whole disc is composed of many local cascades. While this assumption is justified for larger bodies, migration is strongly influencing bodies below 1 km in size. Hence we extend our analysis to examine the influence of migration on the (no longer) local collisional processes. We assume that the collisional evolution of the system leads to an equilibrium planetesimal distribution everywhere in the disc: Σ0(r, m) = Σr,0(r)C0(m) (181) Σr(r) is the total surface density at a given distance r, while C0(m) is the universal equilibrium distribution. Though the planetesimal distribution at larger sizes is likely different at different locations in the disc, we only demand a universal function at smaller sizes, where migration is important. The power law exponent k depends on the details of the invoked physics, but numerical simulations show that k ≈ 2 is a fiducial value. Eq. 181 does not yet include migration effects. If we include migration, the surface density is modified to Σ(r, m) = g(r, m)Σ0(r, m) (182) where the dimensionless function g contains the changes due to migration. The collisional evolution is governed by the continuity equation with an additional collisional term ∂Σ(r, m) ∂t − 1 r ∂ ∂r (v(r, m)rΣ(r, m)) = Σcoll (183) where v(r, m) is the migration velocity (see Eq. 77), defined such that positive v imply an inward migration. We express the collisional term with the help of Eq. 172 and seek for a steady-state solution Σ = 0: 1 τfrag(m, r) ∂g ∂m mΣr,0(r) + 1 r ∂ ∂r (gvrΣr,0(r)) = 0 (184) τfrag(m, r) is the fragmentation timescale of a mass m at a distance r. Since the surface density Σ and the various contributions to the drag force are well described by a power law (with respect to radius), Eq. 184 further simplifies to: 1 τfrag(m, r) ∂g ∂m m + ∂g ∂r v − b r gv = 0 (185) b is a combination of the various invoked power law exponents. As the surface density Σ and the gas density drop with increasing radius in any realistic disc model, it is safe to assume b > 0. We choose a self-similar ansatz for g: g(r, m) = g(ζ) , ζ = mgm(r) (186) Figure 11. Cut-off function g according to Eq. 192. The mass exponent is km = 1/3, while the mass influx exponent is b = 1.75 according to the minimum solar nebula. The new differential equation is 1 τfrag(m, r) dg dζ mgm(r) + m gv = 0 (187) dg dζ dgm dr v − b r which is equivalent to the more concise expression: (cid:18) r (cid:19) d ln(g) d ln(ζ) + d ln(gm) d ln(r) vτfrag = b (188) We assume a power-law dependence for the timescale ratio τmig/τfrag: r vτfrag = τmig τfrag = (m/m0)km (r/r0)kr (189) The cut-off mass m0 at a distance r0 has a timescale ratio τmig/τfrag = 1, which defines a proper lower cut-off within this context. Hence the solution is: (cid:18) (cid:19) ζ km + d ln(g) d ln(ζ) (r/r0)kr /km kr km b = gm(r) = (cid:18) (cid:19)−b/kr m0 kr kmζ km g(ζ) = 1 + (190) (191) (192) Though the analytical solution Eq. 192 provides a complete description of the lower cut-off of the size distribution, it is more appropriate within the frame of this discussion to translate the equilibrium solution to an equilibrium mass loss due to migration: Σmig(r, m) = − bv r Σ + kr/km ζ km + kr/km Σ bv r bΣ = − τmig + kr/kmτfrag (193) An inspection of the timescale ratio shows that the mass exponent km should be positive, whereas simple estimations of kr on the basis of the minimum mass solar nebula are somewhat inconclusive. The value of kr is so close to zero that any change in the assumed equilibrium slope or the impact strength scaling gives easily both positive and neg- ative values. Moreover, Eq. 193 requires a globally relaxed planetesimal disc, but the huge spread in the various in- volved timescales at different radii inhibits any significant relaxation in the early stages. c(cid:13) 2011 RAS, MNRAS 000, 1–42 0 0.2 0.4 0.6 0.8 1 1 10 100 1000 10000 100000 1e+06gζkr=-1kr=0kr=1 However, it is possible to gain valuable information from the two limiting cases kr > 0 and kr < 0. Both values of kr give the proper limit g → 1 at large masses, where the migration timescale is much larger than the fragmentation timescale and we recover the steady-state collisional cascade. A positive exponent kr reduces the effective mass loss due to migration, as fragments from the outer part of the disc replenish the local mass loss. Hence the fragmentation timescale controls the net loss of smaller planetesimals. In contrast, a negative exponent kr leads to a pronounced cut- off in the size distribution, since only larger planetesimals are replenished through inward migration. Though the mass loss rate is singular at some mass m(cid:48), this sharp cut-off is an artifact due to the perturbation approximation. Our analysis is subjected to several restrictions. We ap- plied the perturbation equation to values of g that exceed the limit for a safe application (i. e. g (cid:54)≈ 1) of the perturba- tion expansion. Furthermore, the steady-state solution re- quires a global relaxation of the collisional processes, which is practically never obtained during the disc evolution. De- spite of these restrictions, we gained insight on a more qual- itative level. Numerical calculations indicate that the per- turbation approximation is inappropriate close to the lower cut-off of the size distribution. However, a comparison of dif- ferent exponents kr (see Fig. 11) attributes only a minor role to the replenishment of fragments due to inward migration. Only unrealistic small slopes b of the migrational mass influx would strengthen the importance of this process. Though temporally non-equilibrium phenomena are not ruled out by the previous derivation, their study would require the global simulation of the system. 14 COAGULATION While most coagulation kernels are only restricted to a lim- ited analytical analysis (e. g. scaling relations), there exist some special kernels that allow the closed solution of the coagulation equation. All rely on the assumption of perfect mergers, which allows the reformulation of the general equa- tion 136 to ∂n(m, t) ∂t = 1 2 K(m − m (cid:48) (cid:48) )n(m − m (cid:48) (cid:48) , t)n(m , m (cid:48) , t)dm (cid:90) m 0 (cid:90) ∞ − n(m, t) K(m, m (cid:48) (cid:48) )n(m (cid:48) , t)dm 0 where K is the coagulation kernel. One of these particular kernels was introduced by Safronov (1969): K(m1, m2) = A1(m1 + m2) (194) This coagulation kernel implies perfect mergers, where the coalescence rate of two particles m1 and m2 is assumed to be proportional to the sum of their masses. It seems that this is an artificial choice, devised to allow an analytic solution. However, the Safronov cross section provides an intermedi- ate case between a geometric cross section (σ ∝ m2/3) and strong gravitational focusing (σ ∝ m4/3). A special solution to the initial condition Hybrid methods: A new composite algorithm 25 ρ k 2,700 kg/m3 1/6 Model Gaussian Scatter fKE K 0.1 1.24 Table 8. Main parameters of the collisional model. is the function (Ohtsuki et al. 1990,see e.g.) n(m, τ ) = √ n0g 1 − g m exp(−(2 − g)m/m0)I1(2m/m0 g = exp(−τ ) (cid:90) ∞ τ = A1ρt (cid:112)1 − g) (196) ρ = mn(m)dm = n0m0 (197) 0 where τ is the dimensionless time and I1 is a modified Bessel function of the first kind. 15 MODELS FOR MRED Though we already obtained insight into the nature of col- lisional cascades without a detailed specification of the co- agulation kernel, any detailed study of a collisional system requires the specification of a realistic collisional model. First, we restate the well-known perfect accretion model. While it is a gross oversimplification for collisions among kilometre–sized planetesimals, its simplicity allows a reliable code testing and eases the comparison with other works: Mred(m, m1, m2) = −m1Θ(m − m1) − m2Θ(m − m2)+ (m1 + m2)Θ(m − m1 − m2) (198) Although our fragmentation model (see section 9) pro- vides a very detailed description of the outcome of a colli- sion, we abandon most of the details for the following rea- sons. The computational effort of the numerical solution of the coagulation equation scales with the third power of the number of mass bins. Hence we chose a mass grid whose res- olution is by far smaller than the information provided by the detailed collisional model. As a mismatch of the mass resolution could produce undesired artifacts, a lower resolu- tion of the collisional model is needed for consistence. Thus only the largest fragment fl(f i l , γ) and the second fragment f (2) l , γ)(which contains information on reaccumulation) l enter the fragment size distribution: (f i Mred(xM ) M = 1 − fl (1 − fl − f (2) l )(x/f (2) l )fl if x (cid:62) fl if fl > x (cid:62) f (2) otherwise l (199)  1 n0 m0 c(cid:13) 2011 RAS, MNRAS 000, 1–42 n(m, 0) = exp (−m/m0) (195) Mred(xM )/M = Both values fl and f (2) are interpolated from table 7, where the initial fragment size f i is calculated from the dimen- l sionless impact energy . We used a reduced fragmentation model for test purposes: l (cid:26) 1if x (cid:62) fl (1 − fl)(x/fl)fl otherwise (200) 26 P. Glaschke, P. Amaro-Seoane & R. Spurzem Table 8 summarises the most important model parameters. unique and depends well on the problem at hand (Larson 1970,compare e.g.). 16 STATISTICAL MODEL The direct approach to the integration of an N –body system is, in principle, possible for any particle number. While this procedure becomes computationally too expensive for very large particle numbers, a by far more efficient approach is applicable in this regime. Instead of tracking all particle or- bits, a distribution function f (also phase-space density ), which gives the probability to find a particle at a position x with a velocity v, contains the state of the system: dp = f (x, v)d3xd3v (201) As long as only dynamical interactions are taken into ac- count, the number of all particles (e. g. stars, planetesimals) is conserved. The continuity equation reads: + v · ∇f − ∇Φ · ∂f ∂v ∂f ∂t (202) 0 = This is the collisionless Boltzmann equation. Collisions lead to an additional term = ∂f ∂t + v · ∇f − ∇Φ · ∂f ∂v (203) (cid:18) ∂f (cid:19) ∂t coll which will be discussed later. f is a function of six variables, so an exact solution is usually very complicated or even im- possible. However, it is possible to gain valuable insight into the problem by taking the moments of the distribution func- tion: (cid:104)xn j (cid:105) = i vm f (x, v)xn i vm j d3xd3v n, m > 0 (204) (cid:90) The spatial density (particles per volume) is related to dis- tribution function: (cid:90) ν(x) = f (x, v)d3v (205) Integration of Eq. 203 over all velocities yields the corre- sponding continuity equation: (cid:18) ∂ν (cid:19) ∂t coll ∂ν ∂t + ∂ν ¯vi ∂xi = The first order moment with respect to velocity gives the time evolution of the mean velocity ¯v − ∂(νσ2 ij) ∂xi (cid:18) ∂¯vj = −ν ∂Φ ∂xj ∂¯vj ∂xi ∂¯vj ∂t (cid:19) + ν ¯vi (207) + ν ∂t ν coll (cid:90) (206) (208) f (x, v)vid3v 1 ν ¯vi = ij = vivj − ¯vi¯vj σ2 where σij is the anisotropic velocity dispersion and the con- tinuity equation was used to arrive at a more concise for- mulation. Equations 206 and 207 are the Jeans equations. While the structure of the moment equations is already fa- miliar from hydrodynamics, they do not provide a closed set of differential equations, since each differential equation of a given moment is related to (yet unknown) higher order moments. Hence any finite set of momenta needs a closure relation – additional constraints that relate the highest order moments to known quantities. The choice of this relation is a key element in the validity of the equations, but it is not Owing to the geometry of a planetesimal disc, it is useful to express the Boltzmann equation in cylindrical coordinates (cid:32) (cid:33) ∂f ∂z + v2 φ r − ∂Φ ∂r ∂f ∂t + vr vrvφ r ∂f ∂vφ + vz ∂f ∂r − ∂Φ ∂z ∂f ∂vz = 0 − ∂f ∂vr (209) where all derivatives with respect to φ have been dropped due to the assumed axisymmetry of the disc. 16.1 Distribution Function Any statistical description of a planetesimal disc requires the knowledge of the distribution function. Since the full problem including collisions, encounters and gas drag has no analytic solution, a collisionless planetesimal disc (i. e. no perturbations) is a natural basis for further investigations. The distribution function that describes such a simplified system is a solution of the Boltzmann equation. A special solution to Eq. 209 is a thin homogenous planetesimal disc f (z, v) = ΩΣ 2π2TrTzm exp − v2 r + 4v2 φ 2Tr − v2 z + Ω2z2 2Tz (210) provided that the radial velocity dispersion Tr and the ver- tical dispersion Tz are small compared the mean orbital ve- locity vK . The azimuthal velocity dispersion Tφ is locked to Tr by the local epicyclic frequency κ in a central potential, where the ratio 1 : 4 is a special solution of (Binney 1994,see e.g) κ2Tr = 4Ω2Tφ (211) All velocities vr, vφ and vz refer to the local Keplerian ve- locity. The normalisation is the same as in Stewart (2000): (cid:32) (cid:33) (cid:90) d3vdzf (z, v) = Σ m (212) A planetesimal disc is a slowly evolving system compared to the orbital time, hence it is reasonable to use Eq. 210 as a general solution of the perturbed problem. Σ, Tz and Tr are now functions of time and of the radial distance to the star. All information on the system is contained in these three momenta of the distribution function, where higher order moments can be deduced from Eq. 210. Thus the functional form of the distribution function represents an implicit clo- sure relation. The validity of this approximation can be further as- sessed by a closer examination of the Boltzmann equation. We summarise all perturbations in an evolution timescale Tevol and reduce the radial structure to some typical length scale ∆r to estimate the deviation from the functional form Eq. 210. A comparison with Eq. 209 shows that the differ- ence is small if the migration timescale and the evolution timescale are large compared to the orbital time T0: T0 (cid:28) ∆r/(cid:104)vr(cid:105) T0 (cid:28) Tevol (213) (214) c(cid:13) 2011 RAS, MNRAS 000, 1–42 (cid:114) h = An order–of–magnitude estimate of the evolution time sup- ports condition 213 and 214. Furthermore, numerical calcu- lations confirm that the velocity distribution stays triaxial Gaussian (see Ida 1992). The distribution function is equivalent to an isothermal vertical density structure with scale height h: Tz Ω2 (cid:18) (cid:19) (215) ρ(z) = ρ0 exp (216) Thus the central density ρ0 and the mean density (cid:104)ρ(cid:105) are related to the surface density in a simple way: − z2 2h2 ρ0 = (cid:104)ρ(cid:105) = Σ√ 2πh ρ0√ 2 (217) The triaxial Gaussian velocity distribution is equivalent to a Rayleigh distribution of the orbital elements e and i21: (cid:18) (cid:19) dn(e2, i2) = (cid:104)e2(cid:105) = 1 (cid:104)e2(cid:105)(cid:104)i2(cid:105) exp 2Tr (Ωr0)2 (cid:104)e2(cid:105) − i2 − e2 (cid:104)i2(cid:105) de2di2 (cid:104)i2(cid:105) = 2Tz (Ωr0)2 (218) Planetesimal encounters couple the time evolution of eccen- tricity and inclination, so that the ratio i2/e2 tends to an equilibrium value after a few relaxation times. It is close to 1/4 in a Keplerian potential, but the precise value also depends on the potential itself (Ida et al. 1993). 16.2 Dynamical Friction Planetesimal–planetesimal scatterings change the velocity distribution through two different processes. Firstly, it is unlikely that two planetesimals scatter each other on circu- lar orbits. Thus we expect a steady increase of the velocity dispersion due to this viscous stirring. Secondly, encounters between unequal masses lead successively to energy equipar- tition, slowing down the larger bodies through dynamical friction. The later mechanism is not related to the disc ge- ometry at all, but operates in any multi-mass system. A special case is the systematic deceleration of a massive body M in a homogeneous sea of lighter particles m with density n0, which is given by the Chandrasekhar dynamical friction formula (Chandrasekhar 1942) 4π ln ΛG2(M + m)n0m v3 M dvM dt = −vM vM√ 2σv X = (cid:18) erf(X) − 2X√ π e −X2(cid:19) (219) where σv is the velocity dispersion of the lighter particles. The Coulomb logarithm Λ arises from an integration over all impact parameters smaller than an upper limit lmax and is given by Λ ≈ σ2 vlmax G(m + M ) (220) 21 Eq. 10–12 provide the coordinate transformation. c(cid:13) 2011 RAS, MNRAS 000, 1–42 Hybrid methods: A new composite algorithm 27 Although encounters in the gravitational field of the sun de- viate from pure two-body scatterings, it is safe to neglect the presence of the sun if the encounter velocity is large compared to the Hill velocity 22 ΩRHill. Thus the classical dynamical friction formula is also applicable to planetesimal encounters in the high velocity regime, though a generalisa- tion to triaxial velocity distributions σi is necessary (Binney 1977,see e.g.): dvM,i dt Bi = exp = −vM,i (cid:90) ∞ (cid:112)(σ2 0 √ 2πG2 ln(Λ)(M + m)n0mBi (cid:18) − 1 2 (cid:19) (cid:88) v2 j σ2 j + u du 1 + u)(σ2 2 + u)(σ2 3 + u)(σ2 i + u) (221) (222) An additional complication is the choice of lmax (i. e. the choice of the Coulomb logarithm). There are several scale lengths, which could determine the largest impact parame- ter lmax: The scale height of the planetesimal disc, the radial excursion due to the excentric motion of the planetesimals and the Hill radius of the planetesimals. As it is not possible to derive a unique expression for lmax from first principles, a proper formula is often fitted to N –body calculations (com- pare Eq. 235). The velocity dispersion of a planetesimal disc is triaxial with Tφ/Tr = 1/4 and Tz/Tr ≈ 1/4. We take these values and expand Eq. 221 for small velocities vM : dvM,r dt dvM,φ dt dvM,z dt ≈ −1.389 vM,r ≈ −3.306 vM,φ ≈ −3.306 vM,z √ 2πG2 ln(Λ)(M + m)n0m √ 2πG2 ln(Λ)(M + m)n0m T 3/2 r √ 2πG2 ln(Λ)(M + m)n0m T 3/2 r T 3/2 r (223) The derived expressions provide a compact tool to analyse dynamical friction in disc systems. However, the involved approximations are too severe compared to the needs of an accurate description. While these concise expressions are valuable for basic estimations, the following sections derive viscous stirring and dynamical friction formulae for a plan- etesimal system in a rigorous way. 16.3 High Speed Encounters We return to the Boltzmann equation as a starting point for the derivation of the scattering coefficients: = ∂f ∂t + v · ∇f − ∇Φ · ∂f ∂v (224) (cid:18) ∂f (cid:19) ∂t coll In virtue of the ansatz for the distribution function (see Eq. 210), it is sufficient to derive the time derivative of the second order velocity moments Tr and Tz. Since the distri- bution function is time independent in the absence of en- counters, only the collisional term contributes to the time derivative of the velocity dispersions Tk (k ∈ (r, z, φ) in the 22 Whenever relative velocities are classified as "high" or "low" in the following sections, a comparison with the Hill velocity is implied. 28 P. Glaschke, P. Amaro-Seoane & R. Spurzem following): (cid:90) dρTk dt = (cid:18) ∂f (cid:19) ∂t coll d3vmv2 k (225) u–integration: d(cid:104)ρv2 r(cid:105) dt (cid:90) (cid:90) (cid:90) (cid:90) (cid:90) d3w r − mTr)u2 r r )u3 + ∗2 = 2πG2mm (cid:20) 2A(m∗T ∗ (cid:34) = 2πG2mm 2A(m∗T ∗ m∗(Tr + T ∗ ∗2 (cid:90) m∗(Tr + T ∗ r )u3 + ∗2 = 2πG2mm (cid:20) 2A(m∗T ∗ d3w z − mTz)u2 z m∗(Tz + T ∗ z )u3 + ∗× d3uf f B(u2 − 3u2 r) (cid:21) u3 ∗× u3 ∗× u3 (cid:35) (cid:21) d3uf f B(u2 − 3u2 z) d3w r − mTr)u2 φ d3uf f B(u2 − 3u2 φ) d(cid:104)ρv2 φ(cid:105) dt d(cid:104)ρv2 z(cid:105) dt The collisional term invokes the averaging over many differ- ent scattering trajectories and is, given that the underlying encounter model is analytically solvable, still too complex to derive an exact expression. If most of the encounters are weak – a realistic assumption in a planetesimal disc – it is possible to expand the collisional contribution in terms of the velocity change ∆vi. This is the Fokker-Planck approx- imation (see e.g. Binney 1994)) (cid:18) ∂f (cid:19) ∂t coll = −(cid:88) (cid:88) i 1 2 ∂2 ∂vi∂vj i,j ∂ ∂vi [f D(∆vi)]+ [D(∆vi, ∆vj)] (226) where the diffusion coefficients D contain all information on the underlying scattering process. Next we consider two in- teracting planetesimal populations m, m∗ with distribution functions (cid:32) − v2 r + 4v2 (cid:32) φ 2Tr − v2 − v2 z + Ω2z2 2Tz (cid:33) (cid:33) r + 4v2 φ 2T ∗ r − v2 z + Ω2z2 2T ∗ z (227) f = ∗ f = ΩΣ 2π2TrTzm exp ΩΣ∗ r T ∗ 2π2T ∗ z m∗ exp to evaluate the terms in equation 226. We follow Stewart (2000) except some minor changes in the notation. The col- lisional term requires an averaging over the velocities of the two interacting planetesimals m and m∗: (cid:90) (cid:90) + d(cid:104)ρv2 k(cid:105) dt (cid:20) ∗2 = 2πG2mm d3v − 2A(m + m∗)ukvk uk = vk − v m∗u3 ∗ k ∗ ∗× f f d3v Bu2 + (2C − B)3u2 k) (cid:21) u3 A = ln(Λ2 + 1) C = Λ2 Λ2 + 1 B = A − C (228) A coordinate transformation to the relative velocity u and the modified centre-of-mass velocity w  Vk + Vk + mv + m∗v∗ m + m∗ wk = V = r − mTr)uk (m∗T ∗ (m + m∗)(Tr + T ∗ r ) z − mTz)uk (m∗T ∗ (m + m∗)(Tz + T ∗ z ) for k ∈ {r, φ} for k = z (229) All integrals are solvable and give the result r,φ(cid:105) d(cid:104)v2 dt G2ρ∗ = 2 √ 2(Tr + T ∗ r )3/2 ∗ ∗ [B(T r + Tr)m Jr,φ(β) + 2A(T G2ρ∗ √ 2(Tr + T ∗ r )1/2(Tz + T ∗ z ) ∗ ∗ ∗ Jz(β) + 2A(T z + Tz)m z m Tz + T ∗ Tr + T ∗ [B(T = 2 z r β2 := d(cid:104)v2 z(cid:105) dt ∗ r m ∗ − mTr)Hr,φ(β)] ∗ − mTz)Hz(β)] (230) where six auxiliary functions are introduced to arrive at a more compact notation: b =(cid:112)1 − (1 − β2)x2 (cid:112) a = 4 − 3x2 √ (cid:90) 1 (cid:90) 1 (cid:90) 1 0 0 π π Hr := 8 x2 dx ab √ 1 − x2 Hφ := 8 a(βa + b) √ β(1 − x2) Hz := 8 b(βa + b) Jr := −2Hr + Hφ + Hz Jφ := Hr − 2Hφ + Hz Jz := Hr + Hφ − 2Hz π 0 dx dx (231) Since these are non-trivial functions, we apply a standard Chebyshev approximation for β ∈ [0, 1]: f (x) ≈ 5(cid:88) k=0 ckTk(x) − 1 2 c0 (232) 4 = Table 9 summarises the Chebyshev coefficients. A final z– averaging yields the expressions: r,φ(cid:105) d(cid:104)v2 dt G2ΩΣ∗ √ π(Tr + T ∗ r )3/2(Tz + T ∗ ∗ ∗ [B(T r + Tr)m Jr,φ(β) + 2A(T G2ΩΣ∗ √ π(Tr + T ∗ r )1/2(Tz + T ∗ ∗ ∗ Jz(β) + 2A(T z + Tz)m × (233) z )1/2 ∗ − mTr)Hr,φ(β)] ∗ r m × ∗ − mTz)Hz(β)] z )3/2 ∗ z m d(cid:104)v2 z(cid:105) dt (234) [B(T = 4 further simplifies the double integral. Thus the integration separates in a simple integral over w and a more demanding The determination of a proper Coulomb logarithm Λ leaves room for further optimisation. A careful comparison with N –body models gives rise to the empirical choice (Ohtsuki c(cid:13) 2011 RAS, MNRAS 000, 1–42 Hybrid methods: A new composite algorithm 29 f c0 c1 c2 c3 c4 c5 ∆ Jr Jφ Jz Hr Hφ Hz -10.34660733 1.81674741 8.52985992 11.00434580 6.94989422 4.71219005 4.69990443 2.95397208 -7.65387651 -2.64707927 -2.06510182 1.47084771 -1.25533220 -1.18724874 2.44258094 0.60969641 0.58700192 -0.62294130 0.30288875 0.37775788 -0.68064662 -0.13815856 -0.16311494 0.18968657 -0.07040537 -0.11070339 0.18110876 0.03112047 0.04455314 -0.05271757 0.01540098 0.02922947 -0.04463045 -0.00669979 -0.01130929 0.01331068 0.006 0.015 0.022 0.0025 0.0058 0.0066 Table 9. Chebyshev coefficients of the auxiliary functions Jk and Hk. et al. 2002): Λ = e = 1 √ 12 2Tr ΩRHill ((cid:104)e2(cid:105) + (cid:104)i2(cid:105))(cid:104)i2(cid:105)1/2 √ 2Tz ΩRHill i = (235) (236) Ohtsuki et al. 2002 also report a further improvement by setting B ≡ A. 16.4 Low Speed Encounters to the initial orbital plane preventing any excitation of in- clinations. The respective expressions for the dynamical friction rates are: dTr dt dTz dt = = C3 := 6(m + m∗)(Tr + T ∗ r ) Gr0ΩhΣ∗ Gr0ΩhΣ∗ 6(m + m∗)(Tz + T ∗ z ) ln(10Λ2 + 1) 10Λ2 10C3(cid:104)e2(cid:105)(T ∗ r m ∗ − Trm) 10C3(cid:104)i2(cid:105)(T ∗ z m ∗ − Tzm) (238) Encounters in the low velocity regime exhibit a wealth of dif- ferent orbits, as the solar gravity field perturbs the two-body scattering. Only a small subset of the trajectories represents simple, regular orbits like Tadpole or Horseshoe orbits23. Hence an examination of this velocity regime is done best with a numerical study of the parameter space by integrat- ing the equations of motions numerically (see Eq. 16). As the stirring rates are only valid in the low velocity regime, Ohtsuki et al. (2002) introduced special interpolation coef- ficients Ci. These coefficients tend to unity for very small velocity dispersions, and drop to zero in the high veloc- ity regime. Thus the interpolation formulae are properly "switched off" in the high velocity regime, so they do not interfere with the known high velocity stirring rates. Ohtsuki et al. (2002) integrated a large set of planetesi- mal encounters and extracted fitting formulae that cover the low velocity regime. Their expressions for viscous stirring are: dTr dt dTz dt = = Gr0ΩhΣ∗ 6(m + m∗) Gr0ΩhΣ∗ 6(m + m∗) ∗ 73C1m ∗(cid:18) 4(cid:104)i2(cid:105) + 0.2((cid:104)e2(cid:105))3/2(cid:113) C2m (cid:104)i2(cid:105) (cid:19) e = e/h i = i/h C1 := C2 := ln(10Λ2/e2 + 1) 10Λ2/e2 √ √ e2 + 1) e2 ln(10Λ2 10Λ2 16.5 Distant Encounters All formulae include only the stirring rates due to close en- counters, but non-crossing orbits also contribute to the over- all change of the velocity distribution. As these distant en- counters lead to small changes of the orbital elements, the problem is accessible to perturbation theory; see Hasegawa (1990) for a detailed treatment. Stewart (2000) integrated the perturbation solution over all impact parameters to de- rive the collective effect of all distant encounters: (237) d(cid:104)e2(cid:105) dt = Ωm∗Σ∗r2 (m + m∗)2 (cid:104)PVS,dist(cid:105) 0 (cid:104)PVS,dist(cid:105) = 7.6 The stirring rate of the radial velocity dispersion approaches a finite value for very low velocity dispersions, while the stir- ring rate for the vertical velocity dispersion drops to zero as the velocity dispersion decreases. This different behaviour of the two limits is due to the encounter geometry: If two plan- etesimals have zero inclination, they may still excite higher eccentricities during an encounter, but they remain confined EXINT α α(m + m∗)2 (cid:16) M 2 c h2 α ((cid:104)e2(cid:105)+(cid:104)e∗ 2(cid:105)) (cid:104)e2(cid:105) + (cid:104)e∗2(cid:105) − (cid:104)i2(cid:105) − (cid:104)i∗2(cid:105) (cid:17) − EXINT (cid:16) (cid:114) m + m∗ h = 3 (cid:17) h2 ((cid:104)i2(cid:105)+(cid:104)i∗ 2(cid:105)) EXINT(x) := exp(x)Γ(0, x) α ≈ 1 3Mc (239) 23 The most famous example of such a regular orbit are the two saturnian moons Janus and Epimetheus which share nearly the same orbit. α accounts for the uncertainty in the smallest impact param- eter that is regarded as a distant encounter. While distant encounters are already included in the interpolation formula c(cid:13) 2011 RAS, MNRAS 000, 1–42 30 P. Glaschke, P. Amaro-Seoane & R. Spurzem of the low–velocity regime, we use the modified expression: (cid:18) dTr (cid:19) dt dist (cid:18) d(cid:104)e2(cid:105) (cid:19) = = (1 − C1) 1 (Ωr0)2 2 dt GMcr0Ωm∗Σ∗ 2(m + m∗)2 (cid:104)PVS,dist(cid:105)(1 − C1) dist (240) Stewart (2000) omitted the change in the inclination, as it is small due to the encounter geometry. Nevertheless we de- rived the integrated stirring rate for completeness, which we give in 241 A close inspection of the integrated perturbation shows that the above formula is roughly a factor (cid:104)i2(cid:105) +(cid:104)i∗2(cid:105) smaller than the corresponding changes in the eccentricity. 16.6 Gas Damping The presence of a gaseous disc damps the velocity dispersion of the planetesimals and introduces a slow inward migration. Adachi et al. (1976) used the drag law Eq. 75 to approxi- mate24 the average change of the orbital elements: vK − vg 2m τ0 = πCDρgR2vK ηg = (cid:18) (cid:18) d dt e2 ≈ − 2e2 τ0 i2 ≈ − 2i2 τ0 a ≈ − 2a τ0 d dt d dt vK (cid:19) (cid:19) 3 2 1 2 0.77 e + 0.64 i + ηg 0.39 e + 0.43 i + ηg ηg (0.97 e + 0.64 i + ηg) (242) (243) ηg is the dimensionless velocity lag of the sub–Keplerian rotating gaseous disc. 16.7 Unified Expressions All expressions for the different velocity regimes are con- structed such that a smooth transition between the different regimes is assured. Thus, a simple addition of all contribu- tions yields already the unified expressions (cid:18) dTr (cid:18) dTz dt (cid:19) (cid:19) dTr dt = dTz dt = (cid:18) dTr (cid:18) dTz dt (cid:19) (cid:19) + + (cid:18) dTr (cid:18) dTz dt (cid:19) (cid:19) + + high low gas dt high dt low dt gas (cid:18) dTr (cid:18) dTz dt (cid:19) (cid:19) + + dist (244) dt dist (245) which cover the full range of relative velocities. Although only two populations m and m∗ were assumed, Eq. 244 and Eq. 245 are readily generalised to a multi-mass system by adding a summation over all masses. 16.8 Inhomogeneous Disc The preceding derivations assumed a homogeneous disc, which simplified the calculation, since the integration over all impact parameters needed no special precaution. A more sophisticated consequence is that the spatial density and the density in semimajor axis space are equal: Σ(r) = Σ(a) = Σ0 (246) Density inhomogeneities break this simple relation, as par- ticles at the same radial distance could have different semi- major axes, and particles with the same semimajor axis are located at different positions at a given time. While both representations are equivalent (i. e. describe the same sys- tem in different ways), we chose the density in semimajor axis space as the primary density25. The spatial density is derived as: 1(cid:112)2πa2(cid:104)e2(a)(cid:105) exp (cid:18) − (a − r)2 2a2(cid:104)e2(a)(cid:105) (cid:19) (cid:90) Σ(r) = Σ(a)da (247) Likewise, Tr and Tz are also functions of the semimajor axis. Furthermore, an inhomogeneous surface density invali- dates the averaging over all impact parameters. Planetesimal encounter are most efficient for impact parameters smaller than a few Hill radii, so the derivation is still valid if the sur- face density is roughly constant on that length scale. How- ever, a planetesimal that is large enough will "feel" the spa- tial inhomogeneities or even generates density fluctuations. Hence it is essential to extend the validity of the averaged expressions to inhomogeneous systems. We use the averaged expressions (cid:28) dTr,z (cid:29) (cid:90) ∞ = Σ(a) dt d Tr,z(b) −∞ dt db (248) as a starting point (d Tr,z/dt excludes the surface density, as opposed to the averaged expressions). The (yet unknown) scattering contribution d Tr,z/dt as a function of the impact parameter b is our starting point for a general expression for a varying surface density: dT (a0)r,z = Σ(a0 + b) d Tr,z(b) dt db (249) (cid:90) ∞ (cid:28) dTr,z −∞ (cid:29)−1 We restate Eq. 249 in terms of a weight function w(b): dT (a0)r,z dt = dt Σ(a0) d Tr,z(b) w(b) = Σ(a0) dt dt Σ(a0 + b)w(b)db (250) (251) The numerical solution of the Hill problem (see Eq. 16) gives some insight into how the weight function w(b) changes with the impact parameter. Fig. 12 illustrates the change in e2 of the relative motion during an encounter of two plan- etesimals that were initially on circular orbits. While the details depend on the initial inclination and eccentricity as well as on the selected orbital element, all result share some basic features. Small (compared to the Hill radius) impact parameters allow for a horseshoe orbit and the change in the orbital elements is small except a change in the semima- jor axis. Intermediate impact parameters which lead to close encounters provide the strongest perturbation, but they are also more susceptible to complicated dynamics (compare the (cid:90) ∞ (cid:29) 1 −∞ dt (cid:28) dTr,z 24 A formal expansion at e = 0, i = 0, ηg = 0 is not possible, since the drag law involves the modulus of the relative velocity. Kary et al. (1993) corrected a missing factor 3/2 in Eq. 242. 25 We denote Σ(a) also as "surface density" and refer to a as a radial coordinate. However, all formulae are precise in discrimi- nating both representations in r and a. c(cid:13) 2011 RAS, MNRAS 000, 1–42 d(cid:104)i2(cid:105) dt Ωm∗Σ∗r2 (m + m∗)2 0 = (cid:104)QVS,dist(cid:105)(cid:104)QVS,dist(cid:105) = 0.4 α2(m + m∗)2 M 2 c × − ((cid:104)i2(cid:105) + (cid:104)i∗2(cid:105)) EXINT Hybrid methods: A new composite algorithm 31 1 (cid:34) (cid:104)e2(cid:105) + (cid:104)e∗2(cid:105) − (cid:104)i2(cid:105) − (cid:104)i∗2(cid:105) × (cid:17) − EXINT (cid:16) (cid:16) α ((cid:104)e2(cid:105)+(cid:104)e∗ 2(cid:105)) (cid:104)e2(cid:105) + (cid:104)e∗2(cid:105) − (cid:104)i2(cid:105) − (cid:104)i∗2(cid:105) h2 α 1 − αh2 (cid:104)i2(cid:105) + (cid:104)i∗2(cid:105) EXINT h2 (cid:35) (cid:17) ((cid:104)i2(cid:105)+(cid:104)i∗ 2(cid:105)) (cid:18) α h2 ((cid:104)i2(cid:105) + (cid:104)i∗2(cid:105)) (cid:19) (241) which makes the weight function readily applicable to the summation on an equidistant radial grid with spacing ∆a. 16.9 Diffusion Coefficient We concentrated on the evolution of the velocity dispersion so far, but scatterings among planetesimals also change the semimajor axis of the disc particles, inducing a diffusive evo- lution of the surface density: ∂Σ ∂t = ∆a(DΣ) (255) The diffusion coefficient D is related to the typical change in semimajor axis ∆a and the timescale T2Body on which planetesimal encounters operate: D ≈ (∆a)2 T2Body (256) If we neglect the radial displacement during an encounter, the change in semimajor axis is solely due to the change of the velocity: Figure 12. Change of the relative eccentricity e2 due to an en- counter of two bodies initially on circular orbits. b/H is the impact parameter in units of the Hill radius. The plot was obtained by integrating Eq. 16. resonant structures in Fig. 12). As the gravitational attrac- tion drops with increasing distance, non-crossing orbits yield ever smaller perturbations with increasing impact parame- ter. Aside from this qualitative behaviour, it is very difficult to derive precise expressions. While the limit of high ve- locities is accessible through the two-body approximation, any general formula involves some empiric interpolation to cover the full parameter space (Rafikov 2003b,a,see the ap- proximations of). Therefore we decided to approximate the weight function such that the main features of the true weight function w(b) are reproduced. While this approach is less accurate, it provides better insight into the involved approximations. We expand the surface density under the integral in Eq. 250 and compare the expansion coefficients for w(b) and the approximation w(b) to derive constraints on the choice of w(b). The lowest non-vanishing order is: l2 = b2w(b)db = b2 w(b)db (252) (cid:90) ∞ −∞ l can be interpreted as the width of the heating zone. Con- dition 252 inspires our choice of the weight function w(b) exp l2 = w(b) = − b2 2l2 1√ 2πl 1 Ω2 (T (i) where T (i) are the radial velocity dispersions of the interacting radial bins and l is adjusted to the findings of Ida (1993). The advantage of the bell curve is that it has a discrete counterpart r + T (j) and T (j) ) + R2 (253) Hill r r r (cid:32) (cid:33) w(b) ≈ 1 ∆a2N N b/∆a + N/2 c(cid:13) 2011 RAS, MNRAS 000, 1–42 (cid:90) ∞ −∞ (cid:18) (cid:19) − GM = − GM 2a r ∆a ≈ 2a2 GM + 1 2 v · ∆v v2 (257) An average over all orientations of the velocity v and the velocity change ∆v yields the mean square change in semi- major axis: (cid:104)(∆v)2(cid:105) (cid:104)(∆a)2(cid:105) ≈ 4a3 3GM 4 3Ω2 (∆Tr + ∆Tφ + ∆Tz) = (cid:18) 5 (cid:19) D ≈ 4 3Ω2 d dt Tr + d dt Tz 4 This yields the mean diffusion coefficient (258) (259) where the time derivatives of the velocity dispersions Tr and Tz are taken with respect to encounters. 16.10 Coagulation Equation We already stated the coagulation equation for a multi-mass system: (cid:90)(cid:90) 0 = ∂ ∂t Fm = − mn(t, m) + ∂ ∂m Fm(t, m) n1(t, m1)n2(t, m2)σ(m1, m2)vrel Mred(m, m1, m2)dm1dm2 (260) (261) N = 4l2/(∆a)2 (254) Since the vertical density profile of a planetesimal disc is specified by the known distribution function, we insert the 0 5 10 15 20 25 30 35 0 0.5 1 1.5 2 2.5 3 3.5 4∆(e2/h2)b/H 32 P. Glaschke, P. Amaro-Seoane & R. Spurzem isothermal density profile (see Eq. 216) in the coagulation equation 260 and integrate over z: (cid:90)(cid:90) 0 = ∂ ∂t Fm = − ∂ ∂m Σ(t, m) + Fm(t, m) (cid:112)2π(h(m1)2 + h(m2)2) 1 (262) Σ1(t, m1) Σ2(t, m2) m1 m2 σ(m1, m2)vrelMred(m, m1, m2)dm1dm2 (263) Σ(m) is a short-hand notation for the differential surface density dΣ dm . Further integration over all masses gives the total mass balance: (cid:90) mmax Σtot = Σ(t, m)dm mmin Σtot = Fm(mmin) − Fm(mmax) (264) ∂ ∂t The calculation of collisional cross sections is closely related to the underlying encounter dynamics. A homogenous sys- tem introduces no systematic perturbation, hence an en- counter is a pure two-body problem which is analytically solvable. Thus it is possible to derive the cross section with- out any approximation. Since encounters in the field of a central star deviate noticeably from the pure Kepler solu- tion, the cross sections are also modified. While the cross section in the high velocity regime reduces to the two-body formula (except minor corrections), the low velocity regime is explored best by numerical calculations. It is not appropri- ate to disentangle the different contributions in Eq. 263, but to combine the various terms to the collisional probability (cid:112)2π(h(m1)2 + h(m2)2) σ(m1, m2)vrel Pcoll = (265) which can be easily deduced from the fraction of colliding orbits in Monte–Carlo simulations. An accurate expression for the collisional probability should include the two-body cross section in the limit of high velocities and the numerical data for the low velocity regime as well. We use numerical calculations from Greenberg (1991); Greenzweig (1992) 26 as a basis for a unified fitting formula σ = σ × 0.572 (1 + 3.67vHill/vrel) 1 + 1.0 (cid:18) (cid:19) (cid:19)−1/2 σΩ2 Tz (266) (267) (268) (cid:18) σ = σgeom 1 + v2∞ v2 rel + 1.8v2 Hill v2 rel = 1 2 (Tr + Tφ + Tz) vHill = ΩrHill The main differences to the two-body cross section 269 is a finite gravitational focusing factor, since the Keplerian shear inhibits a zero relative velocity, and a finite collisional prob- ability for very small velocities, again due to the shear which provides a finite influx of particles. The precise calculation of the coagulation kernel should include an integration over all semimajor axes with a proper weighting kernel. As collisions among particles in the statis- tical model play only a major role when the system is still homogenous, we omitted this contribution. In addition, this helps saving computational time, since the solution of the coagulation equation is very costly. However, interactions between N –body particles and the statistical model include spatial inhomogeneities properly (see section 17). 16.11 Collisional Damping Collisions are a dissipative process that removes kinetic en- ergy from the planetesimal system and damps the eccentric- ity. Low speed encounters leave the colliding bodies intact and damp the relative velocities through inelastic collisions. In contrast, high velocity encounters disrupt the colliding bodies and turn them into an expanding cloud of fragments. As a major part of the initial kinetic energy is converted into heat, the fragments disperse with rather low velocities thus reducing the overall velocity dispersion. We formulate the dissipation due to collisions analogue to Eq. 262: (cid:90)(cid:90) 0 = ∂ ∂t FQ,k = − TkΣ(t, m) + ∂ ∂m 1 FQ,k(t, m) (cid:112)2π(h(m1)2 + h(m2)2) Σ1(t, m1) Σ2(t, m2) m1 m2 σ(m1, m2)vrelQred,k(m, m1, m2)dm1dm2 k ∈ {r, z} (271) Qred,k is the kinetic energy redistribution function and FQ,k is the associated flux across the mass distribution. k indi- cates the two velocity dispersions. Qred,k is a complex func- tion, since the disruption of a planetesimal produces frag- ments with a large scatter in velocities and a complicated velocity field. The velocity of a fragment consists of two con- tributions: The ejection velocity relative to the target and the velocity ¯v of the target within the corotating coordinate system. Owing to the strong dissipation, fragment velocities are dominated by the motion of the centre-of-mass of the two colliding bodies. Thus we neglect the ejection velocities and estimate the centre-of-mass motion. The initial kinetic energy of two colliding bodies is: Eini = 1 2 m1v2 1 + 1 2 m2v2 2 (272) We separate Eini into centre-of-mass motion and relative motion and average over v1, v2: (cid:16) m1v1 + m2v2 (cid:17)2 (cid:1) + (cid:0)m2 1T1 + m2 2T2 M + 1 2 µ(v2 1 + v2 2) 1 2 µ(T1 + T2) (273) which gives an effective cross section σ for planetesimal– planetesimal encounters. Eq. 266 reduces to the well-known gravitational focusing formula in the limit of high velocities: (cid:18) (cid:19) σ ∝ σgeom 1 + v2∞ v2 rel 2G(m1 + m2) v2∞ = R1 + R2 (269) (270) Eini = (cid:104)Eini(cid:105) = 1 2 M 1 2M If the vertical velocity dispersion is small, the disc becomes two-dimensional and the cross section is proportional to R. Most of the relative kinetic energy is dissipated during the collision, so we neglect the relative motion after the incident. The final energy is 26 Their work includes an averaging over the Rayleigh distributed inclinations and eccentricities of the colliding planetesimals. (cid:104)Efinal(cid:105) (cid:104)Eini(cid:105) = m2 1T1 + m2 2T2 M (m1T1 + m2T2) (274) c(cid:13) 2011 RAS, MNRAS 000, 1–42 which gives the drift motion ¯v of the expanding fragment cloud: ¯v2 = (cid:104)Efinal(cid:105) 2 M (275) Qred is therefore coupled to the fragment redistribution func- tion Mred Qred = ¯v2Mred + Qdiss (276) where the additional function Qdiss removes the dissipated energy. 16.12 Correlation The statistical model of a planetesimal disc does not only re- quire a large particle number to assure a proper description of the system by a distribution function, but also the un- correlated motion of the planetesimals. Each of the formulae derived before involves the averaging over different impact parameters to some extend, in combination with the vital assumption that all distances are equally probable. As long as all particles are subjected to perturbations by surround- ing bodies, strong correlations are suppressed. This applies to the early stages, but the formation of protoplanets intro- duces a few dominant bodies that are not susceptible to the perturbation of the field planetesimals. Orbit repulsion gives rise to a regular spacing of the protoplanets, which prevents mutual collisions. Therefore not all impact parameters are equally probable due to this strong correlation. Hence a sta- tistical model is inherently not applicable to the late stages of protoplanetary growth. Since statistical models are superior to N –body calcula- tions with respect to speed and (effective) particle number, modifications have been proposed to remedy this problem. The statistical model by Wetherill (1993) uses the fol- lowing solution: A gravitational range ∆a (or buffer zone) is attached to each planetesimal, which represents the min- imal spacing that allows for stable orbits. They propose the expression ∆a = f∆RHill +(cid:112)2Tr/Ω2 (277) where f∆ is the minimal spacing in terms of the Hill radius. The value of f∆ is adopted from Birn (1973), who derived the minimal spacing that allows for stable orbits: √ 3 f∆ = 2 (278) Thus it is possible to define a minimum mass msep by the as- sumption that all bodies larger than this critical mass main- tain a clear buffer zone: (cid:90) ∞ Hybrid methods: A new composite algorithm 33 Figure 13. Covered fraction fC as a function of f for simulation S1FB at T = 105 years. The protoplanets are already formed and grow oligarchically. f∆ = 10, which is the mean distance of protoplanets accord- ing to the orbit repulsion mechanism. To shed light on the proper gravitational range and the validity of this approach, we defined an additional quantity fC which is the true area (i. e. overlapping is handled prop- erly) covered by the buffer zones in terms of the total area: (cid:90) ∞ msep fC = fC (cid:54) 1 dΣ dm 2πa(∆a)C m dm (280) Fig. 13 shows the covered fraction fC (m) as a function of the integrated buffer zones f (m) for one of our hybrid cal- culations. Both values for f∆ are included as well as the two limiting cases random placing and perfect ordering. Though we tested also other values of f∆, a spacing of ten Hill radii proved to be the best choice. Our own experience with this method indicates that it works reasonably well and agrees with Inaba et al. (2001) who used the same technique. However, this modification in- cludes the regular spacing of the protoplanets in a prescribed way, so any exploration of later stages, like the initiation of orbit crossing, is not accessible through this approach. Therefore we use it for comparison purposes only, since the hybrid code (see section 17) provides a much more general framework. f = msep f = 1 dΣ dm 2πa∆a m dm (279) 16.13 Discretisation f is the area covered by the buffer zones (overlapping is not taken into account, therefore f > 1 is possible), normalised to the ring area. Planetesimals smaller than msep can not en- force a minimum distance to their neighbours, as the whole disc surface is already covered by the buffer zones of the largest bodies. Owing to the regular spacing introduced by the buffer zones, planetesimals larger than the critical mass are not allowed to collide with each other. This approach has also been employed by Inaba et al. (2001), who adopted All involved quantities are only functions of a and m. There- fore we introduce a two dimensional grid, where Σ, Tr and Tz are cell centred quantities. Fig. 14 summarises the defini- tion of the two-dimensional grid. Since the full planetesimal size range covers several orders of magnitude in mass, we chose a logarithmically equidistant discretisation in mass to cover the necessary mass range in a reliable way. The ra- dial spacing of the grid cells is equidistant. Thus the grid setup for the mass discretisation reads (N grid cells from c(cid:13) 2011 RAS, MNRAS 000, 1–42 0 0.2 0.4 0.6 0.8 1 1.2 0 0.5 1 1.5 2 2.5 3fCff∆ =10f∆ =2√3Perfect orderingRandom placing 34 P. Glaschke, P. Amaro-Seoane & R. Spurzem this expression to identify the most important terms, as we can see in Eq. 283. The strongest varying contribution FV is now approxi- mated by a power law with respect to mj: FV (m) ≈ FV (mj) (284) (cid:18) m (cid:19)q mj Thus Eq. 284 is used to provide improved partial fluxes F (jk) and thus we come to Eq. 285. i Since the fragment redistribution function is a piecewise power law, an analytical solution of the integral is possible. Eq. 285 gives reliable results even with a spacing δ = 2. The time derivative of the surface density reads m = −Fi+1 + Fi Σi (286) which assures the conservation of mass within numerical ac- curacy. 16.14 Integrator All contributions to the evolution of the surface density Σ and the velocity dispersions Tr and Tz are summarised by the following set of differential equations: (cid:18) 5 (cid:19) 4 D = 3Ω2 4 Tr,enc + Tz,enc dΣ dt dΣTr dt dΣTz dt = ∆a(DΣ) + Σcoll = ∆a(DΣTr) + Σ Tr + = ∆a(DΣTz) + Σ Tz + d dt d dt (ΣTr)coll (ΣTz)coll (287) The Laplace operator is approximated in accordance with the equidistant radial grid: (cid:18) (cid:19) ∆af = 1 a ∂ ∂a a ∂ ∂a f ∆af ≈ fi+1(1 + ∆a/(2ai)) − 2fi + (1 − ∆a/(2ai))fi−1 (∆a)2 We chose the Heun method 28 as the basic integrator for the statistical model. It is a second order accurate predictor– corrector scheme (X is a vector containing all the above quantities): (288) = f (X) dX dt X p = Xn + ∆tf (Xn) Xn+1 = X p + 1 2 ∆t(f (X p) − f (Xn)) + O(∆t3) (289) The Heun method is readily extended to an iterate scheme, which is equivalent to the implicit expression: Xn+1 = Xn + 1 2 ∆t(f (Xn+1) + f (Xn)) (290) This adds stability to the method and allows the secure in- tegration of stiff configurations that may appear during the Figure 14. Numerical grid. The arrows indicate transport of ki- netic energy (red), spatial transport of mass(green) and accretion (black). Non-neighbouring cells are coupled by the coagulation kernel and the radial interpolation kernel. mmin . . . mmax): mi = mminδ ∆mi = mminδ −i(1/2 + δ/2) −i(1 − δ) i = 1, . . . , N Σi = δ = ∆mi dΣ dm (cid:18) mmin (cid:19)1/N mmax (281) The grid spacing δ controls the number of cells which are necessary to cover a specified mass range. As the evaluation of the coagulation equation scales with the third power of the number of grid cells, δ should be as large as possible. If the flux integral (see Eq. 260) is approximated in a standard way N(cid:88) Fi = − N(cid:88) 1(cid:113) k=1 j=1 = F (jk) i F (jk) i Σj mj Σk mk × σ(mj, mk) 2π(h2 j + h2 k) vrelMred(mi − ∆mi/2, mj, mk) (282) a spacing δ much smaller than 2 is required to guarantee a sufficient accuracy27. However, it is possible to use a spacing of 2 if special precaution is taken. Spaute et al. (1991) ap- proximated the surface density with a power law, thus tak- ing the gradient with respect to mass into account. While they reached only a sufficient accuracy with further special adaptations, we use a more rigorous approach. A large spac- ing δ reduces the accuracy, since the partial flux (Eq. 282) is strongly varying even inside one grid cell. Thus we rearrange 27 Ohtsuki et al. (1990) give a thorough analysis of the impor- tance of the resolution. 28 The name of this method is not unique. Some texts denote it as the modified Euler method. The Heun method is a special case of the Runge–Kutta methods. c(cid:13) 2011 RAS, MNRAS 000, 1–42 mmimi+1aaiai+1FiFi+1ΣiTrTzCollisionsEncounters Hybrid methods: A new composite algorithm 35 F (jk) i = (cid:113) vrel 2π(h2 j + h2 k) Σk mk Σj mj (cid:124) × σ(mj , mk)mj Mred(mi − ∆mi/2, mj , mk) 1 mj (cid:125) (cid:123)(cid:122) FV (mj ) j (cid:62) k F (jk) i = (cid:113) vrel 2π(h2 j + h2 k) (cid:90) mj +∆mj /2 mj−∆mj /2 Σk mk FV (m) m∆mj Mred(mi − ∆mi/2, m, mk)dm (283) (285) j (cid:62) k Figure 15. Interplay between the N –body component and the statistical component of the hybrid code. Black arrows indicate mass transfer, red arrows exchange of kinetic energy and green arrows indicate spatial structuring, respectively. runaway accretion phase. In practice, three iterations are sufficient to guarantee a stable integration. As the diffu- sive part is discretised with a first order accurate formula (see Eq. 288), the whole iterated scheme is equivalent to the Crank–Nicolsen method. We choose a global time step for the statistical model according to the expression , X ∈ {Σ, Tr, Tz} ∆t = min (cid:19) (cid:18) (291) ηDisc X X where the hybrid code (see next section) applies an addi- tional discretisation in powers of two to achieve a better synchronisation with the N –body code component. 17 BRINGING THE TWO SCHEMES TOGETHER: THE HYBRID CODE We introduced two different methods to solve the plan- etesimal growth problem. On the one hand, we modified Nbody6++, which has been used so far mainly for the sim- ulation of stellar clusters, to adapt it to the special require- ments of a long-term integration of planetesimal orbits. On the other hand, we developed a new statistical code with a consistent evolution of the velocity dispersion, the capabil- ity to treat spatial inhomogeneities and a thoroughly con- structed collision treatment. Neither of the two approaches c(cid:13) 2011 RAS, MNRAS 000, 1–42 is powerful enough to provide a complete and accurate de- scription of the planetesimal problem, since each method is confined to a certain range of the particle number. However, these restrictions are complementary in the sense that each method covers a regime where the other method fails. This intriguing relation stimulated the construction of a hybrid code which combines the benefits of both methods. The basic idea is to introduce a transition mass mtrans, which separates the two mass regimes. Particles with a lower mass are treated by the statistical model, whereas larger par- ticles belong to the N –body model. Though both parts are clearly divided in different mass ranges, they are connected by various interdependencies: (i) Direct collisions between particles lead to a mass ex- change. One process is the accretion of small particles by N –body particles, but agglomeration within the statistical model can also produce particles larger than the transition mass. This requires the generation of new N –body parti- cles. Energetic impacts may erode larger particles, so a cor- responding particle removal is also required for consistency. (ii) Mutual scatterings among N –body particles and smaller planetesimals transfer kinetic energy. While energy equipartition leads to a systematic heating of the smaller field planetesimals, a consistent treatment has to include both transfer directions. (iii) Accretion and scattering by the N –body particles in- duce spatial inhomogeneities or even gaps in the planetesi- mal component, if the particles have grown massive enough. Likewise, the small particles could induce some structure in the distribution of the N –body particles. Since the spatial structure is dominated by the stirring from few protoplanets, we neglect the latter process. Fig. 15 summarises this brief overview of the interactions between the two code components in a schematic diagram. The following sections explain the implementation of each interaction term in more detail. 17.1 Mass Transfer An N –body particle accretes smaller particles in its vicinity. We already derived expressions which describe agglomera- tion within the statistical model, so it is manifest to apply these formulae to derive the accretion rate of an N –body particle. Most of the material is accreted within the cross- sectional area σ (see Eq. 267), but the finite eccentricity of an orbit extends the accessible radial feeding zone. Thus Pseudo-ForceDiscParticlesximiviΣ (m,a)Tr (m,a)Tz(m,a)mtransxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxmminAccretionNew ParticlesAccretion/ErosionCollisionsHeat DiscScatter Disc?Fragments 36 P. Glaschke, P. Amaro-Seoane & R. Spurzem we assign the following surface density to each particle (cid:18) (cid:19) Σ(a) = √ M 2πa exp − (a − a0)2 2l2 l2 = σ/π + a2e2 + Tr/Ω2 2πl 1 2 (292) (293) by smearing it out over its feeding zone. Tr is the radial velocity dispersion of particles in the statistical model with semimajor axis a. This density distribution is projected onto the radial grid to calculate the accretion rate. As the time step of the statistical model is much larger than the regu- lar step of an N –body particle, the particle mass update is synchronised with the statistical integration. The projection technique allows the calculation of the accretion rates in a simple way, which gives the right size of the feeding zone and the proper total accretion rate. Particle generation is included in the following way: A "virtual" mass bin is introduced as the boundary between the statistical grid (denoted by the dashed area in Fig. 15) and the N –body component. Its sole task is to store mass and kinetic energy that drives the statistical model towards higher masses. If the mass content exceeds one mtrans, a new particle is created with inclination and eccentricity accord- ing to the stored velocity dispersions. The masses of the N –body particles are regularly checked to detect any particle which dropped below the tran- sition mass. While this procedure would remove the particle and transfer the associated quantities back to the grid, we never observed such a particle erosion. 17.2 Disc Excitation The projection of an N –body particle onto the grid with the help of a proper weight function is also useful for the calculation of the disc excitation due to stirring by the larger particles. Since the Hill radius sets the proper length scale for planetesimal encounters, the weight function is modified to (cid:18) (cid:19) Σ(a) = √ M 2πa exp − (a − a0)2 2l2 l2 = R2 Hill + a2e2 + Tr/Ω2 (294) 2πl 1 2 where Tr is the radial velocity dispersion of the heated plan- etesimal component. The velocity dispersion of the stirring N –body particle is (in accordance with Eq. 11 and Eq. 12): Tr,0 = 1 2 (Ωa0)2e2Tz,0 = 1 2 (Ωa0)2i2 (295) We employ the orbital elements as mediators between the fast varying instantaneous position and velocity of a parti- cle and the slow evolution of the statistical model, which operates on a longer relaxation timescale. In virtue of the projection of the particle, we readily apply the standard in- teraction terms (see section 16) to evaluate the additional heating due to the presence of N –body particles. 17.3 Pseudo–Force While an N –body particle is moving through the disc, it also interacts gravitationally with the particles in the statistical model. The collective effect of all these encounters leads to a change in the orbital elements of the N –body particle. Again, we project the N –body particle onto the grid and evaluate the stirring rates Tr and Tz, which correspond to a change in the orbital elements: d dt d dt e2 = i2 = 2 Tr (Ωa0)2 2 Tz (Ωa0)2 (296) These time derivatives of eccentricity and inclination are translated to a pseudo-force that effects the desired change of the orbital elements. We chose the ansatz Fx,y = Cr(vx,y − (vK )x,y) Fz = Czvz (297) where vK is the local Keplerian velocity. In addition, we tried a simpler expression r · v r2 Fx,y = 2Crrx,y Fz = Czvz (298) without any significant differences in the accuracy or the simulation outcome. The proper friction coefficients are: Cr = Cz = Tr 2Tr Tz 2Tz (299) Since the relevant quantities are the time derivatives of the orbital elements, any other pseudo-force is also applicable. Though this approach yields the right mean change of the orbital elements, it lacks the statistical fluctuations from the particle disc. Hence the distribution of the orbital ele- ments of the N –body particles is artificially narrowed, which is especially important when the N –body particles and the statistical particles have a comparable mass. As the mass contrast between the two code parts is quite significant in planet formation simulations, it is safe to neglect the fluctu- ating part without major restrictions on the realism of the simulations. The friction coefficients Ci are kept constant between two integration steps of the statistical model. While a more frequent update of the coefficients would be easily possible, a regular update on the basis of the statistical time step is ac- curate enough. Moreover, each update poses a considerable computational effort (roughly equivalent to 1000 force eval- uations), so our approach also saves valuable computational time. 17.4 Spatial Structure The first insight into planetesimal formation was obtained by the particle–in–a–box method, which invokes the under- lying assumption that the planetesimal disc stays homoge- neous throughout the protoplanet growth (Greenberg et al. 1978,see e.g.). While few large bodies introduce some coarse- graininess of the surface density, all smaller bodies are as- sumed to be evenly spread in the disc. Research on the in- teraction of protoplanets showed that this is an oversimpli- fication, as bodies that are massive enough could open gaps in their vicinity (Lin & Papaloizou 1979; Rafikov 2001,see c(cid:13) 2011 RAS, MNRAS 000, 1–42 Hybrid methods: A new composite algorithm 37 No G1 G2 Σ 1.1251 × 10−6 0.2 1406 ∆a N Nrad e2/h2 i2/h2 Perturber 1.1251 × 10−6 - 1 0.2 – Perturber - 1 – – 201 – 0.00135 e = 6.1 × 10−5 0.00135 i = 3.2 × 10−5 0.00135 e = 6.1 × 10−5 0.00135 i = 3.2 × 10−5 Type m 1 × 10−9 N –body 1 × 10−7 1 × 10−9 1 × 10−7 Statistic Table 10. Parameters of the statistical and the N –body gap simulation. The perturber is placed at the centre of the ring. Figure 16. Gap opening in a planetesimal disc. The gap is fully developed after 2000 years. Table 10 summarises the initial con- ditions for the comparative runs. Figure 17. Mean square eccentricity and inclination of the smaller planetesimals in terms of the reduced Hill radius H of the protoplanet according to simulation G1 (N –body) and G2 (Statistic). e.g.). Gap formation does not only change the overall surface density, but also controls the accretion onto the protoplanet through the amount of planetesimals in the feeding zone. If gap formation is too effective, the growth of the protoplanet may well stop before the isolation mass is reached. Hence any hybrid code should provide a framework that allows this mechanism to operate. A necessary condition is a radial density grid with a sufficient resolution to describe possibly emerging gaps. A too low resolution suppresses local pertur- bations from the protoplanets by a simple averaging, thus inhibiting the formation of any spatial inhomogeneities. A second requirement is that the interaction terms relating statistical model and N –body model include the local inter- action between particles and the statistical component in a proper way. Our hybrid approach includes gap formation implicitly through the diffusive terms. A protoplanet heats only the planetesimals in its vicinity (defined by the heating kernel), thus also increasing locally the diffusion coefficient. Hence the surface density drops due to outward diffusion of the planetesimals, given that the protoplanet is massive enough. The minimum gap opening mass is set by the condition that the protoplanet controls the random velocities of the field planetesimals in its heating zone (see Eq. 30), which is equiv- alent to the independently derived gap formation criterion (compare Eq. 32). Although our algorithm invokes a simplified picture of the protoplanet–planetesimal interaction, it is surprisingly accurate with respect to the width of the forming gap and the opening criterion. Fig. 16 shows a simulation which ex- c(cid:13) 2011 RAS, MNRAS 000, 1–42 amines the accuracy of our approach. The overall perfor- mance of the statistical code is quite remarkable, except a significant overestimation of the surface density at the gap boundary compared to the N –body model. This deviation is due to the improper treatment of strong planetesimal– protoplanet encounters, which exceed the diffusive approxi- mation. Moreover, the higher concentration of planetesimals near the gap boundary leads to an additional overestimation of the velocity dispersion of the smaller planetesimals in the statistical calculation (see Fig. 17). While the comparison with the N –body calculation clearly indicates a necessary improvement of the treatment of spatial inhomogeneities, our approach catches the main features of gap formation. 17.5 Transition Mass Since the inventory of the new hybrid code is now completed, we turn to the specification of the transition mass mtrans. The mass boundary between statistical and N –body part has a major influence on the realism and the speed of the simulation. On the one hand, optimisation with respect to speed favours a large transition mass, whereas a reasonable resolution of the transition between the two components in- troduces some upper limit. Hence we identify first the set of large masses, which controls the velocity dispersion of the disc, since these ob- jects are also possible candidates for gap opening. The in- spection of all involved stirring terms gives approximately 0.5 0.6 0.7 0.8 0.9 1 1.1 1.2 1.3 1.4 1.5 0.94 0.96 0.98 1 1.02 1.04 1.06Σ/Σ0rGap at T=2000 yrN-BodyStatistic 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 0.94 0.96 0.98 1 1.02 1.04 1.06<e2/H2>1/2,<i2/H2>1/2rT=2000 yrN-Body (e)Statistic (e)N-Body (i)Statistic (i) 38 P. Glaschke, P. Amaro-Seoane & R. Spurzem the inequality:(cid:90) mtrans 0 (cid:90) ∞ mtrans dΣ dm dΣ dm mdm < mdm (300) While this is a necessary condition to select all potential ma- jor perturbers, criterion 300 does not imply that all particles in the selected mass range exert indeed a strong influence on the disc. The number of possible gaps – and therefore the number of perturbers associated with them – is ultimately limited by the available space. Thus we integrate the area of all potential gaps (width ≈ f∆RHill) and normalise it to the total disc area: fC ≈ (cid:90) ∞ 2πaRHill dm (301) f∆ dΣ dm m mtrans If the covered fraction fC is much larger than one, it is possible to increase the transition mass until the condition fC (cid:47) 1 (302) is fulfilled. Of course, condition 300 and 302 defined only an upper limit of the transition mass, so the adaptation of a lower value is also possible. Though there are two reliable conditions at hand, the transition mass is still a function of time owing to the time evolution of the density Σ(m). Therefore we chose a priori a fiducial value of the transition mass, run the simulation and conduct an a posteriori check, whether the initial choice matches our requirements at any evolutionary stage of the disc. A reliable value for a solar system analogue at 1 AU is mtrans ≈ 3 × 10 −11M(cid:12) (303) which restricts the number of N –body particles to a tractable amount. Later stages would allow an even larger transition mass, but the current hybrid code does not in- clude any dynamical adjustment of the transition mass at runtime. 17.6 Boundary Conditions Any numerical simulation is limited to a finite simulation volume and a finite time interval. Therefore it is mandatory to introduce proper boundary conditions which provide a reasonable closure of the simulation volume. While boundary conditions with respect to time are the familiar initial conditions, the choice of the spatial bound- ary conditions for the various involved quantities depends on the problem at hand and the type of the boundary. A sim- ulation boundary can be due to physical reasons (like walls of a concert hall, surface of a terrestrial planet) or simply due to a limitation in computational power that inhibits the complete numerical coverage of the problem. The current capability of the hybrid code sets limits on the radial range as well as on the covered mass range, which a simulation can handle in a reasonable time. Hence we have to introduce artificial boundaries in radius, and a lower limit for the mass grid. Any migration process couples the evolution of a lo- cal ring area in the planetesimal disc to the evolution of the whole disc. Inward (or outward) migrating material also transports information on the radial zone where the material originated from. As this information is not available within the frame of a local simulation, any choice of the boundary condition alters the evolution to some extend. However, we focus on a formation stage where migra- tion is not a dominant process, but provides only removal of the smaller collisional fragments. Thus we apply closed boundary conditions for the outer and inner radius of the ring area (i. e. all fluxes vanish at the boundary), and an open boundary for the lower end of the mass range. While these conditions exclude the study of migrational processes, we gain clearer insight into the protoplanet growth. 18 DISCUSSION AND CONCLUSIONS The formation of planetary systems represents a challenge from a numerical standpoint. The dynamical problem spans over many orders of magnitudes in length and demands the combination of different techniques. We have presented a composite algorithm that brings together the advantages of direct-summation tools and statistics for the description of the planetesimal disc. Direct-summation N−body tech- niques have been around for some decades and have proven their accuracy in a very large number of studies of stellar clusters such as galactic nuclei and globular and open clus- ters. We deem it to be the numerical tool to integrate the motion of the bodies for the very precise integration of the orbits and treatment of close encounters. Typically, in a sim- ulation of a stellar system, the energy is conserved in each timestep by E/∆E ∼ 10−11 (where E is the total energy and ∆E the difference between the former and current to- tal energy for a specific time), so that even if we integrate for a long time the cluster, the accumulated energy error is negligible. Nevertheless, porting the numerical tool to the problem of planetary dynamics is not straighforward and re- quires important modifications and additions. In this work we present them in detail: the neighbour radius selection for the protoplanets, the Hermite iteration and we introduce for the very first time the new extended Hermite scheme, since the usual Hermite scheme is not sufficient to integrate planetesimal orbits accurately enough. Then we bring in new forces to the problem, namely the introduction of the central potential of the star, as well as the drag forces, which depend on the gas density and size of the planetesimals. Hence, the regularisation scheme, crucial to exactly integrate the close encounters, has to be accordingly modified. We then intro- duce the disc geometry and discuss the required changes to the neighbour scheme and prediction, as well as the commu- nication algorithm and block size distribution. For the statistical description of the planetesimal disc we employ a Fokker-Planck approach. We include dynamical friction, high- and low-speed encounters, the role of distant encounters as well as gas and collisional damping and then generalise the model to inhomogenous discs. We then de- scribe the combination of the two techniques to address the wole problem of planetesimal dynamics in a realistic way via a transition mass to integrate the evolution of the particles according to their masses. In particular, we introduce and describe the extended Hermite scheme, which reduces the the energy error by three orders of magnitude with the same number of force evalua- tions, compared to the standard version of Nbody6++. While the implementation and some code details are c(cid:13) 2011 RAS, MNRAS 000, 1–42 Hybrid methods: A new composite algorithm 39 ∆a N Nrad e2/h2 i2/h2 m Type 0.02 1000 0.02 500 0.02 – 0.08 800 200 0.08 – 200 – – 0.02 10.000 0.02 0.02 – – – – – 10 10 – – 10 – – – 10 10 – 0.04 0.04 0.04 0.01 0.01 0.01 4 4 4 4 – 4 4 4 1 1 1 1 – 1 1 1 620 155 1.41 × 10−10 N –body 1.41 × 10−10 Hybrid 1.41 × 10−10 5 × 10−10 2 × 10−9 5 × 10−10 2 × 10−9 N –body Statistic Hybrid – Statistic 1.41 × 10−11 N –body 1.41 × 10−11 Hybrid 1.41 × 10−11 2.4 × 10−15 Statistic Statistic No. T1a T1b T1c T2a T2b T3 T4a T4b T4c T5 Σ 1.1251 × 10−6 1.1251 × 10−6 1.1251 × 10−6 0.5626 × 10−6 0.5626 × 10−6 0.5626 × 10−6 0.5626 × 10−6 Safronov 1.1251 × 10−6 1.1251 × 10−6 1.1251 × 10−6 1.8789 × 10−6 Table 11. Parameters of all test simulations. The transition mass in T4b is mtrans = 3.1 × 10−10 Only simulations T3, T4a–T4c and T5 include collisions. All values are scaled to Mc = G = r0 = 1. newly introduced to the field of planet formation simula- tions, the first hybrid approach was developed in the early 90's. Spaute et al. (1991) (further improved in Weiden- schilling et al. 1997) constructed a hybrid code with a statis- tical component to treat the smaller particles and a special treatment for the larger particles. A statistical model cov- ers the field planetesimals with the help of a distribution function (similar to Wetherill (1989)), whereas the larger particles are individually stored and characterised by mass, semimajor axis, eccentricity and inclination. While the inter- action between these single particles and the statistical com- ponent is expressed by standard viscous stirring and dynam- ical friction terms, perturbations among the single particles are equated in a different way. First, the probability of an encounter of two neighbouring particles is calculated. This probability is used in a second step to decide whether a (nu- merically integrated) two-body encounter of the neighbour- ing particles is carried out to derive the change in the orbital elements. Though these two well-defined code components justify to speak about a hybrid approach, the Monte–Carlo like integration of the largest particles is still closely related to a statistical treatment. A modified N –body approach is used in the work of Lev- ison & Morbidelli (2007). Their method covers the largest particles by a direct N –body code, which includes the smaller particles as "tracer" particles. The term "tracer" indicates that each particle represents a whole ensemble of planetesimals. In a similar line of approach and inspired by this idea, Levison et al. (2010) modified a symplectic algo- rithm, Symba, to study the formation of giant planet cores. However, they made some assumptions in order to calcu- late the gravitational interaction between the planetesimals. In particular, they ignored totally close encounters between planetesimals. Ormel & Spaans (2008) present in their work a scheme based on Monte Carlo techniques to cover the vast range of c(cid:13) 2011 RAS, MNRAS 000, 1–42 sizes. For this, they assign more resolution to those parti- cles that are more relevant to the interactions, typically the largest bodies. Smaller particles are grouped and treated collectively, which means that they all share the same mass and structural parameters. This classification is done in ac- cordance to the "zoom factor", a free paramenter. Later, Ormel et al. (2010a) presented an detailed comparison of their Monte Carlo code with other techniques, in particular with pure direct-summation N−body results and other sta- tistical studies and found that system leaves the runaway at a larger radius, in particular at the outer disc. With their simulations, the authors propose a new criterion for the run- away growth-oligarchy transition: from several hundreds of km in the inner disk regions up to a thousand km for the outer disc (Ormel et al. 2010b). Bromley & Kenyon (2006) published a description of a hybrid method with a basic approach similar to our work. They employ two velocity dispersions and the surface den- sity of the planetesimals to describe the planetesimal system. The statistical component includes migration of the plan- etesimals and dust particles due to gas drag and Pointing– Robertson drag. In contrast to our approach, they did not include mass transport due to the diffusion of the planetesi- mals, which precludes the study of spatial structures induced by the protoplanets. One must note also that their method uses the standard discretisation of the collisional flux (see Eq. 282) and thus restrict the spacing factor to δ (cid:46) 1.25 (Kenyon 1998). Bromley & Kenyon (2006) chose a set of test calculations which focused less on the technical aspects of their method, but on an overall comparison with a se- lected set of standard works on planet formation. Their test simulations are in good agreement with the references simu- lations, thus indicating a comparable quality of the method. Four years later, the authors presented an updated version of their code for planet formation. The new characteristics of the code included 1D evolution of the viscous disc, gas ac- 40 P. Glaschke, P. Amaro-Seoane & R. Spurzem cretion on to massive cores, as well as accretion of small par- ticles in planetary atmospheres (Bromley & Kenyon 2010). While a variety of hybrid approaches emerged over the past years, this technique is still far from a routinely appli- cation and is still challenged by many open issues. Hybrid codes bear the potential to address the dynamical evolution of a whole planetary system, the later stages of protoplanet formation initiate a strong interaction with the gaseous disc, which may require more diligence than the inclusion of a few additional interaction terms. However, the development is picking up speed, which places our work in a good position for further research. APPENDIX A: CENTRAL FORCE – DERIVATIVES Central force F per mass (i. e. acceleration) and its time derivatives are: F = − xM x3 F(1) = − vM F(2) = − aM F(3) = − aM x3 − 3AF x3 − 6A F − 3BF x3 − 9AF(2) − 9BF(1) − 3CF x · v x2 v2 x2 + 3v · a x2 + x · a x2 + A2 = A + 3A2 x · a x2 + A(3B − 4A2) a = v A = B = C = (A1) The F(i) denote the central force and its time deriva- tives, whereas a and a refer to the total acceleration of the particle. The assumption that x, v, a and a are independent of each other allows the derivation of averaged expressions for particle–particle interactions: (cid:104)(F)2(cid:105) = m2 1 x4 (cid:104)(F(1))2(cid:105) = m2 2v2 x6 (cid:104)(F(2))2(cid:105) = m2 12 (cid:18) (cid:18) (cid:104)(F(3))2(cid:105) = m2 144 (cid:19) a2 x6 v4 x8 + 2 v6 x10 + 126 (cid:19) a2v2 x8 + 2 a2 x6 (A2) We combine these expressions with Aarseth's time step for- mula to derive the regular time step as a function of the neighbour sphere radius Rs: ∆treg ≈ √ ηreg Rs ¯v 1 +(cid:112)Rs/R0 1 R0 = 4 ¯v2 a ≈ 4¯v2 ¯r2 Gm = 4 ¯r2 rclose (A3) ¯r is the average particle distance and rclose is the impact parameter for a 90–degree deflection. APPENDIX B: SCALABLE COLLISIONS FLUX The mass flux according to the perturbation equation 161 is: (cid:90)(cid:90) Fp = − (n(m2)∆n(m1) + n(m1)∆n(m2)) σ(m1)v(m2)m1fm(m1/m, )dm1dm2 = F (1) + F (2) (B1) Firstly, we employ the substitution m1 = mx1 m2 = m0 (cid:18) m1 m0 (cid:19) 1+α 1+2w (cid:16) S (cid:17) 1 (cid:90) to solve for the partial flux F (1): 1+2w  1 1+2w (B2) (cid:18) mx1 (cid:90) m0 F (1) = −n2 0m3 −k(cid:48)(cid:90) F1(x1) = S 0σ0v0 g(mx1)F1(x1)dx1 − w+s+3+α  2+α+2w fm(x1, ) x1(1 + 2w) d (B3) The second contribution F (2) requires a slightly different transformation: m1 = mx1 −1/(1+α) (cid:19) 1+α (cid:17) 1 1+2w (cid:16) S 1+2w m2 = m0 (B4) Thus the partial flux F (2) is: F (2) = −n2 0m3 −k(cid:48)(cid:90) F2(x1) = S 0σ0v0 g(m2)F2(x1)dx1 − w+s+3+α 2+α+2w  fm(x1−1/(1+α), ) x1(1 + 2w) d (B5) We change to a new set of logarithmic coordinates ln( S) 1 + α u = ln(m/m0) u1 = ln(x1) s = (B6) which transforms the total flux Fm to a convolution integral: Fp = −n2 0m3 0σ0v0 [g(u + u1)G1(u1)+ g (p(u + u1 + s1)) G2(u1)]du1 p = 1 + α 1 + 2w (B7) (B8) p = 1 refers to the already derived solution for self-similar collisions. Hence we expand Eq. B7 at p = 1 and retain only the zeroth-order moment of the fragmentation kernel: 0m3 0σ0v0 g(u)G1,0 + (g(u)+ Fp = −n2 u(p − 1) ∂g ∂u )G2,0 This expression is equivalent to Fp = −n2 0m3 0σ0v0(g(u)G1,0+ [g(u) + (p − 1)(g(u) − g(0))]G2,0) (B9) (B10) where higher derivatives of g(u) are neglected. Hence we recover the same functional form of the perturbed mass flux Fp as for self-similar collisions: 0σ0v0g(u) (G1,0 + p G2,0) + const. Fp = −n2 −k(cid:48) ∝ S 0m3 (B11) c(cid:13) 2011 RAS, MNRAS 000, 1–42 (cid:90) (cid:104) (cid:105) APPENDIX C: COAGULATION EQUATION While the success of a general approximation of the coag- ulation equation depends heavily on the used coagulation kernel, we nevertheless provide a more general approach to embed section 10 in a broader context. The standard coag- ulation equation is: (cid:90)(cid:90) 0 = ∂ ∂t Fm = − mn(t, m) + ∂ ∂m Fm(t, m) n(t, m1)n(t, m2)σ(m1, m2) vrelMred(m, m1, m2)dm1dm2 (C1) In virtue of our experience drawn from the perturbation expansion, we transform the coagulation equation to loga- rithmic coordinates u = ln(m) (C2) and employ the size distribution g(u) relative to the steady- state solution neq(m): g(u, t)neq(u)e2u + ∂ ∂u Fu(t, m) (cid:90)(cid:90) ∂ 0 = ∂t Fu = − g(t, u1)g(t, u2)K(u, u1, u2)du1du2 (C3) K(u, u1, u2) is the properly transformed new coagulation kernel. g(u) is expanded under the integral to arrive at a moment expansion of the flux Fu: (cid:90)(cid:90) Fu = −K00(u)g(u)2 − (K10(u) + K01(u))g(u) Kij = K(u, u1, u2)ui 1uj 2du1du2 ∂g ∂u + . . . (C4) Retaining only the leading order terms, we recover an ap- proximate coagulation equation which is similar to the in- viscid Burgers' Equation 29: ACKNOWLEDGMENTS It is a pleasure to thank Sverre Aarseth, Cornelis Dullemond and Phil Armitage for comments on the manuscript. PAS thanks the National Astronomical Observatories of China, the Chinese Academy of Sciences and the Kavli Institute for Astronomy and Astrophysics in Beijing, for an extended visit, as well as the Aspen Center of Physics and the orga- nizers of the summer meeting, where this work was finished. PAS expresses his utmost gratitude to Hong Qi, Wenhua Ju and Xian Chen for their hospitality during his stay in Beijing. He is also somehow marginally indebted with the 2011 winter strain of German H1N2, which allowed him to skip all possible duties and focus for an extended period of time on this work at home. RS acknowledges support by the Chinese Academy of Sciences Visiting Professorship for Senior International Scientists, Grant Number 2009S1-5 (The Silk Road Project). The special supercomputer Laohu Hybrid methods: A new composite algorithm 41 at the High Performance Computing Center at National As- tronomical Observatories, funded by Ministry of Finance un- der the grant ZDYZ2008-2, has been used. Simulations were also performed on the GRACE supercomputer (grants I/80 041-043 and I/84 678-680 of the Volkswagen Foundation and 823.219-439/30 and /36 of the Ministry of Science, Research and the Arts of Baden-urttemberg). Computing time on the IBM Jump Supercomputer at FZ Julich is acknowledged. REFERENCES Aarseth S., 1985, Academic Press Orlando, 378 -, 1999, PASP, 111, 1333 -, 2003, Cambridge University Press Adachi I., Hayashi C., Nakazawa K., 1976, Progress of The- oretical Physics, 56, 1756 Ahmad A., Cohen L., 1973, Journal of Computational Physics, 12, 389 Armitage P. J., 2010, ArXiv e-prints Astakhov S., Lee E., Farrelly D., 2005, MNRAS, 360, 401 Bagatin A., Cellino A., Davis D., Farinella P., Paolicchi P., 1994, Planet. Space. Sci., 42, 1079 Balbus. S., 2003, Ann. Rev. Astron. Astrophys., 41, 555 Balbus S., Hawley J., 1998, Reviews of Modern Physics, 70, 1 Bateman H., 1915, Monthly Weather Review, 43, 163 Beauge C., Aarseth S. J., 1990, MNRAS, 245, 30 Beckwith S., 1996, Nature, 383, 139 Benz W., 1999, Icarus, 142, 5 Binney J., 1977, MNRAS, 181, 735 -, 1994, Princeton University Press, Princeton, New Jer- sey Birn J., 1973, Astronomy and Astrophysics, 24, 283 Blum J., Wurm G., 2000, Icarus, 143, 138 Bodenheimer P., 1986, Icarus, 67, 391 Bodenheimer P., Hubickyj O., Lissauer J., 2000, Icarus, Bromley B. C., Kenyon S. J., 2006, AJ, 131, 2737 -, 2010, ArXiv e-prints -, 2011, ArXiv e-prints Burgers J., 1948, Adv. Appl. Mech., 1, 171 Cameron A. G. W., 1973, Space Science Reviews, 15, 121 Cazenave A., Lago B., Dominh K., 1982, Icarus, 51, 133 Chambers J., 2006, Icarus, 180, 496 Chandrasekhar S., 1942, Astrophysical Journal, 97, 255 Connaughton C., Rajesh R., Zaboronski O., 2004, Physical Review E, 69, 061114 Dohnanyi J., 1969, Journal of Geophys. Research, 74, 2531 Duncan M. J., Levison H. F., Lee M. H., 1998, 116, 2067 Glaschke P., 2003, in ESA Special Publication, Vol. 539, Earths: DARWIN/TPF and the Search for Extrasolar Terrestrial Planets, M. Fridlund, T. Henning, & H. La- coste, ed., pp. 425–428 Goldreich P., Lithwick Y., Sari R., 2004, Astrophysical Journal, 614, 497 Goldreich P., Ward W. R., 1973, ApJ, 183, 1051 Greenberg R., 1991, Icarus, 94, 98 Greenberg R., Hartmann W., Chapman C., Wacker J., 1978, Icarus, 35, 1 0 = ∂ ∂t g(u, t)neq(u)e2u − ∂ ∂u (C5) 143, 2 (cid:0)K00(u)g(u)2(cid:1) 29 This notion goes back to Burgers (1948), but the equation was already introduced by Bateman (1915). Greenzweig Y., 1992, Icarus, 100, 440 Hasegawa M., 1990, Astronomy and Astrophysics, 227, 619 c(cid:13) 2011 RAS, MNRAS 000, 1–42 42 P. Glaschke, P. Amaro-Seoane & R. Spurzem Hayashi C., 1981, Progress of Theoretical Physics Supple- ment, 70, 35 H´enon M., 1986, Celestial Mechanics, 38, 67 Hernquist L., Hut P., Makino J., 1993, ApJ Lett., 402, L85+ Hill G., 1878, American J. Math., 1, 5 Housen K., 1990, Icarus, 84, 226 Ida S., 1990, Icarus, 88, 129 -, 1992, Icarus, 96, 107 -, 1993, Icarus, 106, 210 Ida S., Guillot T., Morbidelli A., 2008, ApJ, 686, 1292 Ida S., Kokubo E., Makino J., 1993, MNRAS, 263, 875 Inaba S., Barge P., Daniel E., Guillard H., 2005, Astronomy Papaloizou J., 2006, Rep. Prog. Phys., 69, 119 Paszun D., Dominik C., 2009, A&A, 507, 1023 Petit J.-M., 1986, Icarus, 66, 536 Plummer H. C., 1911, MNRAS, 71, 460 Rafikov R., 2001, Astronomical Journal, 122, 2713 -, 2003a, Astronomical Journal, 125, 922 -, 2003b, Astronomical Journal, 125, 906 Reipurth B., Jewitt D., Keil K. E., 2007, Lunar and Plan- etary Information Bulletin, issue 110, p. 25 (May 2007), 110, 25 Safronov V., 1969, NASA-TTF Skeel R., 1999, Applied Numerical Mathematics, 29, 3 Spaute D., Weidenschilling S., Davis D., Marzani F., 1991, and Astrophysics, 431, 365 Icarus, 92, 147 Inaba S., Tanaka H., Nakazawa K., Wetherill G., Kokubo E., 2001, Icarus, 149, 235 Johansen A., Klahr H., Henning T., 2006, ApJ, 636, 1121 Johansen A., Oishi J. S., Mac Low M.-M., Klahr H., Hen- ning T., Youdin A., 2007, Nat, 448, 1022 Kary D., Lissauer J., Greenzweig Y., 1993, Icarus, 106, 288 Kempf S., Pfalzner S., Henning T. K., 1999, Icarus, 141, 388 Kenyon S., 1998, Astronomical Journal, 115, 2136 Kobayashi H., Tanaka H., Krivov A. V., Inaba S., 2010, Icarus, 209, 836 Kokubo E., 1995, Icarus, 114, 247 -, 1996, Icarus, 123, 180 -, 1997, Icarus, 131, 171 Kokubo E., Ida S., 2000, Icarus, 143, 15 -, 2002, ApJ, 581, 666 Kokubo E., Kominami J., Ida S., 2006, ApJ, 642, 1131 Kokubo E., Yoshinaga K., Makino J., 1998, MNRAS, 297, 1067 Kornet K., Stepinski T., R´ozyczka M., 2001, Astronomy and Astrophysics, 378, 180 Kustaanheimo P. E., Stiefel E. L., 1965, J. Reine Angew. Math. Larson R., 1970, MNRAS, 147, 323 Levison H., Morbidelli A., 2007, Icarus, 189, 196 Levison H. F., Thommes E., Duncan M. J., 2010, 139, 1297 Lin D., Papaloizou J., 1979, MNRAS, 188, 191 Makino J., 1988, Astrophysical Journal Supplement, 68, 833 -, 1992, PASJ, 44, 141 Masset F. S., Papaloizou J. C. B., 2003, ApJ, 588, 494 Mikkola S., 1997, Celestial Mechanics and Dynamical As- tronomy, 67, 145 Mikkola S., Aarseth S. J., 1998, New Astronomy, 3, 309 Moore A., Quillen A. C., 2010, ArXiv e-prints Nakagawa Y., H., N., 1983, Icarus, 54, 361 O'Brien D., 2003, Icarus, 164, 334 O'dell C., Wen Z., Hu X., 1993, Astrophysical Journal, 410, 696 Ohtsuki K., Nakagawa Y., Nakazawa K., 1990, Icarus, 83, 205 Ohtsuki K., Stewart G., Ida S., 2002, Icarus, 155, 436 Ormel C. W., Dullemond C. P., Spaans M., 2010a, ApJ Lett., 714, L103 -, 2010b, Icarus, 210, 507 Ormel C. W., Spaans M., 2008, ApJ, 684, 1291 Paardekooper S.-J., Baruteau C., Kley W., 2011, MNRAS, 410, 293 Spurzem R., 1999, The Journal of Computational and Ap- plied Mathematics, Special Volume Computational Astro- physics, 109, 407 Stewart G., 2000, Icarus, 143, 28 Tanaka H., Inaba S., Nakazawa K., 1996, Icarus, 123, 450 Wada K., Tanaka H., Suyama T., Kimura H., Yamamoto T., 2009, ApJ, 702, 1490 Ward W., 1986, Icarus, 67, 164 Weidenschilling S., Spaute D., Davis D., Marzari F., Oht- suki K., 1997, Icarus, 128, 429 Weidenschilling S. J., 1977, MNRAS, 180, 57 Wetherill G., 1989, Icarus, 77, 330 -, 1993, Icarus, 106, 190 Wisdom J., 1991, Astronomical Journal, 102, 1528 c(cid:13) 2011 RAS, MNRAS 000, 1–42
1311.0201
2
1311
2013-11-21T07:56:23
Exploring atmospheres of hot mini-Neptunes and extrasolar giant planets orbiting different stars with application to HD 97658b, WASP-12b, CoRoT-2b, XO-1b and HD 189733b
[ "astro-ph.EP", "astro-ph.SR" ]
We calculated an atmospheric grid for hot mini-Neptune and giant exoplanets, that links astrophysical observable parameters- orbital distance and stellar type- with the chemical atmospheric species expected. The grid can be applied to current and future observations to characterize exoplanet atmospheres and serves as a reference to interpret atmospheric retrieval analysis results. To build the grid, we developed a 1D code for calculating the atmospheric thermal structure and link it to a photochemical model that includes disequilibrium chemistry (molecular diffusion, vertical mixing and photochemistry). We compare thermal profiles and atmospheric composition of planets at different semimajor axis (0.01$\leq$a$\leq$0.1AU) orbiting F, G, K and M stars. Temperature and UV flux affect chemical species in the atmosphere. We explore which effects are due to temperature and which due to stellar characteristics, showing the species most affected in each case. CH$_4$ and H$_2$O are the most sensitive to UV flux, H displaces H$_2$ as the most abundant gas in the upper atmosphere for planets receiving a high UV flux. CH$_4$ is more abundant for cooler planets. We explore vertical mixing, to inform degeneracies on our models and in the resulting spectral observables. For lower pressures observable species like H$_2$O or CO$_2$ can indicate the efficiency of vertical mixing, with larger mixing ratios for a stronger mixing. By establishing the grid, testing the sensitivity of the results and comparing our model to published results, our paper provides a tool to estimate what observations could yield. We apply our model to WASP-12b, CoRoT-2b, XO-1b, HD189733b and HD97658b.
astro-ph.EP
astro-ph
Draft version May 31, 2018 Preprint typeset using LATEX style emulateapj v. 5/2/11 EXPLORING ATMOSPHERES OF HOT MINI-NEPTUNES AND EXTRASOLAR GIANT PLANETS ORBITING DIFFERENT STARS WITH APPLICATION TO HD 97658B, WASP-12B, COROT-2B, XO-1B AND HD 189733B Max Planck Institut fur Astronomie, Konigstuhl 17, 69117, Heidelberg, Germany Y. Miguel Max Planck Institut fur Astronomie, Konigstuhl 17, 69117, Heidelberg, Germany Harvard Smithsonian Center for Astrophysics, 60 Garden St., 02138 MA,Cambridge, USA L. Kaltenegger Draft version May 31, 2018 ABSTRACT We calculated an atmospheric grid for hot mini-Neptune and giant exoplanets, that links astrophys- ical observable parameters- orbital distance and stellar type- with the chemical atmospheric species expected. The grid can be applied to current and future observations to characterize exoplanet at- mospheres and serves as a reference to interpret atmospheric retrieval analysis results. To build the grid, we developed a 1D code for calculating the atmospheric thermal structure and link it to a pho- tochemical model that includes disequilibrium chemistry (molecular diffusion, vertical mixing and photochemistry). We compare thermal profiles and atmospheric composition of planets at different semimajor axis (0.01≤a≤0.1AU) orbiting F, G, K and M stars. Temperature and UV flux affect chemical species in the atmosphere. We explore which effects are due to temperature and which due to stellar characteristics, showing the species most affected in each case. CH4 and H2O are the most sensitive to UV flux, H displaces H2 as the most abundant gas in the upper atmosphere for planets receiving a high UV flux. CH4 is more abundant for cooler planets. We explore vertical mixing, to inform degeneracies on our models and in the resulting spectral observables. For lower pressures observable species like H2O or CO2 can indicate the efficiency of vertical mixing, with larger mixing ratios for a stronger mixing. By establishing the grid, testing the sensitivity of the results and com- paring our model to published results, our paper provides a tool to estimate what observations could yield. We apply our model to WASP-12b, CoRoT-2b, XO-1b, HD189733b and HD97658b. Subject headings: planetary systems – planets and satellites: atmospheres 1. INTRODUCTION Characterization of hot mini-Neptunes (planets with masses larger than 10 M⊕ and a primary atmosphere) and giant planets's atmospheres has been shown for transiting exoplanets, either by secondary eclipse mea- surements or by transmission spectroscopy (see e.g. Seager & Deming (2010), and references therein). Some chemical species (Na, CO and H2O) have been con- firmed to exist in the outer atmosphere of some tran- siting exoplanets. Sodium lines were detected for HD 209458b (Charbonneau et al. 2002) and HD 189733b (Redfield et al. 2008). For HD 209458b carbon monox- ide (Snellen et al. 2010) and water (Deming et al. 2013) were observed. For HD 189733b, CO was also de- tected (de Kok et al. 2013). Carbon monoxide was also found in the atmosphere of tau Bootis b (Brogi et al. 2012), water in XO-1b (Deming et al. 2013) and both CO and water vapor was detected in HR 8799c 's at- mosphere (Konopacky et al. 2013). These detections along with atmospheric parameter retrieval using broad- band photometric data (e.g.,Stevenson et al. (2010); Madhusudhan & Seager (2011); Line et al. (2013)) ex- plore the underlying chemistry in hot exoplanet atmo- spheres, showing a variety of atmospheric composition of hot exoplanets. Our paper explores the change in chemical atmospheric [email protected] species in the observable region of the atmosphere, as a function of astrophysical observable parameters, like orbital distance as well as stellar type. We devel- oped a 1D code for calculating the thermal profile and link it to an atmospheric model (see Kopparapu et al. (2012)) to explore the atmospheric chemistry of exo- planets. We perform simulations comparing thermo- chemical equilibrium to disequilibrium chemistry driven by vertical mixing, molecular diffusion and photochem- istry, in order to explore the differences in the at- mospheres (see, Hubeny & Burrows (2007) for a study on disequilibrium chemistry in brown dwarfs' atmo- spheres). Most of the photochemical models developed in the last years were applied to specific highly irradi- ated exoplanets, like the hot giant planet HD 209458b (Liang et al. 2003; Moses et al. 2011; Venot et al. 2012) and HD 189733b (Line et al. 2010; Moses et al. 2011; Venot et al. 2012). Other studies explored CoRoT-2b, XO-1b and WASP-12b (Moses et al. 2012) and WASP- 12b (Kopparapu et al. 2012). Adopting some simplifi- cations, like the imposition of chemical equilibrium at an arbitrary lower boundary and the assumption of con- stant temperature in the entire atmosphere, Zahnle et al. (2009a,b) used their photochemical kinetic code, to per- form the first parameter space explorations in the study of hot giant planet atmospheres. Zahnle et al. (2009a,b) studied photochemistry of sulfur products for planets with different temperatures and showed photochemical 2 models for planets with three different temperatures: 800, 1000 and 1200 K, including variations in the eddy diffusion coefficient and different metallicities (0.7 ≤ [M/H] ≤ 1.7) for a planet located around HD 189733, respectively. Photochemical models in the literature, explain the chemistry for specific planets, but do not explore broader trends as a function of stellar flux and semimajor axis of hot planets and their atmospheric chemistry. In this pa- per we present a parameter space exploration, linking astrophysical observables like orbital distance and stellar spectral type with the change in mixing ratios of chemi- cal species in hot exoplanet atmospheres. Different ver- tical mixing in the atmospheres is also explored, show- ing possible degeneracies in the mixing ratios of chem- ical species. We present a grid of thermal profiles and photochemical mixing ratios of short orbital period H- dominated exoplanet atmospheres, in order to explore what to expect in future observations. Our grid shows planets with semimajor axis of 0.01, 0.015, 0.025, 0.05 and 0.1 AU orbiting different host stars (F, G, K and M). Section 2 describes our model, section 3 presents our results - for five known extrasolar planets (3.1) as well as for the whole grid (3.2). Section 4 discusses the influence of vertical mixing (4.1), clouds and hazes (4.2), the pro- file adopted for the thermal structure (4.3) and different elemental abundances (4.4) on our results and section 5 summarizes the paper. 2. THE MODEL The study of hot exoplanet atmospheres is a com- plex problem that requires understanding of the ther- mal structure, photochemistry and hydrodynamics of the planetary atmosphere. In this paper we developed a 1D model for calculating thermal atmospheric profiles, described in section 2.1 and link it to an atmospheric chemical model developed for studying WASP-12b and differences between the photochemical and thermochem- ical equilibrium models (see, Kopparapu et al. (2012)). The model takes the effect of photochemistry induced by stellar irradiation, considering equilibrium and disequi- librium chemistry as well as vertical diffusion and molec- ular diffusion into account. In this section we describe the model used for computing exoplanet atmospheres. 2.1. Thermal profile of the atmosphere For the atmospheric thermal structure, we developed a 1D model for highly irradiated exoplanets, which as- sumes a gray atmosphere in hydrostatic equilibrium (Hansen 2008; Guillot 2010; Burrows & Orton 2010). We focus on modeling the upper observable region of the at- mosphere where the opacities are low and therefore we neglect convection and assume a pure radiative model (see Section 4). The upper region of the atmosphere is characterized by temperatures significantly smaller than the effective temperature of the star, therefore we can assume that both radiation fields are decoupled and adopt character- istic mean opacities for each one. Assuming that the thermal contribution from the atmosphere in the visible is negligible, the model employs a two-stream approxi- mation, that treats the absorption of stellar irradiation (shortwave) and the subsequent reradiation (longwave) separately. Assuming isotropic radiation, the average temperature profile in the atmosphere as a function of the optical depth (τ ) is computed in equation 1 (Guillot 2010): + τ(cid:19)+ T 4 p = 3T 4 int 3 4 (cid:18) 2 +(cid:18) γ 3T 4 irr 4 f(cid:18) 2 3 + 1 γ√3 √3 − 1 γ√3(cid:19)e−γτ√3(cid:19) (1) where Tint is the internal temperature of the planet and Tirr is the irradiation temperature, that characterizes the flux received from the star and is proportional to the equilibrium temperature (Teq). The proportionality con- stant is f −1/4, where f has a value of 1 4 when assuming global average over the entire planetary surface. γ is the greenhouse factor defined as the ratio between mean opacities in the short and long wavelengths. γ < 1 im- plies that the stellar radiation is more transparent than the emerging one, while a value of γ > 1 implies a more opaque atmosphere. We follow Guillot (2010) and adopt a value of γ = 0.6q Tirr 2000 and a thermal mean opacity of κth = 10−2g/cm2 for the hottest planets. For the coolest planets in our grid (Teq ≃1000 K), we use γ=0.07 and the same value of κth (following Miller-Ricci & Fortney (2010) for GJ 1214b). In our model we divide the atmosphere in 100 layers between 10−7 and 100 bars, with the high pressure level chosen to show the transition between the region of chem- ical equilibrium and disequilibrium in the atmosphere. Finally we integrate the equation of hydrostatic equilib- rium at each step for the pressure structure calculation using equation 2: dP dz = −gρ (2) where the change in pressure (P ) with heigh (z) is equal to the density (ρ), multiplied by the gravity of the planet (g, set to be a constant in the thin upper region we are modeling). 2.2. Atmospheric chemistry calculations In our models, we assume a C/O=0.54 and solar elemental abundances, following Asplund et al. (2005). We calculate the equilibrium abundances at each temperature-pressure level, based on the minimization of Gibbs free energy, solving simultaneously the sys- tem of chemical equilibrium equations to derive equilib- rium mixing ratios as shown in White et al. (1958). We use those values as initial conditions and recalculate the mixing ratios using photochemistry and disequilibrium chemistry. To compute the photochemical mixing ratios, we use a 1D photochemistry code developed by Kopparapu et al. (2012) for WASP-12b that includes disequilibrium chem- istry driven by photochemistry, molecular diffusion and vertical mixing. The photochemistry code employs the reverse Euler method in order to solve the system of dif- ferential equations that determines the mixing ratios of all species at different heights in the atmosphere, adopt- ing as lower boundary condition the mixing ratios of the species at thermodynamic equilibrium and zero flux for all the long-lived species at the upper boundary. Our models include 19 chemical species (O,O(1D), O2, H2O, H, OH, CO2, CO, HCO, H2CO, CH4, CH3, CH3O, CH3OH, CH, CH2, H2COH, C, H2) in 179 reactions. The reaction list adopted in this code allows us to perform simulations in hot regimes, for 700.T.2800 K, note that kinetic rates might not be reliable at temperatures higher than 2800K. Our model grid covers planets with equilib- rium temperature from 2800 K to 700 K, cooler planets will be explored in a future paper. The complete list of chemical reactions can be found in Kopparapu et al. (2012) and for detailed information about the numerical analysis used in the code see Pavlov et al. (2001). Vertical mixing time scale is parametrized using the eddy diffusion coefficient (KZZ ), as usual in 1D pho- tochemical models (Line et al. 2010; Moses et al. 2011, 2012). In our calculations we use a nominal value of KZZ = 109 cm2/s, which is a value adopted by Moses et al. (2012) and explore the dependence of the results on this parameter in section 4.1. This coefficient is a big uncertain in 1D models and a different value can lead to significantly different results (see section 4.1). 3. RESULTS We first apply our model to five known extrasolar plan- ets in this temperature range and compare our results to the four atmospheric models published (section 3.1). We show the first model developed specifically for the re- cently discovered HD 97658b (section 3.1.5). Then, we show a grid of planet models, that can be used for hot planets ( 700< Teq <2800 K) with a solar composition H-dominated atmosphere (section 3.2). The planets modeled in our paper are shown schemat- ically in Figure 1, which shows the stellar effective tem- perature vs semimajor axis of the planets assuming a planetary albedo of 0.01 (see e.g., Rowe et al. (2008); Burrows et al. (2008); Sudarsky et al. (2000)). The in- verted triangles indicate the known planets modeled (sec- tion 3.1) and the dots the sample planets in the grid (section 3.2), with a color scale according to their equi- 4a2 (cid:19) librium temperatures (Teq = Tef f,⋆(cid:18) (1−A)R2 ⋆ 1 4 , where A is the planet albedo, set to 0.01 here). Hotter planets are shown in blue, cooler planets in yellow. Planets with temperatures below 700 K (e.g. a=0.05 and 0.1 around an M star, and a=0.1 AU orbiting a K star) or above 2800 K (e.g. a=0.01 AU orbiting a G star and a=0.01 and 0.015 around an F star) have equilibrium temper- atures which are beyond the temperature range of our model (see section 2.2). 3.1. Application of our model to known extrasolar planets We calculate thermal and photochemical models for five known exoplanets. We use the four published models to compare our results and validate our model. CoRoT- 2b, WASP-12b, XO-1b and HD 189733b fall within the temperature range of our model. Table 1 shows the main characteristics of these planets and references to recent atmospheric calculations for comparison to our results. Table 2 gives the corresponding stellar information. As seen in the tables, these four planets have semimajor axis 3 WASP 12b T=2800 K CoRoT 2b XO-1b HD189733b HD 97658b T=700 K 0.01 0.015 0.025 0.05 0.1 Semimajor axis (AU) Semi major axis (AU) 2800 2600 2400 2200 2000 1800 1600 1400 1200 1000 800 ) K ( e r u t a r e p m e t m u i r b i l i u q e y r a e n a P t l ) K ( e r u t a r e p m e t e v i t c e f f e r a l l t e S F star 7000 ) K ( 6500 6000 t e r u G star a r e p m e t 5500 e v i t K star c e 5000 f f e r a l l t e S 4500 4000 M star 3500 Fig. 1.- Semimajor axis vs. stellar effective temperature for the planets and stars adopted in this paper. The dotted lines show the approximate limits in planetary equilibrium temperature modeled. The dots and triangles represent the planets modeled shown in different color palette according to their equilibrium temperature with hotter planets shown in blue while cooler planets shown in yellow.The 5 known extrasolar planets modeled in this paper are indicated with inverted triangles. between 0.01 and 0.1 AU, orbit stars with effective tem- perature between 3800 and 7000 K and have a potentially H-dominated atmosphere. Cooler planets like GJ 1214b and GJ 436b were also studied (Miller-Ricci & Fortney 2010; Moses et al. 2013), but they have cool tempera- tures that put them outside the range of our study. We also applied our results to HD 97658b, whose discovery was recently announced (Dragomir et al. 2013), where no atmosphere models exist to compare to yet. For the four exoplanets with published photochem- ical models we use the same stellar flux libraries as in Kopparapu et al. (2012), to facilitate the compar- ison. HD 189733b is orbiting a K2V star, as tem- plate we use HD 22049 (following Segura et al. (2003); Kopparapu et al. (2012)). CoRoT-2b, WASP-12b and XO-1b orbit G0 stars, for these planets we use G0 spectra from Pickles stellar spectral flux library (Pickles 1998). For HD 97658b, there are no thermal or photochemi- cal models to compare to, therefore we use as input one of the stars used in our grid, with Tef f,⋆ =5000 K, R⋆ = 0.8R⊙ (Rugheimer et al. 2013). Figures 2 and 3 show the thermal structure and pho- tochemistry profiles, respectively, calculated using the models described in Sections 2.1 and 2.2 applied to HD 189733b, XO-1b, CoRoT-2b and WASP-12b. The ther- mal profile found for HD 97658b is shown in Figure 4(a) and its photochemical models are shown in Figure 4(b). In all Figures, the photochemical mixing ratios are shown as solid lines, while the chemical equilibrium values are plotted as dotted lines. Our results agree with published models (see table 1). Small differences in individual val- ues are due to slightly different chemical schemes, ther- mal profiles, different metallicity, UV fluxes used, and cross sections adopted. 3.1.1. HD 189733b HD 189733b is the most studied planet among the ones analyzed in this paper. As shown in Figure 2, its ther- mal model profile is isothermal in the upper atmosphere, with an equilibrium temperature of 1192 K, when adopt- 4 Planetary data of the known planets modeled. TABLE 1 Planet a (AU) Mp (MJ up) Rp (RJ up) Log10(g) Reference to recent atmospheric calculations HD 189733b XO-1b CoRoT-2b WASP-12b HD 97658b 0.031 0.04928 0.02809 0.02253 0.0796 1.14 0.918 3.270 1.139 0.0247 1.138 1.206 1.466 1.79 0.2088 3.341 3.211 3.548 3. 3.19 Line et al. (2011); Moses et al. (2011); Venot et al. (2012), this paper Moses et al. (2012), this paper Moses et al. (2012), this paper Kopparapu et al. (2012); Moses et al. (2012), this paper This paper TABLE 2 Stellar data of the known planets modeled. Star R⋆/R⊙ Tef f,⋆ (K) HD 189733 XO-1 CoRoT-2 WASP-12 HD 97658 0.756 0.934 0.902 1.63 0.703 5040 5750 5630 6300 5119 1e-07 1e-7 1e-06 1e-6 1e-5 1e-05 0.0001 0.0001 CoRoT 2b CoRoT2b 0.001 0.001 0.01 0.01 0.1 0.1 XO-1b X0-1b 1 1 HD189733b 10 10 HD189733b ) s r a b ( e r u s s e r P ) s r a b ( e r u s s e r P 1000 1000 1200 1200 1400 1400 1200 1600 1800 1600 1800 2000 2000 2200 2200 2400 2400 2600 2600 2800 2800 Temperature (K) Temperature (K) Fig. 2.- Temperature-pressure profile calculated for HD 189733b, XO-1b, CoRoT-2b and WASP-12b. ing a 0.01 Bond albedo, reaching ∼ 1500 K at about 100 bars. Our thermal profile is in good agreement with the profiles calculated by other authors like Burrows et al. (2008) and Showman et al. (2009) this last authors per- formed detailed simulations with a global 3D model cou- pled with a nongray cloud-free radiative transfer code. Other authors (Line et al. 2010; Venot et al. 2012) use a similar T-P profile. HD 189733b mixing ratios are shown in Figure 3(a), with CO, H2O and H being the major gases in its atmo- sphere. The mixing ratios of H2O varies between 10−3 and 10−4 until it dissociates at 5 × 10−6 bars. The CH4 mixing ratio varies between 10−3 and 10−8 for P> 0.001 bars. OH and O show mixing ratios less than 10−9 for P>0.001 bars. CO mixing ratio vary between 10−3 and 10−4. The CO2 mixing ratio is about ∼ 10−9 for values above 5 × 10−6 bars, for lower pressures it increase (due to water photolysis) and then decrease again due to its own dissociation. When comparing with other studies we find that CO mixing ratio has similar values in all three comparison studies, but the pressure when the mixing ratio starts to decrease as a consequence of water photolysis de- pends on the water dissociation level in the atmosphere. For H2O, our model agree with models by Moses et al. (2011) and Venot et al. (2012). Note that Line et al. (2011) find the dissociation pressure at a lower pres- sure of 10−8. CO2 decreases due to its own photoly- sis at below 5 × 10−6 bars in our model, in agreement with Moses et al. (2011) and Venot et al. (2012). Note that Line et al. (2011) find dissociation at lower pressure of 10−8 bar for this molecule. We find CH4 levels be- tween 10−3 and 10−5 down to pressures of 10−4 bars in agreement with Moses et al. (2011). Other models find methane levels between 10−4 to 10−5 down to pressures of 10−8 bars (Line et al. 2011) and 10−5 and 10−6 down to pressures of 10−4 bar (Venot et al. 2012). WASP12b WASP 12b 3.1.2. XO-1b The thermal profile for XO-1b is shown in Figure 2 and presents a isothermal profile with ∼1200 K (assuming an albedo of 0.01) in the upper atmosphere and tempera- tures close to 1800 K in the optically thick region, in agreement with models adopted by Moses et al. (2012). XO-1b photochemical mixing ratios are shown in Fig- ure 3(b), showing H, CO, H2O and CH4 as the major gases in the atmosphere. H2O has a mixing ratio between 10−3 and 10−4 until its dissociation, at ∼ 5 × 10−5 bars. CO2 shows a mixing ratio between 10−8 and 10−7. Wa- ter photolysis affects the mixing ratios of other molecules (see section 3.2) and for that reason CO2 has a local max- imum at ∼ 5 × 10−5 bars, below that pressure it starts to decrease due to its own photolysis. CO, shows a sim- ilar behavior and a mixing ratio between 10−4 and 10−3 until it starts to decrease at ∼ 10−7 bars. H mixing ratio is also affected by water photolysis, becoming the major gas in the upper atmosphere. CH4 reaches the quench level at a pressure close to 1 bar and starts to decrease at 10−4 bars. Our results agree with Moses et al. (2012), with mi- nor differences caused by different metallicities and stel- lar flux used in both models. While Moses et al. (2012) adopted a metallicity of Fe/H=1.5× solar and use the solar spectral flux as an input in their photochemical calculations, we use a solar metallicity and our incoming stellar flux is from a G0 star whose flux in the UV is higher than the Sun (following Kopparapu et al. (2012), figure 3). As a consequence, H2O starts to dissociate at 10−6 bars for Moses et al. (2012) and at∼ 5 × 10−5 in our simulations. This change in the H2O dissociation also bring minor changes in the CO2 and CO mixing ratios, but does not influence the overall agreement sub- stantially. 3.1.3. CoRoT-2b CoRoT-2b thermal profile is shown in Figure 2. This planet has an equilibrium temperature of 1536 K (as- 1e-06 1e-05 CO2 H2O H2 1e-06 1e-05 0.001 0.0001 ) s r a b ( e r u s s e r P 0.01 0.1 HD 189733b O CH4 H CO OH 0.001 0.0001 ) s r a b ( e r u s s e r P 0.01 0.1 5 CO2 H2O CH4 H2 H O OH CO 1 10 1e-20 1e-15 1e-10 1e-05 Volume Mixing ratio 1 XO-1b 10 1 1e-20 1e-15 1e-10 1e-05 Volume Mixing ratio (a) 1e-06 1e-05 H2 1e-06 1e-05 0.001 0.0001 ) s r a b ( e r u s s e r P 0.01 0.1 CO2 CO H H2O CH4 OH O 0.001 0.0001 ) s r a b ( e r u s s e r P 0.01 0.1 1 H2 (b) CO2 H2O CO O H OH CH4 1 CoRoT-2b 10 1e-20 1e-15 1e-10 1e-05 Volume Mixing ratio 1 WASP12b 10 1 1e-20 1e-15 1e-10 1e-05 Volume Mixing ratio 1 (c) (d) Fig. 3.- Mixing ratios vs. pressures for HD 189733b (Figure 3(a)), XO-1b (Figure 3(b)), CoRoT-2b (Figure 3(c)) and WASP-12b (Figure 3(d)). The solid lines show the results when disequilibrium chemistry due to vertical mixing and photochemistry is included, while the dotted lines show the mixing ratios in equilibrium. suming an albedo of 0.01). Our thermal profile is sim- ilar to retrieval models (Madhusudhan & Seager 2009; Madhusudhan et al. 2011; Madhusudhan 2012), but dif- fers from the ones adopted by (Moses et al. 2012). Figure 3(c) shows the photochemical and equilibrium mixing ratios of CoRoT-2b. H, CO and H2O are the most abundant species in the model atmosphere. H2O has a mixing ratio between 10−3 and 10−4 until it gets pho- tolyzed at ∼ 10−4 bars. Water photolysis affects other molecules like CO2 which has a local maximum due to H2O dissociation at ∼ 10−4 bars. CO mixing ratio is be- tween 10−3 and 10−4 in our models and methane quench level is at 1 bar and its abundance is lower than in the previous cases, because this planet is hotter than HD 189733b and XO-1b. Moses et al. (2012) adopted a metallicity of 0.5× so- lar for the case of solar C/O. They also use as input the solar flux scaled at CoRoT 2b semimajor axis, while we use solar metallicity and the same G0 star as for XO-1b. This difference in the stellar flux explains the difference in the mixing ratios of the major molecules. H2O pho- tolysis occurs at a lower pressure for Moses et al. (2012), at 10−6 bars. This also affects the mixing ratios of other molecules as CO2 whose local maximum consequence of water dissociation occurs at 10−7 bars in these models. CO and methane present minor changes. 3.1.4. WASP-12b WASP-12b is the hottest planet among the known planets modeled in this paper. It has an equilibrium temperature of 2577 K (adopting an albedo of 0.01) and its thermal profile is shown in Figure 2. Our ther- mal profile for WASP-12b agrees with retrieval mod- els (Madhusudhan & Seager 2009; Madhusudhan et al. 2011; Madhusudhan 2012), but differs from to the ones adopted by other authors (Kopparapu et al. 2012; Moses et al. 2012). Mixing ratios of WASP-12b are shown in Figure 3(d). H, CO and H2O are the major gases in the atmosphere, while CH4 has a very low mixing ratio as a consequence of the high temperature of the planet and therefore is not expected to have significant abundance at the ob- servable pressure levels. H2O has a mixing ratio between 10−3 and 10−4 for P> 10−4 bars and start to dissociate at 5 × 10−4 bars. As seen in the previous cases, H2O dissociation also affects the mixing ratios of the rest of 6 the major species in the upper atmosphere. CH4 shows a mixing ratio less than 10−5, CO a mixing ratio between 10−3 and 10−4 and CO2 a mixing ratio close to 10−9 until its dissociation at 10−4 bars. When comparing with other authors, we notice more differences than the other three cases because the initial assumptions and thermal profiles of Moses et al. (2012); Kopparapu et al. (2012) are different from ours. To compare between models we keep metallicity and C/O ratio constant in our grid as explained in section 2.2. Moses et al. (2012) adopted 3 times solar metallicity for this planet which influences in the comparison. Another difference is the eddy diffusion coefficient adopted, which is KZZ = 1010 cm2/s for Moses et al. (2012), and KZZ > 1010 cm2/s for Kopparapu et al. (2012) in the upper at- mosphere and close to KZZ = 109 cm2/s for P> 10−2 bars. Both values are higher than KZZ = 109 cm2/s adopted in our paper. Despite these differences, the CO profile is very similar in the three cases, while there are some differences in the H2O, CH4 and CO2 profiles. H2O has a similar mixing ratio in the three cases, but its dis- sociation starts at a lower pressure in our model, as a consequence of the hotter thermal planetary profile. The CO2 mixing ratio profile by Kopparapu et al. (2012) is very similar to the one we find, except in the upper at- mosphere, where it is affected by the differences in the H2O photolysis. Moses et al. (2012) found a larger mix- ing ratio of CO2, related to the larger metallicity they adopt in their model. Our methane mixing ratio profile is very similar to the one by Moses et al. (2012), being ∼ 10−5 at 100 bars, and decreasing close to the chemical equilibrium curve until ∼ 10−4 bars. As a consequence of their cooler thermal structure, Kopparapu et al. (2012) have a larger mixing ratio in the region between 10−2 to 10−4 bars, which is close to 10−10 while Moses et al. (2012) and our results show mixing ratios of ∼ 10−14. 3.1.5. HD 97658b: a planet with an atmosphere in thermochemical disequilibrium Dragomir et al. (2013) announced the discovery of HD 97658b, which is a transiting planet of 7.862 M⊕, radius of 2.34 R⊕ and whose host star has T=5119 K and R = 0.703 R⊙. The mass and radius of HD 97658b implies a low density, which suggests that this planet might have an atmosphere of volatiles. Howe & Burrows (2012) made some predictions about HD 97658b's atmospheric observables, by assuming sim- ilar atmospheric properties as the cool GJ 1214b. They didn't present photochemical models specifically for this planet. Here, we assume that it has an atmosphere with significant H/He (and solar elemental abundances) and model this planet and its potential atmosphere. This is a first model for HD 97658b's atmosphere. Note that other C/O ratios, albedos and metallicities will influence these results. Figure 1 shows the location of this planet in our grid, close to the lower temperature limit. Figure 4(a) shows the thermal profile model for HD 97658b, with an equilibrium temperature of Teq = 731 K when adopting a 0.01 Bond albedo. The photochemical mixing ratios for this planet are shown as solid lines in Figure 4(b). As dotted lines we show the equilibrium values, where the model shows a thermochemical dise- quilibrium for pressures P<1 bar. Our model shows that ) s r a b ( e r u s s e r P ) s r a b ( e r u s s e r P 1e-7 1e-07 1e-6 1e-06 1e-5 1e-05 0.0001 0.0001 0.001 0.001 0.01 0.01 0.1 0.1 1 1 10 10 100 500 500 1e-6 1e-06 1e-5 1e-05 0.0001 0.0001 0.001 0.001 0.01 0.01 0.1 0.1 1 1 10 10 100 ) s ) r s a r b a ( b P ( e r u s s e r P HD 97658b 1000 1000 1500 1500 2000 2000 2500 2500 3000 3000 3500 Temperature (K) Temperature (K) (a) HD 97658b O CO2 H CH4 CO OH H2 H2O 1e-20 1e-20 1e-15 1e-15 1e-10 1e-10 1e-05 1e-5 1 1 Volume Mixing Ratio Volume Mixing ratio (b) Fig. 4.- Thermal profile (Figure 4(a)) and mixing ratios of the major chemical species (Figure 4(b)) for HD 97658b . The solid lines show the results when disequilibrium chemistry due to vertical mixing and photochemistry is included, while the dotted lines show the mixing ratios in equilibrium. A value of KZZ = 109 cm2/s was adopted. H2, H2O, CO, H and CH4 are the most abundant gases in HD 97658b's atmosphere. H2O has a mixing ratio between 10−3 and 10−4, until it starts to dissociate. The planet is relatively far from its host star (a=0.0796 AU), when compared with other planets considered in this paper, receiving a lower UV flux from its host star. As a consequence, H2O starts to dissociate relatively high in the atmosphere (∼ 10−6 bars) and its dissociation is not as efficient as for planets closer to their host stars. Since this is a cool planet, CH4 abundance is relatively high, with a mixing ratio between 10−4 and 10−5, un- til it starts to decrease at 10−5 bars. CO2 is relatively abundant, with a relatively constant profile close to 10−4 down to 10−5 bars, that increases as a side effect of water photolysis (see 3.2). CO has a mixing ratio between 10−3 and 10−4 in all the pressure range explored. H replaces H2 as the major species in its atmosphere at P< 5× 10−6 bars. We adopted an eddy diffusion coefficient of KZZ = 109 cm2/s here. Section 4.1 discusses how different val- ues influence the chemical abundance of major species in the upper atmosphere. 10 10 exoplanets.org 7/10/2013 1 1 0.1 0.1 ] s s a ) s M s a r M e t i r e t i p u p J u [ J s ( s s a s a M m m e n u m a P n M t l i i 0.01 0.01 -3 0.001 10 2.5 2.5 3.0 3 3.5 3.5 Log (g) Surface Gravity 4.0 4 Fig. 5.- Planetary mass vs. surface gravity for all the plan- ets with calculated values in the literature (http://exoplanets.org). The dotted line shows the intermediate value (log(g)=3.4) adopted in our grid. 3.2. Grid of planetary atmospheres between 0.01 and 0.1 AU around MKGF host stars We use our model to build a grid of mini-Neptune and giant planet atmospheres to explore the characteristics of a wide range of exoplanets. This grid links astrophysi- cal observables with exoplanets atmospheric composition and can be used as a reference for current and future ob- servations and retrieval analysis. This grid is applicable to any hot exoplanet ( 700<T<2800 K) with H-dominated atmosphere of solar composition and a surface gravity of log(g) ≃ 3.4. This value was chosen as a mean value that can be applied to model mini-Neptunes or giant planets, as shown in Fig- ure 5, where the surface gravity is shown as a function of the planetary mass1. We assume that the primary atmosphere is mainly composed of H and He, with elemental abundances of solar composition. Our grid models the thermal profile and mixing ratios of planets between 700 K and 2800 K, what corresponds to planets located at 0.01, 0.015, 0.025, 0.05 and 0.1 AU of their hosts stars, assuming a planetary albedo of 0.01. We use the stellar fluxes for an F (Tef f,⋆ =7000 K, R⋆ = 1.5R⊙), G (Tef f,⋆ =6000 K, R⋆ = 1.1R⊙) and K stars (Tef f,⋆ =5000 K, R⋆ = 0.8R⊙) calculated by Rugheimer et al. (2013), based on stellar models taken from the ATLAS synthetic spectra (Kurucz 1979) com- bined with UV observations from the International Ul- traviolet Explorer (IUE) archive 2. For the M star (Tef f,⋆ =3800 K, R⋆ = 0.62R⊙), we use the emission flux of an inactive M star from the dust model devel- oped by Allard et al. (2001). We use the luminosity of a main sequence star based on its effective temperature and calculated the flux received at the top of the planet atmosphere using the inverse square law of the distance to get the correct flux for each planet's semimajor axis. Note that the inactive Mstar model we use assumes no chromospheric activity (Allard et al. 2001), what might overestimate the differences between active and inactive 1 Figure 5 was made using the data published in http://exoplanets.org/ (Wright et al. 2011) 2 http://archive.stsci.edu/iue/ 7 stars if some chromospheric activity is present in realis- tic inactive Mstars (see also Seager et al. (2013)). Many main sequence M stars present strong chromospheric ac- tivity that produces high-energy radiation. We will ad- dress the effect of activity on a planet's spectra in a future paper. 3.2.1. Grid of thermal profiles Figure 6 show the grid of atmospheric temperature vs. pressure. Planetary semimajor axis are indicated at the top and the different host star types are shown on the right side of the grid. As expected, temperatures are cooler when the planets orbit further away from the star (see Spiegel et al. (2010) for similar results for plan- ets around a solar-type star), with the planet located at 0.025 AU from a M star being the coolest and the one orbiting a F star at 0.025 AU the hottest planet consid- ered in our grid (assuming the same albedo of 0.01 for all planets). Figure 6 shows that the planetary thermal profiles typically show an isothermal layer in the upper atmosphere, which is optically thin and the most rele- vant part when considering the gases that are observable in the spectra. While the incident stellar flux is absorbed, the atmosphere goes from the optically thin to optically thick regime in a transition region, followed by a final, op- tically thick, nearly isothermal profile below that. Con- vection is likely to be the most efficient energy transport mechanism at these depths, which is not consider in this paper (see discussion in Section 4.3), therefore we do not consider our results below ∼1 bar reliable. Note that our radiative models give similar results as observations and models by other groups (see section 3.1), even at higher pressures. Chemical kinetics timescales required to maintain the species in equilibrium with each other are shorter than vertical mixing for high temperatures and pressures. Therefore, chemical equilibrium is reached faster at high pressures, deep in the atmosphere of these gaseous plan- ets. Chemical disequilibrium dominates the upper atmo- sphere of highly irradiated planets, where densities are low and disequilibrium processes have shorter timescales. Furthermore, in highly irradiated atmospheres stellar UV radiation photodissociate molecules in the upper atmo- sphere, causing escape and recombination of atoms. 3.2.2. The effect of UV flux only on a exoplanet's atmosphere In order to study the effect of stellar flux alone on photochemical mixing ratios in the atmosphere, we per- formed simulations using the same thermal profile and vertical mixing, but adopting different stellar fluxes. Fig- ure 7 shows the mixing ratios of different species as a function of pressure for planets with the same thermal structure and eddy coefficient (KZZ = 109cm2/s), but different stellar fluxes. For the thermal structure, we use the temperature-pressure profile of a planet located at 0.025 AU of a K star as an example (row 3, column 3 in Figure 6), the stellar fluxes are the same adopted in the grid calculations. The different photochemical mixing ratios found in each case are shown in different colors in Figure 7, or- ange (F star), red (G star), brown (K star) and black (M star). H2O and CH4 are the species most affected by the stellar UV flux and the resulting photochemistry. 8 ) s r a b ( e r u s s e r P ) s r a b ( e r u s s e r P 0.01 AU 0.015 AU 0.025 AU 0.05 AU 0.1 AU 1e-05 1e-06 1e-6 1e-5 1e-4 0.001 0.01 ) s r a b ( 0.0001 0.001 0.01 e r u s s e r P 0.1 0.1 1 10 10 1 ) s r a b ( e r u s s e r P Teq=2607.6 K 1e-06 1e-05 0.0001 0.001 0.01 0.1 1 10 ) s r a b ( e r u s s e r P 1e-06 Teq=1843.89 K 1e-05 Teq=1303.83 K ) s r a b ( e r u s s e r P 0.0001 0.001 0.01 0.1 1 10 1e-06 1e-05 1e-6 1e-5 1e-4 0.001 0.01 0.0001 0.001 0.01 ) s r a b ( e r u s s e r P 0.1 0.1 1 10 1 10 ) s r a b ( e r u s s e r P 100 1e-06 500 1000 1500 2000 2500 3000 Temperature (K) 100 1e-06 3500 500 Teq=2468.6 K ) s r a b ( e r u s s e r P 1e-05 0.0001 0.001 0.01 0.1 1 10 Teq=1914.06 K 1e-05 ) s r a b ( e r u s s e r P 0.0001 0.001 0.01 0.1 1 10 1e-06 1e-05 1e-6 1e-5 1e-4 0.001 0.01 0.0001 0.001 ) s r a b ( e r u s s e r P 0.01 0.1 0.1 1 10 1 10 100 1e-06 500 1000 1500 2000 2500 3000 Temperature (K) 100 1e-06 3500 500 1000 1500 2000 2500 3000 Temperature (K) 100 1e-06 3500 500 Teq=2150.76 K ) s r a b ( e r u s s e r P 1e-05 0.0001 0.001 0.01 0.1 1 10 Teq=1808.4 K 1e-05 Teq=1360.26 K 1e-05 ) s r a b ( e r u s s e r P 0.0001 0.001 0.01 0.1 1 10 ) s r a b ( e r u s s e r P 0.0001 0.001 0.01 0.1 1 10 1000 1500 2000 2500 3000 100 1e-06 3500 500 1000 1500 2000 2500 3000 3500 Temperature (K) Teq=1353.44 K 1e-05 Temperature (K) Teq=957.03 K ) s r a b ( e r u s s e r P 0.0001 0.001 0.01 0.1 1 10 1000 1500 2000 2500 3000 100 3500 500 1000 1500 2000 2500 3000 3500 Temperature (K) Teq=961.85 K 1000 1500 2000 Temperature (K) 2500 3000 Temperature (K) 1000 1500 2000 2500 3000 Temperature (K) 100 1e-06 3500 500 1000 1500 2000 2500 3000 Temperature (K) 100 1e-06 3500 500 1000 1500 2000 2500 3000 100 3500 500 1000 1500 2000 2500 3000 3500 Temperature (K) Teq=910.09 K 1000 1500 2000 Temperature (K) 2500 3000 Temperature (K) 100 1e-06 500 1e-05 1e-6 1e-5 1e-4 0.001 0.01 ) s r a b ( 0.0001 0.001 0.01 e r u s s e r P 0.1 0.1 1 10 1 10 Teq=1439 K 1e-05 Teq=1173.7 K 1e-05 ) s r a b ( e r u s s e r P 0.0001 0.001 0.01 0.1 1 10 ) s r a b ( e r u s s e r P 0.0001 0.001 0.01 0.1 1 10 F G K M 100 500 1000 1500 2000 2500 3000 100 3500 500 1000 1500 2000 2500 3000 100 3500 500 1000 1500 2000 2500 3000 3500 500 1000 1500 2000 2500 3000 Temperature (K) 500 1000 1500 2000 Temperature (K) 2500 3000 500 1000 1500 2000 2500 3000 Temperature (K) Temperature (K) Temperature (K) Temperature (K) Fig. 6.- Grid of planetary temperature-pressure profiles for H-dominated planets as a function of their semimajor axis. The 5 columns present the distance of the planet from their host star for 0.01, 0.015, 0.05, 0.025, and 0.1AU, from left to right respectively. The different rows show the model results for planets orbiting different host stars. Planets around inactive F (Tef f,⋆ =7000 K), G (Tef f,⋆ =6000 K), K (Tef f,⋆ =5000 K) and M (Tef f,⋆ =3800 K) stars are shown in rows 1, 2, 3 and 4, respectively. An albedo of 0.01 was adopted for all planets to calculate the equilibrium temperature with the individual results shown in each panel. The effect of vertical mixing and photochemistry com- pete in the atmosphere. Here we investigate the effects of photodissociation, the effect of vertical mixing is ex- plored in section 4.1. As expected, at lower pressures in the atmosphere (P< 10−4 bars) photolysis has shorter timescales than vertical mixing and its effects are even larger when the planet is irradiated by a higher UV flux and therefore mixing ratios change when going from an M to an F star. When larger amount of photons are irra- diating the atmosphere, they reach lower altitudes, and therefore dissociation starts at higher pressures. Disso- ciation of H2O starts at 10−4 bars when the planet is irradiated by an F star and at ∼ 10−7 bars for an M star case. The differences are even larger for methane, which starts to dissociate at 0.01 bars for the F star case and at ∼ 10−7 bars for the M star. CH4 dissociates creating CH3, H and CH2 and H, therefore methane dissociation creates H. Large amounts of H are also created due to water photolysis. Dissociation of H2O also creates OH, which reacts with H2, destroying it. This effect is impor- tant in the upper atmosphere, with a H2 mixing ratio of 10−2 for a planet irradiated by an F star and close to 1 for a planet irradiated by an M star at ∼ 10−7 bars. Due to this effect, the ratio of H to H2 is higher when going from low to high UV stellar emission (i.e. from M to F stars). CO combines with OH (product of H2O dissociation), forming CO2, that has a local maximum when this happens, at the pressure where water starts to dissociate, then it decreases due to its own dissociation reaching lower mixing ratios for the case of the planet irradiated by the F star. Another effect of water dissoci- ation is the production of O by the reaction between H and OH, which reaches higher mixing ratios in the atmo- sphere, reaching a mixing ratio close to 10−5 at different pressures, according to the pressure levels where water starts to dissociate in the atmosphere. Water photolysis dominates the chemistry in the upper atmosphere, being the effects larger when going from the M to the F star. In summary, high stellar UV flux (e.g. F star) leads to in- crease of mixing ratios of H, OH and O, a small increase in CO2 and the destruction of H2. For planets around stars with low UV fluxes (e.g. inactive M stars) H2, H, H2O, CO and CH4 are the most abundant chemicals in the observable part of the planet's atmosphere. 3.2.3. Grid of photochemical models The volume mixing ratios of the major gases in the at- mosphere as a function of pressure in the grid are shown H2O CO2 1e-6 1e-5 0.0001 ) s r a b ( P 0.001 0.01 9 0.1 1 10 M star K star G star F star M star K star G star F star H M star K star G star F star 1e-6 1e-5 0.0001 ) s r a b ( P 0.001 0.01 0.1 1 10 1e-6 1e-5 0.0001 ) s r a b ( P 0.001 0.01 0.1 1 10 1e-6 1e-5 0.0001 ) s r a b ( P 0.001 0.01 0.1 1 10 M star K star G star F star CO M star K star G star F star H 2 M star K star G star F star OH M star K star G star F star CH4 M star K star G star F star O 1e-15 1e-10 1e-5 1 1e-15 1e-10 1e-5 1 Volume Mixing ratio Volume Mixing ratio Fig. 7.- Pressure vs. volume mixing ratios for different chemical species. To study the effect of stellar UV flux only, we use the same thermal profile for a planet located at 0.025 AU from a K star (row 3, column 3, in Fig. 6), but change the stellar UV fluxes according to the stellar type: M (black), K (brown), G (red) and F (orange) stars. We adopted KZZ = 109 cm2/s. in Figure 8, where the stellar types of the host star and semimajor axis of the planets are shown on the right and top of the grid, respectively. In this figures, dif- ferent colors show the major chemical species in dotted lines when they are driven by equilibrium chemistry only and in solid lines when disequilibrium chemistry is also taken into account. As expected at high temperatures and pressures, equilibrium mixing ratios for all species are maintained in the atmosphere (below ∼1 bar). Equi- librium levels can be maintained even at lower pressures (P< 10−5 bars) for the hottest planet around inactive M stars, due to low UV flux emission. It is also shown in the models of the hottest planets in our grid (with equi- librium temperatures close to 2800 K), where the mixing ratios of all chemical species are close to the equilibrium. This is expected because equilibrium is reached faster at high temperatures and pressures, where the chemical equilibrium timescales (τeq) are shorter than disequilib- rium processes (τdis,eq). Cooler temperatures and low pressures favor disequilibrium (τeq > τdis,eq), which is relevant for high altitudes in the atmosphere. There- fore, the quench level where τeq ∼ τdis,eq, is a function of 10 the temperature, pressure and vertical mixing and will not occur at the same temperature and pressure for all chemical species, because it depends on the time scale for the fastest reactions that produce and destroy the molecule (Fegley & Prinn 1985). For example, methane departs from equilibrium deeper in the atmosphere than any other chemical species in our models for all the plan- ets of our grid. As seen in Figure 7, high in the atmosphere the UV photons coming from the star are absorbed and the re- sulting photochemistry dominates the chemical compo- sition of the planet's atmosphere. One of the main ef- fects is the production of large amounts of H and OH due to the photolysis of H2O, followed by a destruc- tion of molecular hydrogen due to its reaction with the produced OH (see e.g. Liang et al. (2003); Moses et al. (2011); Kopparapu et al. (2012)). Therefore, while large amounts of H are created, H2 is destroyed and replaced by H as the main atmospheric species high in the atmo- sphere. This effect is larger when the photolysis of H2O is larger. Therefore, as seen in Figure 8, there is more H and less H2 when the planets are closer to the star, being exposed to more UV radiation. CO is the dominant carbon compound in all the plan- etary atmospheres in the grid. Its abundance is set by thermochemical equilibrium, except high in the atmo- sphere (P< 10−5) where it has a local maximum product of water photolysis, except for the hottest planets in our grid, where equilibrium chemistry dominates even at low pressures. Methane is more stable at lower temperatures and therefore its abundance increases when going from hot to cooler planets in the grid. In addition, looking at planets with the same semimajor axis but orbiting dif- ferent stellar types, CH4 abundance increases also with decreasing stellar temperature, dominated by the effects of decreasing stellar UV. Methane is less efficiently recy- cled and therefore, even more affected by photodissocia- tion. CH4 is not present in significant amounts high in the atmosphere (above 10−5 bars) in the planets in our grid, with the exception of the cool planets around the M star, which are exposed to very low UV radiation and are cool enough to maintain a large abundance of methane in the atmosphere. The other exceptions to this are the cases of a higher vertical mixing in the atmosphere (see section 4.1). As seen in Figure 8, the mixing ratio of atomic O in- creases at high altitudes in all models, due to its pro- duction through water photolysis in the atmosphere. Atomic oxygen is also lost due to the backwards reac- tion, but can be transported up to very low pressures, where it is distributed over large radial distances due to hydrodynamic winds and other processes present in the thermosphere, as explained by several authors (see e.g., Murray-Clay et al. (2009); Lammer et al. (2009)). Low planetary temperatures (e.g. for large orbital distances) show an increase of CH4. For hot planets (e.g. with small orbital distances) H, H2 and CO are the most abundant chemicals in the observable part of the planet's atmo- sphere. 3.3. Equilibrium vs. Disequilibrium chemistry Figures 3, 4(b) and 8 show the comparison between models that use equilibrium (dotted lines) vs. non equi- librium (solid lines) chemical processes in exoplanet at- mospheres. The chemistry in mini-Neptune and giant exoplanet atmospheres is driven by disequilibrium pro- cesses (vertical mixing, photochemistry, molecular dif- fusion) for all planets with temperatures (T / 2500K) in our grid, with differences of several orders of mag- nitude to the corresponding equilibrium chemistry cal- culations. This occurs at pressures of P/ 10 bars, for all the planets in our grid including the known exoplan- ets HD189733b, XO-1b, CoRoT-2b (Figures 3(a), 3(b), 3(c)) and HD 97658b (Figure 4(b)), except for the hottest planets with T > 2500K, where the entire atmosphere is dominated by equilibrium chemistry: the planet at 0.015 AU around an G star, the planet at 0.025AU around an F star. WASP-12b is dominated by equilibrium chemistry for P¡0.001 bars (Figure 3(d)). Equilibrium chemistry dominates deeper regions (P' 10 bars) for all planets modeled. Only regions with P/ 10 bars are observed re- motely, therefore, we can not ignore the effects of disequi- librium chemistry when exploring the observable species in these exoplanet atmospheres. Figure 8 shows that for all planets in the grid dise- quilibrium chemistry lowers the abundance of H2, H2O and CO and increases the abundance of CH4, H, O and OH in the planet's atmosphere. Ignoring disequilibrium chemistry would severely overestimate the abundance of H and H2O and underestimate the abundance of CH4 (for the intermediate atmosphere, it is dissociated in the upper atmosphere), H, O and OH in the observable pres- sure region for planets with T / 2500K. Figures 3 and 4(b) show that using equilibrium chem- istry only would overestimate the abundance of H2, H2O and CO for the know exoplanets HD 189733b, XO-1b, CoRoT-2b, WASP-12b and HD 97658b. Note that the effects are smaller for WASP-12b because it is a hotter planet with a Teq of 2577 K (for A=0.01). Especially the abundance of CH4, H2O, H and H2 are critically depen- dent on disequilibrium chemistry as shown in Figures 3 and 4(b). 4. DISCUSSION Here we discuss some model parameters that influence our results, vertical mixing in section 4.1, clouds and hazes in section 4.2, our assumption of radiative profile in section 4.3 and elemental abundances in section 4.4. 4.1. Effect of vertical mixing The eddy diffusion coefficient is included in 1D pho- tochemical models (Line et al. 2010; Moses et al. 2011, 2012), to represent the vertical mixing processes that occur in planetary atmospheres. This coefficient is hard to determine observationally or through experi- ments. Some authors use it as a free parameter (see e.g., Moses et al. (2012)). Others, adopt a coefficient propor- tional to a globally averaged wind profile from GCM's results, where the proportional factor is the scale height (Line et al. 2010; Moses et al. 2011). In a recent paper, Parmentier et al. (2013) compared 1D and 3D models and derived a KZZ value that represents the averaged tracer profiles for HD 209458b. These recent results are two orders of magnitude smaller than previous ones ob- tained using the root mean square of the vertical veloc- ity. These new results suggest that mixing in hot-Jupiter atmospheres is potentially not as strong as previously 0.01 AU 0.015 AU 0.025 AU 0.05 AU 0.1 AU 1e-06 1e-05 0.0001 ) s r a b ( P 0.001 0.01 0.1 1 10 1e-06 1e-05 0.0001 0.001 0.01 0.1 1 10 CH4 CO2 CO H2 ) s r a b ( P H H2O O OH ) s r a b ( e r u s s e r P 0.0001 0.001 1e-06 1e-05 1e-6 1e-5 1e-4 0.001 0.01 0.1 1 10 ) s r a b ( P 1 10 0.01 0.1 CO2 H2O CH4 O CO H2 ) s r a b ( P H OH 1e-06 1e-05 0.0001 0.001 0.01 0.1 1 10 CO2 O H2O CH4 OH CO H2 ) s r a b ( P H 1e-20 1e-15 1e-10 1e-05 1 1e-20 1e-15 1e-10 1e-05 Volume Mixing ratio CO2 OH H2O CO Volume Mixing ratio CO2 H2O CO CH4 O OH H2 ) s r a b ( P H 1e-06 1e-05 0.0001 0.001 0.01 0.1 1 10 100 1 1e-06 Volume Mixing ratio CO2 CH4 O CO 1e-05 H2O H 0.0001 0.001 0.01 H2 ) s r a b ( P CH4 1e-05 CO CH4 O 1e-20 1e-15 1e-10 Volume Mixing ratio CO2 OH 0.1 O OH 1 10 1e-06 1e-05 0.0001 0.001 0.01 0.1 1 10 1e-06 1e-05 0.0001 0.001 0.01 0.1 1 10 CO2 CO CH4 H H2 O OH H2O 1e-20 1e-15 1e-10 1e-05 1 Volume Mixing ratio CO CH4 O CO2 OH H H2 H2O 1 H H2 ) s r a b ( P 1 1e-20 1e-15 1e-10 1e-05 1e-20 1e-15 1e-10 1e-5 Volume Mixing ratio 1 1 Volume Mixing Ratio H H2 H2O ) s r a b ( e r u s s e r P ) s r a b ( e r u s s e r P 1e-20 1e-15 Volume Mixing ratio 1e-10 CO2 1e-05 1 1e-20 1e-15 1e-10 1e-05 0.001 1e-05 1e-06 0.0001 1e-6 1e-5 1e-4 0.001 0.01 0.1 1 10 ) s r a b ( P 1 10 0.01 0.1 CH4 CO H OH ) s r a b ( P CO2 O H2O 1e-06 1e-05 H2 0.0001 0.001 0.01 0.1 1 10 CH4 H2O O OH CO H2 ) s r a b ( P H 1e-06 1e-05 0.0001 0.001 0.01 0.1 1 10 0.001 1e-05 1e-06 0.0001 1e-6 1e-5 1e-4 0.001 0.01 0.1 1 10 ) s r a b ( P 1 10 0.01 0.1 1e-20 1e-15 1e-10 1e-05 1 1e-20 1e-15 1e-10 1e-05 1 1e-20 1e-15 1e-10 1e-05 1 1e-20 1e-15 1e-10 1e-05 Volume Mixing ratio H 1e-06 Volume Mixing ratio CO2 O H2O O CO2 OH CH4 CO 1e-05 0.0001 0.001 0.01 H2 ) s r a b ( P 0.1 1 10 OH H CH4 H2O H2 CO ) s r a b ( P 1e-06 1e-05 0.0001 O 0.001 0.01 0.1 1 10 Volume Mixing ratio CO2 CH4 H2O H2 H CO OH 1e-20 1e-15 1e-10 1e-05 1e-20 1e-15 1e-10 1e-5 Volume Mixing ratio 1 1 1e-20 1e-15 1e-10 1e-05 1e-20 1e-15 1e-10 1e-5 Volume Mixing ratio 1 1 1e-20 1e-15 1e-10 1e-05 1e-20 1e-15 1e-10 1e-5 Volume Mixing ratio 1 1 Volume Mixing Ratio Volume Mixing Ratio Volume Mixing Ratio 1e-20 1e-15 1e-10 1e-5 Volume Mixing ratio Volume Mixing Ratio 1 1 11 F G K M Fig. 8.- Grid of mixing ratios vs pressure for H-dominated planet models orbiting inactive F (Tef f,⋆ =7000 K), G (Tef f,⋆ =6000 K), K (Tef f,⋆ =5000 K) and M (Tef f,⋆ =3800 K) stars (1, 2, 3 and 4th row, respectively). The 5 columns present the distance of the planet from their host star for 0.01, 0.015, 0.025, 0.05 and 0.1AU, from left to right respectively. The solid lines show the results when disequilibrium chemistry due to vertical mixing and photochemistry is included, while the dotted lines show the mixing ratios in equilibrium. thought. However, these estimates are an approximation from GCM models calculated for one specific planet and can not be taken as a general result for other planets. Given the uncertainty of KZZ, we model two ex- treme and one intermediate value here in order to ex- plore the degeneracies in the abundance of observable species on this parameter. Following Parmentier et al. (2013) we choose two extreme values that are the mini- mum and maximum values derived in their parametriza- tion and one intermediate value: KZZ = 108 cm2/s, KZZ = 1012 cm2/s and KZZ = 1010 cm2/s, respec- tively. Figure 9 shows the influence of the eddy diffu- sion coefficient on the volume mixing ratios as a function of pressure for four planets in our grid located at 0.01, 0.015, 0.025 and 0.05 AU from a K star. Figure 9 shows semimajor axis at the top and the KZZ values on the right side. Figure 9 shows that the pressure where the quenched level is reached depends on the vertical mixing. Strong mixing in the atmosphere implies that molecules like CO and CH4 will be quenched at higher pressures, deeper in the atmosphere, than for weaker vertical mix- ing. CO2 shows the same behavior for semimajor axis of 0.025 and 0.05 AU.The effects of vertical mixing on the mixing ratios of CO and CH4 was also studied by Visscher & Moses (2011). They analyzed the abundance of CO in the atmosphere of Gliese 229b and CH4 in HD 189733b and found similar results. In the all cases shown in Figure 9, the mixing is efficient enough to maintain the concentrations of the species well mixed very deep in the atmosphere, where the equilib- rium chemistry dominates. Two effects dominate dise- quilibrium chemistry at high altitudes: vertical mixing and dissociation driven by the UV flux received from the star. Photodissociation of atmospheric molecules due to absorption of energetic photons maintains chemical dise- quilibrium within the atmosphere, but fast reaction rates in a well mixed atmosphere enable recombination of the molecules split apart by photochemistry. Therefore pho- tochemistry becomes relevant at low pressures in the at- mosphere, and it is more efficient when vertical mixing is weak. Figure 9 shows that for KZZ = 1012 cm2/s to KZZ = 108 cm2/s (from stronger to weaker vertical mixing) dissociation becomes more efficient. Therefore, some molecules that would dissociate at rel- atively low pressures for a small eddy coefficient, will dissociate at even lower pressures when the mixing is stronger in the atmosphere, becoming one of the major species in the upper atmosphere, like CH4 for planets lo- cated at 0.025 and 0.05 AU. Figure 9 shows that CH4 is photodissociated at ∼ 10−5 bars for KZZ = 108 cm2/s, but has a mixing ratio higher than CO for the same pres- sure for KZZ = 1012 cm2/s, showing that the tempera- ture (due to its semimajor axis) is only one of the relevant 12 1e-6 1e-06 ) s r a b ( e r u s s e r P ) s r a b ( e r u s s e r P ) s r a b ( e r u s s e r P 1e-05 1e-5 1e-4 0.001 0.0001 0.001 ) s r a b ( P 0.01 0.01 0.1 0.1 1 1 10 10 1e-6 1e-06 1e-05 1e-5 1e-4 0.001 0.0001 0.001 ) s r a b ( P 0.01 0.01 0.1 0.1 1 1 10 10 1e-06 1e-6 1e-05 1e-5 1e-4 0.001 0.0001 0.001 ) s r a b ( P 0.01 0.01 0.1 0.1 1 1 10 10 0.01 AU CH4 CO2 O CO H ) s r a b ( P OH H2O 1e-06 1e-05 H2 0.0001 0.001 0.01 0.1 1 10 0.015 AU CO2 CH4 H2O O OH CO H2 ) s r a b ( P H 1e-06 1e-05 0.0001 0.001 0.01 0.1 1 10 0.025 AU 0.05 AU CO2 CH4 CO 1e-06 1e-05 0.0001 0.001 0.01 H2 ) s r a b ( P H2O H O CO CH4 CO2 Kzz 1e8 cm /s 2 H H2 H2O OH 0.1 O OH 1 10 1e-20 1e-15 1e-10 1e-05 1 1e-20 1e-15 1e-10 1e-05 1 1e-20 1e-15 1e-10 1e-05 1 1e-20 1e-15 1e-10 1e-05 1 Volume Mixing ratio CH4 CO2 O CO H ) s r a b ( P OH H2O 1e-06 1e-05 H2 0.0001 0.001 0.01 0.1 1 10 Volume Mixing ratio CO2 CH4 H2O O OH CO H2 ) s r a b ( P H 1e-06 1e-05 0.0001 0.001 0.01 0.1 1 10 Volume Mixing ratio CO2 CH4 CO 1e-06 1e-05 0.0001 0.001 0.01 H2 ) s r a b ( P H2O H O Volume Mixing ratio CO CH4 CO2 H H2 H2O OH 0.1 O OH 1 10 Kzz 1e10 cm /s 2 1e-20 1e-15 1e-10 1e-05 1 1e-20 1e-15 1e-10 1e-05 1 1e-20 1e-15 1e-10 1e-05 1 1e-20 1e-15 1e-10 Volume Mixing ratio CH4 CO2 O CO H ) s r a b ( P OH H2O 1e-06 1e-05 H2 0.0001 0.001 0.01 0.1 1 10 Volume Mixing ratio H2O CO H2 ) s r a b ( P CO2 H O CH4 OH 1e-06 1e-05 0.0001 0.001 0.01 0.1 1 10 Volume Mixing ratio CO 1e-06 1e-05 CH4 H2O 0.0001 O H H2 ) s r a b ( P 0.001 0.01 OH CO2 0.1 O OH 1 10 Volume Mixing ratio 1e-05 CO 1 CH4 CO2 H H2 H2O Kzz 1e12 cm /s 2 1e-20 1e-15 1e-10 1e-05 1e-20 1e-15 Volume Mixing ratio 1e-10 1e-5 1 1 1e-20 1e-15 1e-10 1e-05 1e-20 1e-15 Volume Mixing ratio 1e-10 1e-5 1 1 1e-20 1e-15 1e-10 1e-05 1e-20 1e-15 Volume Mixing ratio 1e-10 1e-5 1 1 1e-20 1e-15 1e-10 1e-05 1e-20 1e-15 Volume Mixing ratio 1e-10 1e-5 1 1 Volume Mixing Ratio Volume Mixing Ratio Volume Mixing Ratio Volume Mixing Ratio Fig. 9.- Comparison of the mixing ratios for different eddy diffusion coefficients for planets orbiting a K (Tef f,⋆ =5000 K) star. The different rows show the mixing ratios as a function of pressure for KZZ = 108 cm2/s (top), KZZ = 1010 cm2/s (middle) and KZZ = 1012 cm2/s (bottom). Solid lines show disequilibrium chemistry results, dotted lines show the mixing ratios in equilibrium. factors when determining major observable constituents in an exoplanet's atmosphere. CO2 and H2O also show this behavior. A higher dis- sociation of H2O molecules implies less H2 and more H in the atmosphere, what is shown in Fig. 9 when go- ing from stronger to weaker eddy diffusion coefficients. H2 is the major gas in the atmosphere for KZZ = 1012 cm2/s. H becomes the dominant gas for KZZ = 108 cm2/s, which could be relevant for atmospheric escape studies in hot exoplanets (e.g., Lammer et al. (2013); Kurokawa & Kaltenegger (2013)). Therefore quantitative measurements from observations of these gases in the atmospheres of hot extrasolar giant plan- ets and mini-Neptunes are needed to constrain the value of the vertical mixing in hot exoplanets atmospheres. 4.2. Clouds and hazes We do not consider the existence of atmospheric con- densates in our models. The presence of clouds and atmospheric hazes can lead to variations in the opac- ities, that can alter the thermal profile in the atmo- sphere. The presence of clouds were inferred for Kepler- 7b (Demory et al. 2013) and observations made with the Spitzer space telescope of the dayside atmospheres of some hot-Jupiters that presented an excess of emission that was interpretated as thermal inversions in the atmo- sphere (e.g.Burrows et al. (2007); Knutson et al. (2008)). We do not include thermal inversions in the temperature and profiles presented, in order to limit the number of free parameters we explore in this paper. An exploration of possible inversions in these exoplanet atmospheres will be address in a future paper. 4.3. Radiative profile Our assumption of a pure radiative model in planets with an extended hydrogen and helium atmosphere is supported by models by several groups (Guillot et al. 1996; Guillot & Showman 2002), which show that highly irradiated hot-Jupiters present an extended radiative layer. Note that models by Dobbs-Dixon & Lin (2008), showed that in some cases, the night side of planets might have a convective layer even at low optical depths. Hotter planets may also exhibit weaker global energy redistri- bution giving them stronger day/night temperature con- trasts (e.g.,Cowan & Agol (2011)), not considered in our model grid. 4.4. Different elemental abundances The atmospheric elemental abundances of hot exo- It planets may differ from that of their host stars. is an unknown and therefore another parameter to ex- plore using atmospheric models. Recent studies have suggested that lower-mass planets may also have atmo- spheres highly enriched in heavy elements (Fortney et al. 2013; Moses et al. 2013), which is a plausible explanation to the apparent CO-rich, CH4-poor nature of GJ 436b (Stevenson et al. 2010). High C/O ratios were suggested in order to explain the abundances detected in WASP- 12b (Madhusudhan & Seager 2011; Moses et al. 2012), while studies by Crossfield et al. (2012); Line et al. (2013) show that there is no evidence for C/O≥ 1 in WASP-12b atmosphere. Moses et al. (2012), also ex- plored a range of different C/O ratios (0.1 <C/O< 2) on the atmospheric composition of three exoplanets: CoRoT-2b, XO-1b and HD 189733b, finding the first two consistent with the assumption of high C/O ratios, and HD 189733b consistent with a C/O≤ 1. These results show that the elemental abundances expected for hot exoplanets is so far not known. In this paper we focussed on the influence of semimajor axis, stellar flux and vertical mixing on the atmospheric models of gas planets. To limit the parameter space to explore, we adopted solar elemental abundances in the atmospheres. The atmospheric composition and possible deviations of the elemental abundances from solar values on the grid will be explored in a future paper. 5. CONCLUSION In this paper we present a grid that links astrophys- ical observable data (orbital distance and effective stel- lar temperature) with the atmospheric composition ex- pected in hot-exoplanet atmospheres. The thermal and photochemical atmospheric grid presented in this paper can be applied to current and future planetary observa- tions to inform what observable atmospheric species are expected for H-dominated exoplanet's atmospheres with 700 < Teq <2800 K. We explored the effect of temperature and UV flux linking them to observables like semimajor axis and stel- lar type in the composition of hot and cool mini-Neptune and giant planet atmospheres models. We used a one di- mensional climate and photochemical code to study these effects for five known extrasolar planets (HD97658b, HD 189733b, XO-1b, CoRoT-2b and WASP-12b) as well as provide a grid of planets from 2800 K < Teq <700 K. We build a grid of planets with different temperatures assuming an albedo of 0.01, calculated based on their semimajor axis (0.01, 0.015, 0.025, 0.05 and 0.1 AU), for a grid of inactive M, K, G and F main sequence stars (Tef f,⋆ =3800, 5000, 6000 and 7000 K, respectively). The different semimajor axis dominate thermal profiles of the planets while the stellar spectral energy distribu- tion affects the photochemistry of the atmospheres due to UV radiation received by the planet. 13 We found that the atmospheres of the hottest planets in the grid are dominated by H, H2, O, H2O and CO, with the last one being the major carbon compound species for all the planets in the grid. For the cooler planets other gases like CH4 become dominant, due to the fact that methane is more stable at lower temperatures. Different stellar spectral types supply different levels of UV flux, going from low flux for the inactive M star (∼ 103 photons/s/cm2/A for 1500A) to high UV flux for the F star (∼ 1010 photons/s/cm2/A for 1500A). This in turn affects the photochemistry in a planet's atmosphere due to the dissociation of molecules occurring down to higher pressure when going from M to F stars. Our results show that CH4 and H2O are the most sen- sitive species to UV flux, and this in turn affects the mix- ing ratios of other major observable species like CO2, H, H2, OH, having a higher photolysis rate, and therefore a lower mixing ratio for pressures P/ 10−5 bars for plan- ets orbiting an F star than an M star. As a result of the photolysis of H2O, and the subsequent destruction of H2 in the reaction with OH, large amounts of H are produced, which replaces H2 as the major gas in the at- mosphere, for planets receiving high UV radiation. This effect is important when the photolysis of H2O is large and therefore is higher when going from M to F host stars. Vertical mixing is important in exoplanet atmospheres, since different possible values change the mixing ratios of observable gases in the atmospheres. Mixing ratio of CH4 is indicative of the effect of vertical mixing and therefore departure of equilibrium in the atmosphere, specially for pressures close to 1 bar. When looking at lower pressures (10−4 >P> 10−7bars), Other species, like H2O, H, H2 or CO2 also indicate the efficiency of vertical mixing in the atmosphere, with larger mixing ratios for stronger mix- ing in the planet's atmosphere. Dissociation becomes less efficient when going from weak to strong vertical mixing in the atmosphere, which affects the observable atmospheric abundances of most of the species. Measure- ments of these gases' abundance for short orbital period exoplanets can be used to explore and constrain vertical mixing in these atmospheres. We would like to thank Ravi Kumar Kopparapu for in depth discussions. We are grateful to Jonathan Fort- ney, Ian Crossfield, Andras Zsom, Brad Hansen and Jim Kasting for stimulating discussions and comments that strengthened our paper, as well as Sarah Rugheimer for providing the stellar fluxes. We thank the referee for fruitful suggestions. LK and YM acknowledge DFG funding ENP Ka 3142/1-1 and the International Space Science Institute ISSI. This research has made use of the Exoplanet Orbit Database and the Exoplanet Data Ex- plorer at exoplanets.org. REFERENCES Allard, F., Hauschildt, P. H., Alexander, D. R., Tamanai, A.,& Brogi, M., Snellen, I. A. G., de Kok, R. J., et al., 2012, Nature, Schweitzer, A., 2001, ApJ, 556, 357 486, 502 Asplund, M., Grevesse, N., & Sauval, A. 2005, in ASP Conf. Ser. Burrows, A., Hubeny, I., Budaj, J., Knutson, H. A., & 336, Cosmic Abundances as Records of Stellar Evolution and Nucleosynthesis, ed. T. G. Barnes, III & F. N. Bash (San Francisco, CA: ASP), 2538 Charbonneau, D., 2007, ApJ, 668, L171 Burrows, A., Budaj, J., & Hubeny, I., 2008, ApJ, 678, 1436 Burrows, A., Ibgui, L., & Hubeny, I., 2008, ApJ, 682, 1277 Batalha, N. M. et al., 2011,ApJ, 729, 27 14 Burrows, A. & Orton, G., 2010, chapter in Exoplanets,edited by S. Seager, Space Science Series of the University of Arizona Press (Tucson, AZ) de Kok, R. J., Brogi, M., Snellen, I. A. G., et al., 2013, arXiv:1304.4014 Charbonneau, D., Brown, T. M., Noyes, R. W., & Gilliland, R. L., 2002, ApJ, 568, 377 Charbonneau, D., Allen, L. E., Megeath, S. T., et al., 2005, ApJ, 626, 523 Cowan, N. B. & Agol, E., 2011,ApJ,729,54 Crossfield, I., et al, 2012 Deming, D., Seager, S., Richardson, L. J., & Harrington, J., 2005, Nature, 434, 740 Lissauer, J. J., et al., 2011, Nature, 470,53 Madhusudhan, N. & Seager, S. 2009, ApJ, 707, 24 Madhusudhan, N., Harrington, J., Stevenson, K. B., et al. 2011a, Nature, 469,64 Madhusudhan, N. & Seager S., 2011, ApJ, 729, 41 Madhusudhan, N., 2012, ApJ, 758, 36 Miller-Ricci, E., & Fortney, J. J. 2010, ApJ, 716, 74 Moses, J. I., Visscher, C., Fortney, J. J., et al. 2011, ApJ, 737, 15 Moses, J. I., Madhusudhan, N., Visscher, C., & Freedman R.,S., 2012, in press Moses, J. I., Line, M. R., Visscher, C., et al. 2013, submitted to ApJ (arXiv:1306.5178v1) Murray-Clay, R. A., Chiang, E. I., & Murray, N. 2009, ApJ, 693, Deming, D., Wilkins, A., McCullough, P., et al., 2013, 23 arXiv:1302.1141 Demory, B.-O., de Wit, J., Lewis, N., et al. 2013, ApJ, 776, L25 Dragomir, D., Matthews, J. M., Eastman, J. D., et al., 2013, ApJ, Parmentier, V., Showman A. & Lian Y., 2013, A&A, in press, arXiv:1301.4522v1 Pavlov, A., Brown, L., & Kasting, J. F. 2001, J. Geophys. Res., 772, L2 106, 23267 Dobbs-Dixon, I., & Lin, D. N. C. 2008, ApJ, 673, 513 Fegley, Jr., B., & Prinn, R. G. 1985, Astrophysical Journal, 299, Pickles, A. J. 1998, PASP, 110, 863 Redfield, S., Endl, M., Cochran, W. D., & Koesterke, L., 2008, 1067 ApJ, 673, L87 Fortney, J. J., Mordasini, C., Nettelmann, N., Kempton, E., & Rowe, J. F., Matthews, J. M., Seager, S., et al., 2008, ApJ, 689, Zahnle, K., 2013, ApJ, submitted Guillot, T., 2010, A&A 520, A27 Guillot, T., & Showman, A. P. 2002, A&A, 385, 156 Guillot, T., Burrows, A., Hubbard,W. B., Lunine, J. I., & Saumon, D. 1996, ApJ, 459,L35 Hansen B., 2008, ApJ, 179, 484 Howard, A. W., 2013, Science, 340,572 Howe, A. R., & Burrows, A. S., 2012, ApJ, 756, 176 Hubeny, I., & Burrows, A., 2007, ApJ, 669, 1248 Kurokawa, H., & Kaltenegger, L., 2013, MNRAS, 433, 3239 Knutson, H. A., Charbonneau, D., Allen, L. E., Burrows, A., & Megeath, S. T., 2008, ApJ, 673, 526 Konopacky, Q. M., Barman, T. S., Macintosh, B. A., & Marois, C., 2013, Science, 339, 1398 Kopparapu, R. k., Kasting, J. F., & Zahnle, K. J., 2012, ApJ, 745, 77 Kurucz, R.L. (1979) Model atmospheres for G, F, A, B, and O stars. Astrophys. J. Suppl. Ser. 1340. Lammer, H., Odert, P., Leitzinger, M., et al., 2009, A&A, 506, 399 Lammer, H., Erkaev, N. V., Odert, P., Kislyakova, K. G., Leitzinger, M., & Khodachenko, M. L., 2013, MNRAS 430, 1247 Liang, M. C., Parkinson, C. D., Lee, A. Y. T., Yung, Y. L., & Seager, S., 2003, ApJ, 596, L247 Line, M. R., Liang, M. C., Yung, Y. L., et al. 2010, ApJ, 717, 496 Line, M. R., Vasisht, G., Chen, P., Angerhausen, D., & Yung, Y. L. 2011, ApJ, 738, 32 Line, M. R., Wolf, A. S., Zhang, X., et al., 2013, (arXiv:1304.5561) 1345 Rugheimer, S., Kaltenegger, L., Zsom, A., Segura, A., & Sasselov, D., 2013, Astrobiology, 13, 251 Seager, S., Bains, W.,& Hu, R., 2013, ApJ, 777, 95 Seager, S., & Deming, D., 2010, ARA&A, 48, 631 Schaefer, L., & Fegley, B., Jr. 2009, ApJ, 703, 113 Segura, A., Krelove, K., Kasting, J. F., Sommerlatt, D., Meadows, V., Crisp, D., Cohen, M., & Mlawer, E. 2003, Astrobiology, 3, 689 Showman, A.P., Fortney, J.J., Lian, Y., et al., 2009, ApJ, 699, 564 Snellen, I. A. G., de Kok, R. J., de Mooij, E. J. W., & Albrecht, S., 2010, Nature, 465, 1049 Spiegel, D. S., Burrows, A., Ibgui, L., Hubeny, I., & Milsom, J. A., 2010, ApJ, 709, 149 Stevenson, K. B., Harrington, J., Nymeyer, S., et al., 2010, Nature, 464, 1161 Sudarsky, D., Burrows, A., & Pinto, P. 2000, ApJ, 538, 885 Valencia, D., Sasselov, D. D., & O'Connell, R. J., 2007, ApJ, 665, 141 Venot, O., Hebrard, G., Agundez, M., et al. 2012, A&A, in press Visscher, C., & Moses, J. I., 2011, ApJ, 738, 72 White, W. B., Johnson, S. M. & Dantzig G. B., 1958, J. Chem. Phys. 28, 751 Wright, J. T., Fakhouri, O., Marcy, G. W., et al., 2011, PASP, 123, 412 Zahnle, K., Marley, M. S., Freedman, R. S., Lodders, K., & Fortney, J. J., 2009, ApJ, 701, L20 Zahnle, K., Marley, M. S., & Fortney, J. J. 2009a, arXiv:0911.0728
1104.4322
1
1104
2011-03-17T17:39:39
On the Origin and Evolution of Life in the Galaxy
[ "astro-ph.EP", "astro-ph.CO" ]
A simple stochastic model for evolution, based upon the need to pass a sequence of n critical steps (Carter 1983, Watson 2008) is applied to both terrestrial and extraterrestrial origins of life. In the former case, the time at which humans have emerged during the habitable period of the Earth suggests a value of n = 4. Progressively adding earlier evolutionary transitions (Maynard Smith and Szathmary, 1995) gives an optimum fit when n = 5, implying either that their initial transitions are not critical or that habitability began around 6 Ga ago. The origin of life on Mars or elsewhere within the Solar System is excluded by the latter case and the simple anthropic argument is that extraterrestrial life is scarce in the Universe because it does not have time to evolve. Alternatively, the timescale can be extended if the migration of basic progenotic material to Earth is possible. If extra transitions are included in the model to allow for Earth migration, then the start of habitability needs to be even earlier than 6 Ga ago. Our present understanding of Galactic habitability and dynamics does not exclude this possibility. We conclude that Galactic punctuated equilibrium (Cirkovic et al. 2009), proposed as a way round the anthropic problem, is not the only way of making life more common in the Galaxy.
astro-ph.EP
astro-ph
On the Origin and Evolution of Life in the Galaxy Michael McCabe and Ho lly Lucas Department of Mathemat ics, University o f Portsmouth, Lion Terrace, Portsmouth, Hants PO41 3HF UK Abstract A simple stochastic model for evolut ion, based upon the need to pass a sequence of n crit ical steps (Carter 1983, Watson 2008) is applied to both terrestrial and extraterrestrial origins o f life. In the former case, the time at which humans have emerged during the habitable period of the Earth suggests a value o f n = 4. Progressively adding earlier evolut ionary transit ions (Maynard Smith and Szathmary, 1995) gives an optimum fit when n = 5, implying either that their init ial transit ions are not critical or that habitability began around 6 Ga ago. The origin o f life on Mars or elsewhere within the Solar System is excluded by the latter case and the simple anthropic argument is that extraterrestrial life is scarce in the Universe because it does not have time to evolve. Alternat ively, the timescale can be extended if the migrat ion of basic progenotic material to Earth is possible. If extra transit ions are included in the model to allow for Earth migration, then the start of habitability needs to be even earlier than 6 Ga ago. Our present understanding of Galact ic habitability and dynamics does not exclude this possibility. We conclude that Galact ic punctuated equilibrium (Cirkovic et al. 2009), proposed as a way round the anthropic problem, is not the only way o f making life more common in the Galaxy. Keywords Evolut ion, origin o f life, Galaxy, stochastic, mathemat ical model, crit ical evo lut ionary transit ions, Galact ic Habitable Zone, habitability, Earth migrat ion Introduction It is usually accepted that our atomic origins are extraterrestrial and that physics is universal. The chemical elements required for life were not created on Earth, but through nucleosynthesis during the Big Bang (H,He,Li), inside the cores of stars (He, C, N, O …) and within supernovae (Ca, P, S, Na, K, Cl …). The origin o f terrestrial mo lecules required for life, notably water, is less certain, but comets and meteorites impacting the young Earth must have made a significant contribution and extraterrestrial origins are certainly not excluded. Water ice and organic material has recent ly been detected on a main belt asteroid (Rivkin and 1 Emery, 2010) increasing their possible role in delivering mo lecules required for life. Chemistry is regarded as universal and our molecular origins are accepted as being both terrestrial and extraterrestrial. Bio logy is not generally regarded as universal and the question o f whether our biological origins are terrestrial or extraterrestrial is highly controversial (Table 1). Standard evolut ionary bio logy adopts a time on Earth between 3.5 and 4 Ga ago as its starting point with deep ocean hydrothermal vents as one possible location (Martin et al., 2008) and recognises that the rapid emergence of life is a puzzle. Maynard Smith and Szathmary (1995) have suggested eight steps on Earth leading from replicat ing mo lecules to mankind, of which seven are regarded as crit ical. Table 1 includes the crit ical transit ions (1) to (7), but omits the faster transit ion expected from solitary individuals to colonies between (6) and (7). Modern Examples H to He Li to Fe above Fe H2O glycine lipids bilayer vesicles retrovirus (RNA) bacteria algae protozoa primates authors MM/HL Transition ET Origin nucleosynthesis Big Bang nucleosynthesis stellar cores nucleosynthesis supernovae chemical react ion ISM, comets ISM, meteorites chemical react ion comets/meteorites chemical react ion self organisat ion (1) least unlikely less unlikely synthesis (2) RNA to DNA (3) unlikely endosymbio sis (4) more unlikely most unlikely asexual to sexual (5) cell different iat ion (6) science fict ion science fict ion language (7) Entity small atoms medium atoms large atoms smaller mo lecules larger mo lecules replicat ing mo lecules protocells chromosomes prokaryotes eukaryotes protists animals/plants/fungi humans Table 1 Extraterrestrial Origins and Transit ions Required for Human Life to Emerge A simple stochastic model, first proposed by Carter (1983) and discussed by Barrow and T ipler (1986), Hanson (1998), Flambaum (2003), assumes that evo lut ion is governed by a series or crit ical transit ions or bottlenecks, which must be passed to advance from basic mo lecules to simple, complex and then intelligent life. The anthropic model allowed the derivat ion of a probability density funct ion and expectation time for the final step, based upon the habitable lifet ime o f the planet. Watson (2008) has further analysed the model to derive probability density functions and expectation times for every step. He ident ifies 2 seven crit ical transit ions in bio logical evo lut ion from amongst the eight proposed by Maynard Smith and Szathmary (1995) and compares their probabilit ies with their est imated times o f occurrence. He also discusses geochemical transit ions relevant to bio logy, although they do not form part of the strict ly bio logical evo lut ionary sequence based upon informat ion. 3 Critical Step Model for Evolution The crit ical step model assumes that there is an ordered sequence of transit ions which govern the pace of evo lut ion. Any intermediate steps or stages are assumed to occur rapidly without increasing the time between steps. The crit ical steps are assumed to occur stochast ically with uniform, small, but unequal probabilit ies, 1 … n where n << 1 / th where th is the habitable lifet ime of the planet. One alternative is a “long, slow fuse” model, which allows for a much larger sequence of more likely events. The probability density for the m-th crit ical step in an n crit ical step model is: The expectation time for the m-th critical step in an n critical step model is: (2) (1) i.e. the spacing of the steps will therefore tend to be even throughout the habitability period, although they will be subject to statist ical variation. Watson adopted 4 Ga ago as the start of habitability, corresponding roughly to the period of late, heavy bombardment. The end of habitability is taken as 1 Ga in the future, corresponding to the time when so lar luminosity and atmospheric models predict an extreme temperature rise result ing from faster weathering of rocks to carbonates and subsequent CO2 loss. The probability density funct ions for the n-step models with n = 1,2 .. 5 are the polynomials shown in Table 2. The fit between the probability density distribut ions in the 7 -step model and the estimated times at which the crit ical steps actua lly occurred is poor. The main cause of the discrepancy is that prokaryotic life emerged relat ively soon after habitability. The first three steps are much more closely spaced than the subsequent steps, in disagreement with the expectation of closer spacing predicted by the model. Watson shows that the fit can be significant ly improved by assuming that the emergence of prokaryotic life only required a single init ial step in a 5-step model. 4 hmnhmnmttttmmnntP)()!1()!(!)(1/ Pm/n(t) n = 1 n = 2 n = 3 n = 4 n = 5 m = 1 m = 2 m = 3 m = 4 m = 5 m = 3 m = 4 m = 5 4 20/6 m = 2 25/6 15/4 3 15/6 10/3 10/4 2 10/6 m = 1 5/2 5/3 5/4 1 5/6 Table 2 Probability Density Funct ions for 1 to 5 Step Models (habitability lifet ime th = 5 Ga) <tm/n> n = 1 n = 2 n = 3 n = 4 n = 5 Table 3 Expectation Time for Critical Transit ions, <t m/n> in Ga (habitable lifet ime th = 5 Ga) Considering the final transit ion alone, the emergence of humans after 4 Ga of a 5 Ga habitable lifet ime on Earth implies that the nth out of n transit ions occurred after 4 Ga. Table 3 shows the expectation times derived from equat ion (2) and indicates that this is most likely for n = 4, since m = n = 4 gives <t4/4> = 4 Ga. An alternative to a single step leading to prokaryotes is the suggestion that the timescale for evolut ion might be extended back by starting evo lut ion earlier (Carter 2008, Watson 2008). Evidence of water on Earth 4.3 Ga ago (Valley at al. 2002) could signify the start of habitability. Meteorites from Mars suggest that life might have originated there and migrated to Earth (Treeman et al., 2000). If the origin o f life occurred elsewhere in the Solar System, it must have been within the past 4.6 Ga. This extens ion of 0.6 Ga to the time over which life evo lved helps to resolve the discrepancy between the model and the est imated transit ion times, but it does not alter the conclusion that the optimum number of crit ical transit ions is less than seven. Carter (2008) suggests 5 an optimum of six transit ions if early evo lut ion began on Mars, but stops short of proposing any earlier steps towards life. Origins within our Galaxy Our Galaxy appears to be almo st as old as the Universe itself, since the oldest known halo star is est imated at 13.2 Ga old (Frebel, 2007). This ancient star is metal-poor and lacking in the heavy elements needed for planetary format ion, while the oldest brown dwarfs, intermediates between stars and planets, are estimated to be 10 Ga old (Schilbach and Röser, 2009). The Galactic thin disc, where conditions are more favourable for life, is estimated to be between 6.5 and 11.1 Ga old (Del Peloso 2005). The concept of a Galactic Habitable Zone, a region within the Galaxy favourable for the development of life, was first introduced by Gonzalez (2001) and expanded upon in Gonzalez (2005). Lineweaver et al. (2004) identified the present GHZ as an annular region between 7 and 9 kpc, which has moved outwards from the Galactic centre and is composed of stars that formed between 8 Ga and 4 Ga ago. His model of Galactic evolution and habitability considered the need for adequate star formation, heavy elements for planets to form and time for life to evolve, together with an absence of supernovae. Prantzos (2008) concludes that that habitability is far less clear -cut, because the factors involved are poorly understood. In particular, the effect of heavy metals on planet formation, hot Jupiters on terrestrial planet survival and supernovae radiation on life are extremely uncertain. His own model predicts a narrow annulus for Galactic habitability starting at 3 kpc around 13 Ga ago which then expands as it moves outwards, until it eventually encompasses the whole Galaxy within the lifetime of the Solar Syst em and has a peak probability of Earth-like planets with life approaching 0.1 at 10kpc from the Galactic centre today. More recent work (Mattson, 2009) has modeled the build-up of carbon in the Galaxy predicting a peak at ~5.5 kpc around 8 Ga ago and at the solar distance ~9 kpc around 5 Ga ago, close to the time when the Earth formed. Some of the necessary ingredients and conditions for life may therefore have existed within the Galaxy, albeit at different locations, for many billions of years before the formation of the Earth. Based upon current understanding, it is reasonable to consider the possibility of life originating elsewhere in the Galaxy before it began on Earth. Proponents of the Rare Earth Hypothesis (Brownlee and Ward, 2000) may place many further conditions upon the emergence of complex life, but we are concerned with the earliest, simple stages. Table 4 summarises some of the timings for key events and transitions through the history of the Universe and the Earth. Although many may be rega rded as controversial, they are used as best estimates in subsequent calculations. The origin of life in the Galaxy, or more specifically the start of Galactic habitability, is taken to have occurred 8.5 ± 5 Ga ago, where the central value of 8.5 Ga reflects the work of Lineweaver et al 6 (2004), Prantzos (2008) and Mattson (2009) and the range of uncertainty allows for the most extreme conceivable values from the first stars in the Universe to the origin of life on Earth. 7 Time in past (Ga) 13.7 13.7 - 13.7 - 13.6 13.2 1 [13.5] 8.5 [3.5] 2 [6.5] 8.3 [11.1] 3 4.57 4.55 [4.4] 4.3 4 [4.1] 4.0 [3.8] 5 [3.85] 3.5 - 3.0 [2.6] [3.2] 2.8 – 2.6 2.2 – 2.0 [2.7] 1.9 – 1.2 [0.8] [1.9] 1.2 – 0.6 Time since habitability on Earth (Ga) 0.0 0.2 0.4 0.5 2.1 2.5 2.8 Critical Event / Transition Type C Big Bang C nucleosynthesis o f H/He after ~100s C atoms formed after 0.3 My (CMBR) C first stars expected A C first (halo) stars in Galaxy observed A existence of Galact ic Habitable Zone A Galactic thin disc formed A E format ion of Sun E format ion of Earth liquid water on Earth E late heavy bombardment (Earth habitability) L proto-cells (mo lecules in compartments) B (1) B (2) chromosomes (RNA) B (3) emergence o f prokaryotes (life) E oxygen photosynthesis atmosphere becomes oxidising E B (4) prokaryotes to eukaryotes B (5) asexual to sexual populat ions protists to animals/plants/fungi B (6) (cell differentiat ion) microscopic to macroscopic life (Cambrian) E B (7) primates to man predicted CO2 loss (inhabitability) B 4.0 5.0 0.55 0.0001 – 0.0002 (1.0) in future C = Cosmo logy A = Astronomy / Galaxy L = Lunar Science B = Bio logical Informat ion [extreme est imates] best estimates 1 Frebel (2007) 2 See preceding text 3 Del Pelosa (2005) 4 Valley et al (2002) 5 Cohen et al. (2000) (1) to (7) Crit ical Evo lut ionary Transit ions (Maynard-Smith and Szathmary 1995, Watson 2008) Table 4 Astrobio logical T imings E = Earth Science / Geology 8 Extending the Timescale in the Critical Transition Model Calculations of the crit ical transit ion model have been carried out using the Maple computer algebra system. Maple was used to plot and integrate the probability density distributions defined by equation (1) and hence determine the probability that a transit ion would occur before its estimated time. Since the distribut ions are polynomial funct ions, no numerical techniques have been required to perform the integrat ions. Figure 1 shows the probability density distr ibut ions for the 7-step and 5-step models with the best estimates for the actual transit ion t imes marked on each curve by a cross. For the intermediate curves a cross towards the peak of the distribut ion suggests a good fit. The integrated area below the probability density curve to the left of the cross provides a more precise measure for establishing the goodness of fit. The adoption of the values in Table 4, which differ within reasonable limits from those adopted by Watson (2008), yields different probabilit ies that a step occurred at or before the observed time, but confirms his conclusions. A 7 -step model fits poorly (probabilit ies 25%, 10%, 3%, 33%, 23%, 11%, 21% respectively ) while a 5-step model fits much better (probabilit ies 41%, 70%, 50%, 27%, 33% respectively), where values clo ser to 50% imply a more acceptable model. The essent ial dist inguishing factor between the models is the rapid emergence of life, which is allowed for in the 5-step model by combining the first three transit ions of the 7-step model into one. 9 Figure 1 Crit ical Transit ion Models for (a) 7-steps (b) 5-steps Compared with Estimated Transit ion T imes (start of habitability = 4 Ga ago) 10 An alternative explanat ion is that the earlier transit ions have occurred elsewhere in the So lar System or beyond. By considering the goodness of fit (3) where Pi = probability o f the ith step occurring before its actual est imated time (%) n = total number of steps we have examined how well the model performs when different numbers o f steps and earlier transit ions on extraterrestrial locat ions are allowed. If all the probabilit ies were 50% , giving F = 0, the fit would be regarded as ideal. Figure 2 shows that F is minimized for n = 5 and that the 5-step model does indeed fit the data best for purely Earth-based evo lut ion. Figure 2 Optimisat ion of the Number of Transit ions in the Crit ical Step Model 11 nPFnii12)50( If these earlier transit ions have taken place “off-site” they can be re-incorporated into the model by extending the timescale. Given that this needs to be longer than the age of the Solar System to make any significant difference to the model, a time of 8 Ga ago, corresponding to the estimated age of the thin disc and start of Galactic habitability according to the Lineweaver (2004) model, is adopted. The probability density distributions and estimated transit ion times are shown in Figure 3, where the probabilit ies for the 5 terrestrial transit ions become 77%, 84%, 70%, 46%, 44%. The first two extraterrestrial transit ions need to be excluded from the goodness of fit, since there is no sensible estimate for their times. Figure 3 Crit ical Transit ion Models for 7-steps Compared with Est imated Transit ion T imes (start of habitability = 8 Ga ago) During the habitability period the first two transit ions would have taken place on an extraterrestrial locat ion and the final five on Earth, where we assume tha t the migration to Earth is not a critical step. However we can consider the consequences if the physical transfer to Earth is regarded as another crit ical step, result ing from an unlikely event over a long timescale? For example, a randomly occurring collision of a suitable comet or meteorite with Earth might be 12 necessary. Alternat ively, one collision or event might be required to init iate a transfer from one locat ion and another to complete it by co llision with the Earth. Lithopanspermia, an idea first suggested by Lord Kelvin in 1871, has been well discussed in subsequent years. Some calculat ions have suggested that interstellar panspermia is unlikely (Melo sh 2003), but viable mechanisms have subsequent ly been suggested (Napier 2004, Wallis and Wickramasinghe 2004). The effect of adding one or two critical transit ions to the model is shown in Figure 4. 13 Figure 4 Crit ical Transit ion Models for (a) 8-steps (b) 9-steps Compared with Estimated Transit ion T imes (start of habitability = 8 Ga ago) 14 The best fit is obtained with the 9-step model, i.e. by having two Earth migrat ion transit ions (Figure 5a). By reducing the start of habitability to 6 Ga ago, a similar goodness of fit can be achieved without any Earth migration transit ions (Figure 5b). Any further reduction in the start of habitability towards the age of the Solar System (Mars) decreases the goodness of fit (Figure 5c). Figure 5 Model Dependence upon the Number of Earth Migrat ion Transit ions for Start of Habitability at (a) 8 Ga (b) 6 Ga (c) 5 Ga 15 Discussion The equal spacing property of the crit ical transit ion model can be used to explain what is happening. If the transit ions 1 - 7 are equally spaced upon the x-axis together with 0 = start of Earth habitability and 8 = end of Earth habitability and the estimated times from Table 4 are plotted, then a best fit line through the origin can be used for comparison (Figure 6). The fit is poor because of the rapid early transit ions. Figure 6 The Enforcement of Crit ical Spacing in a 7-Step Evolutionary Model If the first two transit ions are removed and the best fit line is no longer required to pass through the origin, the y-intercept of -1.6 Ga gives an est imate for the origin of life, i.e. around 5.6 Ga ago (Figure 7). This co rresponds to the value of 6 Ga ident ified for the start of habitability when there were no Earth migration transit ions. If addit ional transit ions are introduced for Earth migration, then the y- intercept, will decrease and life emerge further back in the past. 16 Figure 7 The Enforcement of Crit ical Spacing in a 5-Step Evolutionary Model Lépine et al (2003) have shown that resonance with the spiral gravitat ion field, can cause stars to move radially through the Galaxy over distances o f 2-3 kpc in significant ly less than 1 Ga. Roškar et al. (2008) have modeled this stellar migration and considered its astrophysical implications. A serious possibility is that it could account for the migrat ion of simple life from closer to the Galact ic centre. Haywood (2009) have analysed solar neighbourhood stars showing that around 50% have come from elsewhere, primarily by migrat ion from the inner disc, where Galactic habitability was greater at an earlier time. Not only does the Galactic Habitable Zone move radially outwards from the Galactic centre according to most models (Lineweaver, 2004, Prantzos 2008), but so do a significant number of stars. Thus from both a dynamic and chemical perspective, it may be possible for simple life to migrate from the inner disc o f the Galaxy on realist ic t imescales and hence be included within the crit ical transit ion model. Wesson (2010) argues that, despite viable mechanisms for distribut ing organic material throughout the Galaxy, life could not survive the UV and cosmic ray bombardment. He does though suggest that necropanspermia, the transfer of informat ion via damaged bio logical mo lecules, is feasible. Conclusions A simple stochastic model for evolut ion, based upon the need to pass a sequence of n crit ical steps (Carter 1983, Watson 2008) has been applied to both terrestrial and extraterrestrial origins o f life. In the former case, the time at which humans have emerged during the habitable period of the Earth suggests a value of n = 4. 17 Progressively adding earlier evolut ionary transit ions (Maynard Smith and Szathmary, 1995) gives an optimum fit when n = 5, implying either that their init ial transit ions are not critical or that habitability began around 6 Ga ago. The origin o f life on Mars or elsewhere within the Solar System is excluded by the latter case. Carter (2008) in reviewing his original work, argues for a 5 -step model for terrestrial evo lut ion and a 6-step model if Mars evo lution is included. If extra transit ions are included to allow for Earth migrat ion by simple life, the start of habitability needs to be at an even earlier time. Our present understanding of Galactic habitability and dynamics does not exclude this possibility. Each addit ional migrat ion transit ion pushes back the start of habitability by around 1 Ga. The Drake equation has long been used as a touchstone for astrobio logy. Its mix of astrophysical, bio logical and social science, with an increasing level of uncertainty, continues to provoke discussion (Burchell, 2006). The critical transit ion model addresses an issue for any Drake-type equat ion, e.g. the Rare Earth equation (Brownlee and Ward, 2000). If any one of the factors is zero, then the result is zero. The crit ical transit ion model breaks down the crucial bio logica l steps, typically only represented by one or two factors, and introduces time dependence. In particular, it provides a reasonable fit to the timescales for the emergence o f life on Earth. Furthermore, if the possibility of life originating beyond the Earth is considered, it can provide useful clues as to the time that life might have emerged. Whether the assumpt ions o f the model and the ident ification o f the transit ions are reasonable is open to debate. For example, the likelihood of crit ical transit ions could be governed by causal links between astrophysical and bio logical events. Transit ions might occur relat ively quickly with a high probability if life emerges through stages of Galact ic punctuated equilibrium separated by irregular catastrophic (catalyt ic?) events in the Galaxy such as gamma ray bursts or supernovae as suggested by Circovic (2009). The Earth is certainly not a closed system, being exposed to electromagnet ic radiat ion, neutrinos and cosmic rays from beyond the Solar System and material from within it . The extent to which the Earth is an open system and the degree of correlat ion between astrophysics and bio logy are key issues, which “grand” astrobiological models can help to resolve. Crit ical transit ion models, incorporating both bio logical and astrophysical concepts, provide a standard against which other models can be judged. 18 References Barrow, J.D. and Tipler, F.J., (1986). The Anthropic Cosmological Principle. Oxford University Press. Burchell, M.J., (2006). W(h)ither the Drake Equation?. International Journal of Astrobio logy, 5, 243-250 Carter, B., (1983). The Anthropic Principle and its Implications for Biological Evolution. Phil. Trans. R. Soc. Series A, v. 310, p. 347-363 Carter, B., (2008). Five- or Six-Step Scenario for Evolution?. International Journal of Astrobio logy, 7:177-182, Cambridge University Press Cirkovic, M.M., Vukotic, B., Dragicevic, I., (2009). Galactic Punctuated Equilibrium: How to Undermine Carter’s Anthropic Argument in Astrobiology. Astrobio logy, vo l 9, no. 5, 491 - 581 Cohen, B.A. et al., (2000). Support for the Lunar Cataclysm Hypothesis from Lunar Meteorite Impact Melt Ages. Science 290, (5497):1754-1755 Del Peloso, E. F., (2005). The Age of the Galactic Thin Disk from Th/Eu Nucleocosmochronology. A&A, 440: 1153 Flambaum, V.V., (2003). Comment on Does the Rapid Appearance of Life on Earth Suggest that Life is Common in the Universe? Astrobio logy 3, 237-239 Frebel, A., (2007). Discovery of HE 1523-0901, a Strongly r-Process-enhanced Metal-poor Star with Detected Uranium, Ap. J. 660: L117 Hanson, R., (1998). Must Early Life Be Easy? The Rhythm of Major Evolutionary Transitions. http://hanson.gmu.edu/hardstep.pdf Haywood, M., (2009). Radial Mixing and the Transition between the Thick and Thin Galactic Discs. MNRAS, Volume 388 Issue 3, 1175 - 1184 Gonzalez, G. et al., (2001). The Galactic Habitable Zone: Galactic Chemical Evolution. Icarus, 152, 185 - 200 Gonzalez, G. ,(2005). Habitable Zones in the Universe. Origins Life Evo l. Biospheres 33 555 19 Lépine , J. R. D., Acharova , I. A. and Mishurov, Yu. N. (2003). Corotation, Stellar Wandering, and Fine Structure of the Galactic Abundance Pattern. Ap. J., 589:210-216 Lineweaver, C. H., Fenner, Y. and Gibson, B. K. (2004). The Galactic Habitable Zone and the Age Distribution of Complex Life in the Milky Way. Science, 303, 59-62. Martin, W, et al., (2008). Hydrothermal Vents and the Origin of Life. Nature Reviews Microbiology, 6, 805 - 814 Mattson, L., (2009). On the Existence of a Galactic Habitable Zone and the Origin of Carbon. Swedish Astrobio logy Meet ing, Lund http://videos.nordita.org/conference/SwAN2009/Mattsson.pdf Maynard Smith, J. and Szathmary, E., (1995). The Major Transitions in Evolution. Oxford University Press, ISBN 019850294X Melosh, H.J., (2003). Exchange of Meteorites (and Life?) Between Stellar Systems. Astrobio logy, 3(1): 207-215 Napier, W.M. (2004). A Mechanism for Interstellar Panspermia. Monthly Notices of the Royal Astronomical Society, 348, 1, 46-51 Prantzos, N, (2008). On the “Galactic Habitable Zone”. Space Sci Rev 135:313 Rivkin, A. S. & Emery, J. P. Detection of ice and organics on an asteroidal surface Nature 464, 1322-1323 (2010). Roškar, R. , Debattista, V.P. , Quinn, T.R. , Stinson, G.S. and Wadsley, J., (2008). Riding the Spiral Waves: Implications of Stellar Migration for the Properties of Galactic Discs. Ap. J, 684: L79–L82 Schilbach E., Roeser S., Scholz R.D., (2009). Trigonometric Parallaxes of Ten Ultracool Subdwarfs. A&A, 493, 27 Treeman, A.H. et al., (2000). The SNC Meteorites are from Mars. Planetary Space Science. 48 (12-14): 1213-1230 Valley, J.W. et al. (2002). A Cool Early Earth. Geology 30: 351-354 20 Wallis, M.K. and Wickramasinghe, C.R. (2004). Interstellar Transfer of Planetary Microbiota. Monthly Notices of the Royal Astronomical Society, 348, 1, 52-61 Ward, P. D., and Brownlee, D., (2000). Rare Earth: Why Complex Life is Uncommon in the Universe. Copernicus Books, New York (Springer Verlag). ISBN 0-387-98701-0. Watson, A. J. ,(2008). Implications of an Anthropic Model for the Evolution of Complex Life and Intelligence. Astrobio logy J. v. 8, p 1-11. Wesson, P.S., (2010). Panspermia, Past and Present: Astrophysical and Biophysical Conditions for the Dissemination of Life in Space. Space Sci Rev, Springer ISSN 1572-9672 (online) 21
1611.00741
1
1611
2016-11-02T19:36:51
The Role of Ice Compositions for Snowlines and the C/N/O Ratios in Active Disks
[ "astro-ph.EP", "astro-ph.SR" ]
The elemental compositions of planets define their chemistry, and could potentially be used as beacons for their formation location if the elemental gas and grain ratios of planet birth environments, i.e. protoplanetary disks, are well understood. In disks, the ratios of volatile elements, such as C/O and N/O, are regulated by the abundance of the main C, N, O carriers, their ice binding environment, and the presence of snowlines of major volatiles at different distances from the central star. We explore the effects of disk dynamical processes, molecular compositions and abundances, and ice compositions on the snowline locations of the main C, O and N carriers, and the C/N/O ratios in gas and dust throughout the disk. The gas-phase N/O ratio enhancement in the outer disk (exterior to the H2O snowline) exceeds the C/O ratio enhancement for all reasonable volatile compositions. Ice compositions and disk dynamics individually change the snowline location of N2, the main nitrogen carrier, by a factor of 2-3, and when considered together the range of possible N2 snowline locations is ~11- ~79 AU in a standard disk model. Observations that anchor snowline locations at different stages of planet formation are therefore key to develop C/N/O ratios as a probe of planet formation zones.
astro-ph.EP
astro-ph
Draft version September 11, 2018 Preprint typeset using LATEX style emulateapj v. 5/2/11 6 1 0 2 v o N 2 . ] P E h p - o r t s a [ 1 v 1 4 7 0 0 . 1 1 6 1 : v i X r a THE ROLE OF ICE COMPOSITIONS FOR SNOWLINES AND THE C/N/O RATIOS IN ACTIVE DISKS Ana-Maria A. Piso1,2, Jamila Pegues1,3, Karin I. Oberg1 Draft version September 11, 2018 ABSTRACT The elemental compositions of planets define their chemistry, and could potentially be used as beacons for their formation location if the elemental gas and grain ratios of planet birth environments, i.e. protoplanetary disks, are well understood. In disks, the ratios of volatile elements, such as C/O and N/O, are regulated by the abundance of the main C, N, O carriers, their ice binding environment, and the presence of snowlines of major volatiles at different distances from the central star. We explore the effects of disk dynamical processes, molecular compositions and abundances, and ice compositions on the snowline locations of the main C, O and N carriers, and the C/N/O ratios in gas and dust throughout the disk. The gas-phase N/O ratio enhancement in the outer disk (exterior to the H2O snowline) exceeds the C/O ratio enhancement for all reasonable volatile compositions. Ice compositions and disk dynamics individually change the snowline location of N2, the main nitrogen carrier, by a factor of 2-3, and when considered together the range of possible N2 snowline locations is ∼11-∼79 AU in a standard disk model. Observations that anchor snowline locations at different stages of planet formation are therefore key to develop C/N/O ratios as a probe of planet formation zones. 1. INTRODUCTION The chemical composition of protoplanetary disks is largely dictated by the freeze-out of volatile species. The snowline locations of volatile molecules are therefore cru- cial in determining disk chemical abundances in gas and dust, as well as planet compositions. Carbon and oxygen bearing molecules, such as H2O, CO2 and CO, as well as the carbon-to-oxygen (C/O) ratio in protoplanetary disks and in giant planet atmo- spheres have been extensively studied from a theoretical standpoint ( Oberg et al. 2011b, Ali-Dib et al. 2014, Mad- husudhan et al. 2014, Molli`ere et al. 2015), and snow- lines of volatiles such as H2O and CO have been detected (Zhang et al. 2013, Qi et al. 2013). However, disk chem- istry involves many other molecular compounds (Hen- ning & Semenov 2013) including nitrogen bearing species and hydrocarbons (e.g., Mandell et al. 2012), which may affect the compositions of nascent planets. Both in Solar system comets and in protoplanetary disks, volatile carbon and oxygen are primarily contained in H2O, CO2 and CO (e.g., Lodders 2003, Mumma & Charnley 2011, Oberg et al. 2011b, Boogert et al. 2015). However, some fraction of carbon may also be carried by CH4 (e.g., Oberg et al. 2008), which may change the C/O ratio in gas and in dust at some disk locations. In the case of nitrogen, chemical models of the protostel- lar nebula (e.g., Owen et al. 2001) and of protoplane- tary disks (e.g., Rodgers & Charnley 2002) suggest that N2 was the dominant form of nitrogen, and that giant planets have accreted their nitrogen content primarily as N2 (Mousis et al. 2014). Due to the high volatility of N2, the gas phase nitrogen-to-oxygen (N/O) ratio in 1 Harvard-Smithsonian Center for Astrophysics, 60 Garden Street, Cambridge, MA 02138 2 UCLA, 595 Charles E. Young Drive East, Los Angeles, CA 90095 3 Department of Astrophysical Sciences, Princeton University, Princeton, NJ 08544 the outer disk is expected to be high, perhaps more en- hanced than the C/O ratio compared to the Solar value. The C/O ratio in gas cannot exceed unity (i.e., a factor of ∼2 enhancement compared to the Solar value) since the major volatile carbon carrier is CO. In contrast, the N/O ratio mainly depends on the relative depletion of N2 and oxygen carriers, and it will increase as each of the oxygen carrier snowlines (H2O, CO2, CO) is crossed. Beyond the CO snowline, there is no strict upper limit to the N/O ratio. The spatial extent of this latter region depends on the relative bond strengths of CO and N2 to ice, but may be quite large (see Section 3). Giant plan- ets that form at wide separations should thus have an excess of nitrogen in their atmospheres, which could be used to trace their formation origin. In addition to N2, a fraction of the nitrogen abundance may also be car- ried by less volatile species such as NH3 (Bottinelli et al. 2010, Mumma & Charnley 2011). The present day N2 in Titan's atmosphere, for example, is thought to originate from accretion of primordial NH3 (Atreya et al. 1978, Mandt et al. 2014). The snowline locations of the main carbon, oxygen and nitrogen carriers strongly depend on the ice grain compo- sition. Very volatile species, such as CO and N2, present binding energies, and therefore snowline locations, that are sensitive to the details of the composition of the icy grain mantles. Spectroscopic observations suggest that up to 90% of the CO ice is frozen in a layer that is thick enough and separated from the H2O ice layer underneath so that it can be considered pure ice (e.g., Pontoppidan et al. 2003). However, if the CO (or N2) ice layer is thin enough (∼monolayer coverage), then it will interact with the H2O ice substrate (e.g., Collings et al. 2003). The ice binding energy is significantly larger in this wa- ter dominated environment than in the pure ice case, as shown by laboratory experiments (Collings et al. 2003, Oberg et al. 2005, Bisschop et al. 2006, Fayolle et al. 2016). This implies that ices in different environments 2 will sublimate at different radii, which will substantially change the disk regions where these volatiles are present in gaseous or solid form (see Section 3.2). In protostellar cores, H2O ice is primarily amorphous (e.g., Williams & Herbst 2002, van Dishoeck et al. 2014). When temper- atures exceed ∼80-90 K, H2O ice acquires a crystalline structure (Schegerer & Wolf 2010). The CO binding en- ergy is larger in an amorphous porous H2O ice environ- ment than in the amorphous compact or crystalline cases (Noble et al. 2012, Fayolle et al. 2016). The N2 binding energy is also larger for a porous versus compact H2O ice substrate. No equivalent studies have yet been per- formed for the deposition of N2 on crystalline H2O ice, but we expect the N2 binding energy in this environment to follow the same trend as the CO binding energy, as CO and N2 display a similar desorption behavior. To explore the range of distances at which CO and N2 in different environments desorb, we consider the limiting cases: pure ices (lowest binding energy) and ices resid- ing on an amorphous porous H2O ice substrate (highest binding energy). We refer to the latter simply as water dominated ices unless noted otherwise. In this work, we expand the coupled drift-desorption model developed in (Piso et al. 2015; hereafter Paper I) by considering additional volatile molecules and abun- dances, ice compositions, as well as nitrogen-to-oxygen (N/O) ratios. This paper is organized as follows. In Sec- tion 2, we review the drift-desorption model developed in Paper I. We discuss the effect of different abundances of the main carbon, oxygen and nitrogen carriers, grain compositions and disk dynamics on snowline locations and the C/N/O ratios in Section 3. We address the im- plications of our results in Section 4 and summarize our findings in Section 5. 2. COUPLED DRIFT-DESORPTION MODEL We begin with a brief review of Paper I's model for the effect of radial drift and viscous gas accretion on volatile snowline locations. We review our disk model in Section 2.1, and summarize our numerical method and results in Section 2.2. 2.1. Disk Model In this work we consider both a static and a viscous disk. The static disk is irradiated by the central star and does not experience redistribution of solids or radial movement of the nebular gas. To quantify the effects of radial drift and gas accretion, we use a viscous disk with M . The a spatially and temporally constant mass flux, viscous disk takes into account radial drift, gas accretion onto the central star, as well as accretion heating. We focus on this disk model which includes all the dynamical and thermal processes we are interested in for the scope of this paper, and do not further consider the other disk models presented in Paper I. Following Chiang & Youdin (2010), the temperature profile for a static disk is Tirr = 120 (r/AU)−3/7 K, (1) where r is the semimajor axis. We use the Shakura & Sunyaev (1973) steady-state disk solution to model the viscous disk. From Paper I, the viscous disk temperature profile is computed as (cid:104) 1 (cid:16) 3Gκ0 M 2M∗µmpΩk (cid:17)1/3(cid:105)4 + T 4 T 4 visc = 4r π2αkBσ (2) Here G is the gravitational constant, κ0 = 2×10−6 is a di- mensionless opacity coefficient, M∗ = M(cid:12) is the mass of the central star, µ = 2.35 is the mean molecular weight of the nebular gas, mp is the proton mass, Ωk =(cid:112)GM(cid:12)/r3 irr. is the Keplerian angular velocity, α = 0.01 is a dimen- sionless coefficient (see below for details), kB is the Boltz- mann constant, and σ is the Stefan-Boltzmann constant. The steady-state disk has an α-viscosity prescription, where the kinematic viscosity is ν = αcH. Here c ≡ (cid:112)kBTvisc/(µmp) is the isothermal sound speed and H ≡ c/Ωk is the disk scale height. We can then determine the gas surface density for a viscous disk as (Shakura & Sunyaev 1973; see also Paper I for a more detailed explanation of these calculations): M 3πν . Σ = (3) M = 10−8M(cid:12) yr−1, consistent with mass We choose flux observations in disks (e.g., Andrews et al. 2010). As described in Paper I, the mass flux rate M and stel- lar luminosity L∗ will vary throughout the disk lifetime (Kennedy et al. 2006, Chambers 2009), in contrast with our simplified model which assumes that both quanti- ties are constant. This effect will be most pronounced in the inner disk ((cid:46) few AU), where accretion heating dominates. We thus acknowledge that the location of the H2O snowline may be determined by the decline in M or L∗ with time, rather than radial drift (see Paper I, Section 2.1 for a more detailed explanation). 2.2. Desorption-Drift Equations and Results The model is described in full in Paper I, here we review and summarize key concepts and results. For a range of initial icy grain sizes composed of a single volatile, we showed in Paper I that the timescale on which these particles desorb is comparable to their radial drift time, as well as to the accretion timescale of the nebu- lar gas onto the central star. We thus have to take into account both drift and gas accretion when we calculate the disk location at which a particle desorbs, since that location may be different from the snowline position in a static disk for a given volatile (see Figure 1 and Oberg et al. 2011b). We determine a particle's final location in the disk by solving the following coupled differential equations: = − 3µxmp ρs NxRdes,x = r, ds dt dr dt (4a) (4b) where s is the particle size, t is time, µx is the mean molecular weight of volatile x, ρs = 2 g cm −3 is the density of an icy particle, Nx ≈ 1015 sites cm −2 is the number of adsorption sites of molecule x per cm−2, Rdes,x is the desorption rate of species x, and r is the particle's radial drift velocity. We calculate R,des and r as follows. The desorption rate Rdes,x (per molecule) is (Hollen- bach et al. 2009) Rdes,x = νx exp (−Ex/Tgrain), (5) where Ex is the adsorption binding energy in units of Kelvin, Tgrain is the grain temperature (assumed to be the same as the disk temperature, see Paper I), and νx = 1.6 × 1011(cid:112)(Ex/µx) s−1 is the molecule's vibra- tional frequency in the surface potential well. We dis- cuss our choices for Ex for the different volatile species in Section 3.1. Following Chiang & Youdin (2010) and Birnstiel et al. (2012), a particle's radial drift velocity can be approxi- mated as (cid:16) τs (cid:17) 1 + τ 2 s r ≈ −2ηΩkr + rgas 1 + τ 2 s , (6) (cid:26) ρss/(ρc), where the first term is the drift velocity in a non-accreting disk and the second term accounts for the radial move- ment of the gas. Here η ≈ c2/(2v2 k), where vk is the Keplerian velocity, and τs ≡ Ωkts is the dimensionless stopping time: ts = 4ρss2/(9ρcλ), s < 9λ/4 Epstein drag s < 9λ/4, Re (cid:46) 1 Stokes drag, (7) where ρ is the disk mid-plane density, λ is the mean free path and Re is the Reynolds number. The gas accretion velocity rgas is determined from M = −2πr rgasΣ, for a fixed M and with Σ given by Equation (3). For a particle of initial size s0, we solve the Equa- tion set (4) with the initial conditions s(t0) = s0 and r(t0) = r0, where t0 is the time at which we start the in- tegration and r0 is the particle's initial location. We stop our simulation after td = 3 Myr, the disk lifetime, since this is roughly the timescale on which planets form, and determine the desorption timescale tdes from s(tdes) = 0, and thus a particle's desorption distance rdes = r(tdes). Our results are insensitive to our choice of t0 as long as t0 (cid:28) td. We note that a particle's size is initially fixed and only changes due to desorption. We thus do not take into account processes such as grain coagulation or fragmentation, which nonetheless occur in disks (e.g., Birnstiel et al. 2012, P´erez et al. 2012). We discuss the effect of these processes on snowline locations in Paper I. As we show in Paper I, a particle of initial size s0 can experience three outcomes after td = 3 Myr: (1) it can remain at its initial location, (2) it can drift towards the host star, then stop without evaporating significantly, and (3) it can completely desorb on a timescale shorter than 3 Myr. Particles in scenarios (1) and (2) are thus not affected by radial drift or gas accretion, and the snow- line locations are those for a static disk. In contrast, the grains in case (3) desorb practically instantaneously and at a fixed particle-size dependent location in the disk, re- gardless of their initial position. The snowline locations for these particles will thus be fixed for a given initial particle size and disk model. We have found that grains with sizes ∼ 0.001 cm (cid:46) s (cid:46) 7 m satisfy this condition for our fiducial disk. 3. RESULTS 3 3.1. Snowlines in a Static Disk: The Importance of Ice Compositions As we note in Section 1, the disk volatile composi- tion and the ice composition determine the location of important snowlines. In this work we focus on the pri- mary carbon, oxygen and nitrogen carriers, i.e. H2O, CO2, CO, N2, and to a lesser extent, CH4 and NH3. Our standard model is based on the median ice abun- dances observed toward Solar-type protostars ( Oberg et al. 2011a), which are nCO2 = 0.29 × nH2O, nCO = 0.38× nH2O, nCH4 = 0.0555× nH2O (hereafter CH4-mid) and nNH3 = 0.055 × nH2O (hereafter NH3-mid). Here nH2O ≈ 10−4 × nH is the total water abundance (van Dishoeck 2006), with nH the hydrogen abundance in the disk midplane. For CO, we also take into account that the observed CO ice only traces some of the CO reservoir due to its high volatility, and similarly to Oberg et al. (2011b) and Paper I we set the total CO abundance to 0.9 × 10−4nH. Finally, we assume that all nitrogen not found in NH3 is in N2 and assume a Solar nitrogen abun- dance, nN = 8 × 10−5nH (Lodders 2003). In effect, this model assumes no chemical evolution between the proto- stellar and disk midplane stages. This is reasonable for material that accretes onto the disk at large radii (Visser et al. 2009), but may overestimate the contribution of the original volatiles to the total volatile budget in the innermost disk. We determine the location of the H2O, CO2, CO, CH4, N2 and NH3 snowlines in our static disk by balancing desorption with readsorption, following Hollenbach et al. (2009). The binding energies of H2O, CO2, CO, CH4, N2 and NH3 as pure ices are 5800 K, 2000 K, 834 K, 1300 K, 767 K and 2965 K, respectively (Fraser et al. 2001, Collings et al. 2004, Fayolle et al. 2016, Garrod & Herbst 2006, Mart´ın-Dom´enech et al. 2014). For our fiducial disk model, these energies correspond to disk temper- atures of 143 K, 48 K, 21 K, 30 K, 18 K and 68 K, respectively. For CO and N2 as water dominated ices, the binding energies are 1388 K and 1266 K, respectively (Fayolle et al. 2016), corresponding to disk temperatures of 34 K and 31 K, respectively. Figure 1 shows the re- sulting snowline locations, assuming CO and N2 pure ices (top panel), and CO and N2 in water dominated ices (bottom panel). The ordinate displays the total carbon, oxygen and nitrogen abundance in solids as a function of the hydrogen total abundance. As expected, the total grain abundance increases with semimajor axis, as more and more species freeze out. Freeze-out at the CO2 and CO snowlines pulls more heavy elements into the grains than in the case of the H2O snowline. Figure 1, bottom panel, displays the snowline locations when CO and N2 are in an amorphous porous water en- vironment (see Section 1). The snowlines move outward by a factor of ∼3.3 if the ices are pure, and by up to a fac- tor of ∼2 for an amorphous compact water substrate (not shown; our estimates are based on the results of Fayolle et al. 2016). The CO snowline moves outward by a factor of ∼1.5 if CO is in a crystalline rather than an amorphous porous water environment (not shown; our estimates are based on the results of Noble et al. 2012), and a simi- lar trend is expected for the N2 snowline (see Section 1). The results of Figure 1 thus represent the limiting cases for the positions of the CO and N2 snowlines, for differ- 4 Fig. 1. -- The total carbon, nitrogen and oxygen abundance in solids as a function of semimajor axis in a static disk, for CO and N2 as pure ices (top panel) and water dominated ices (bottom panel). Relevant volatile snowlines are marked by the vertical dashed lines. The grain abundances are calculated as a function of the observed median CH4 and NH3 abundances in protostellar cores. The total grain abundance increases with semimajor axis as more and more species freeze out. ent compositions of the icy grains and morphology of the water ice substrate. This variation in snowline locations changes the chemical abundances both in gas and dust throughout the disk, directly affecting the compositions of nascent giant planets forming in situ. In our simple model, we ignore the effects of CO and N2 entrapment in water ice through clathrate formation or other processes. Theoretical models aimed at explaining the composition of comet 67P/Churyumov-Gerasimenko suggest that a small fraction of the total CO and N2 reservoir may be trapped in clathrates, and only released upon water sub- limation (Lectez et al. 2015, Mousis et al. 2016). In this case, the CO and N2 snowlines would be closer to the star than in the pure ice case. 3.2. C/N/O Ratios in Static Disks In this section we determine the C/O and N/O ra- tios in gas and dust throughout our static disk, and to what extent they are affected by the presence of CH4 and NH3 over the full range of observed CH4 and NH3 abundances toward low-mass protostars. In this section we only consider pure ices, i.e. ices that are layered on a silicate mantle. In reality, the icy grain will have a layered structure with volatiles residing on top of a H2O ice substrate. However, if the volatile ice layer is thick enough and separated from the H2O ice layer, the inter- action between H2O and the other volatile species will be minimal and thus the ices can be considered pure rather than water dominated (see also Section 1). We explore the parameter space of possible CH4 abun- dances by assuming three different scenarios: (1) no CH4, (2) CH4-mid, and (3) the maximum CH4 observed abun- dance (hereafter CH4-max), nCH4−max = 0.13 × nH2O ( Oberg et al. 2008). Since the abundance of carbon grains is uncertain, we assume that all the carbon that is not in the form of CH4, CO and CO2 is in carbon grains, so that we reproduce the Solar C/O ratio (gas+dust) of 0.54. Figure 2 shows the C/O ratio in gas and dust as a function of semimajor axis in a static disk: no CH4 (top panel), CH4-mid (middle panel) and CH4-max (bottom panel). As in Oberg et al. (2011b) and Paper I, a gaseous C/O ratio of unity can be achieved between the CO2 and CO snowlines, where oxygen gas is significantly depleted. The gas-phase C/O ratio may be further enhanced be- tween the CO2 and CH4 snowlines due to the presence of additional carbon gas from CH4. In this region, the C/O ratio increases by 3% for CH4-mid and by 8% for CH4-max, as displayed in the middle and bottom panels of Figure 2. Based on the range of observed CH4 proto- stellar abundances, its presence in the disk only modestly affects the C/O ratio. We assume that the main nitrogen-bearing species are N2 and NH3, since other volatiles that contain nitrogen have significantly lower abundances in comparison (e.g., Mumma & Charnley 2011). Similarly to the case of CH4, we explore the parameter space of possible NH3 abun- dances using observations toward low-mass protostars, 12345678H2ONH3CO2CH4CO(p)N2(p)pure ice grains(nC+nN+nO) / (104×nH)101100101102r [AU]1234567CO(m)N2(m)on H2O ice grains 5 Fig. 2. -- The C/O ratio in gas (solid lines) and dust (dashed lines) as a function of semimajor axis in a static disk, assuming no carbon is present in the form of CH4 (top panel), the median observed CH4 abundance is assumed (middle panel), and the max- imum observed CH4 abundance is assumed (bottom panel). The C/O estimates are performed assuming that the CO ices are in pure form. The vertical dotted lines mark the snowline locations of the main C and O carriers. The horizontal dotted lines represent the stellar C/O value. The presence of methane only modestly in- creases the C/O ratio in gas between the CO2 and CH4 snowlines. as follows: (1) no NH3, (2) NH3-mid, and (3) the maxi- mum observed NH3 abundance nNH3−max = 0.15× nH2O (Bottinelli et al. 2010). In each case, the N2 abundance then simply follows as nN2 = (nN − nNH3)/2. Figure 3 shows the snowline locations of the main oxy- gen and nitrogen carriers and the N/O ratio in gas and dust as a function of semimajor axis in a static disk, for our three choices of the NH3 abundance: no NH3 (top panel), NH3-mid (middle panel) and NH3-max (bottom panel). For comparison, the horizontal dotted lines show the average N/O ratio in the disk. As expected, the gaseous N/O ratio generally exhibits an increasing trend towards the outer disk as more oxygen gas is depleted, with small decreases between the NH3 and CO2 snow- lines (by 6% for NH3-mid and by 18% for NH3-max, respectively) due to NH3 freeze-out. While the pres- ence of NH3 only moderately affects our results for the N/O ratio, NH3 is important since otherwise the nitro- Fig. 3. -- The N/O ratio in gas (solid lines) and dust (dashed lines) as a function of semimajor axis in a static disk, assuming no nitrogen is present in the form of NH3 (top panel), the median observed NH3 abundance is assumed (middle panel), and the max- imum observed NH3 abundance is assumed (bottom panel). The N/O estimates are performed assuming that the CO and N2 ices are in pure form. The vertical dotted lines mark the snowline loca- tions of the main C,O and N carriers. The horizontal dotted lines represent the average N/O value in the disk. The gas-phase N/O ratio is enhanced by a factor of two between the H2O and CO2 snowlines compared to its average value, and by a factor of three between the CO2 and CO snowlines. The arrows mark a highly elevated N/O ratio in gas between the CO and N2 snowlines due to the depletion of oxygen gas in this region. The presence of NH3 moderately decreases the N/O ratio in gas between the NH3 and CO2 snowlines. gen content in solid bodies would be more depleted than is observed for comets and asteroids (Wyckoff et al. 1991, Mumma & Charnley 2011, Bergin et al. 2015). The gas-phase N/O ratio is enhanced by a factor of two outside the H2O snowline compared to its average value, by more than a factor of three between the CO2 and CO snowlines, and by orders of magnitude between the CO and N2 snowlines. This latter region can span tens of AU depending on disk parameters and the relative CO and N2 ice binding environment. This N/O enhancement is more pronounced than the C/O gas phase enhancement of a factor of two in the outer disk (see Figure 2). 0.00.20.40.60.81.01.2nCH4=0H2OCO2CH4CO(p)0.00.20.40.60.81.01.2C/O rationCH4=5.55%nH2O101100101102r [AU]0.00.20.40.60.81.01.2nCH4=13%nH2O0.00.10.20.30.40.50.6nNH3=0H2OCO2NH3CO(p)N2(p)0.00.10.20.30.40.50.6N/O rationNH3=5.5%nH2O101100101102r [AU]0.00.10.20.30.40.50.6nNH3=15%nH2O 6 3.3. C/N/O Ratios in Dynamic Disks Here we use the model of Section 2 to estimate the movement of the CO and N2 snowlines for different grain morphologies in a viscous disk. Figure 4 shows the H2O, CO2 and CO snowline locations for particles with initial sizes ∼ 0.05 cm (cid:46) s (cid:46) 7 m as well as estimates for the C/O ratio in gas and dust in a viscous disk, with the CO snowline calculated under different grain morphologies as noted above. We assume there is no carbon in the form of CH4. The true snowline for particles that desorb outside the static snowline is the static snowline itself, hence desorbing particles with s < 0.05 cm do not form true snowlines. If the CO binding environment is known, the CO snowline moves inward by up to ∼50 % compared to a static disk for each case (pure and water dominated ices) due to disk dynamics. The full range of potential CO snowlines taking into account both ice compositions and disk dynamics span ∼8.7 AU to ∼61 AU, which is a factor of ∼7 difference. This implies that gas phase C/O ratios of order unity may be reached in the giant planet forming zone, and the CO snowline may be inside 10 AU for certain disk parameters. the N/O ratio (see Figure 3), and since we are primarily interested in the N2 snowline locations rather than exact values for the N/O ratio. The innermost N2 snowlines in the viscous disk, created by particles with s ∼ 7 m for our fiducial model, are located at rN2,pure ≈ 42 AU for N2 as pure ice and at rN2,water ≈ 11 AU for N2 in water dominated ices. Thus for each case (pure versus water dominated ices), the N2 snowline moves inward by up to 50% due to disk dynamics. By taking into account both ice compositions and disk dynamics, the full range of potential N2 snowlines span ∼11 to ∼79 AU, which is a factor of ∼7 difference. Similarly to the case for CO, the N2 snowline may be close to 10 AU for certain disk models. Fig. 5. -- N/O ratio estimates in gas (solid lines) and dust (dashed lines) as function of semimajor axis in a viscous disk, for CO and N2 as pure ices (top panel) or as water dominated ices (bottom panel). The H2O, CO2, CO and N2 snowlines are shown for particles with initial sizes ∼ 0.05 cm (cid:46) s (cid:46) 7 m as indicated by the color bar. The N/O ratio in a static disk (black lines) is shown for comparison. The arrows show that the N/O ratio in gas will decrease inside the H2O and CO2 snowlines in the viscous disk, as the relative fluxes of the desorbed icy particles and the overall nebular gas will cause an excess of oxygen gas inside these snowlines (see Paper I for details). Radial drift and gas accretion move the N2 snowline inward by up to ∼50% compared to a static disk. The presence of N2 in a water ice environment rather than as pure ice moves the N2 snowline significantly inward by ∼70%. Taken together, disk dynamics and ice compositions move the N2 snowline inward by a factor of ∼7. The results of an enhanced gas-phase N/O ratio between the H2O and CO snowlines compared to its average value, and of highly elevated N/O ratios in gas between the CO and N2 snowlines (see Figure 3), are preserved. 4. DISCUSSION This study shows that the gas-phase N/O ratio in pro- toplanetary disks is considerably enhanced throughout most of the disk midplane compared to its average value. As demonstrated in Figure 6, the gaseous N/O ratio is enhanced by a factor of two beyond the H2O snowline, by more than a factor of three between the CO2 and CO snowlines, and by several orders of magnitude between the CO and N2 snowlines. Thus constraining the N/O ratio in a giant planet atmosphere could be used to trace its formation origins. Fig. 4. -- C/O ratio estimates in gas (solid lines) and dust (dashed lines) as function of semimajor axis in a viscous disk, for CO as pure ice (top panel) or as water dominated ices (bottom panel). The H2O, CO2 and CO snowlines are shown for particles with initial sizes ∼ 0.05 cm (cid:46) s (cid:46) 7 m as indicated by the color bar. The C/O ratio in a static disk (black lines) is shown for comparison. The arrows show that the C/O ratio in gas will decrease inside the H2O and CO2 snowlines in the viscous disk, as the relative fluxes of the desorbed icy particles and the overall nebular gas will cause an excess of oxygen gas inside these snowlines (see Paper I for details). The presence of CO in a water ice environment rather than as pure ice moves the CO snowline significantly inward by ∼70%. Taken together, disk dynamics and ice compositions move the CO snowline inward by a factor of ∼7. Figure 5 shows the H2O, CO2, CO and N2 snowline locations in a viscous disk for the same range of initial particle sizes as in Figure 4, and with the CO and N2 snowlines calculated assuming different grain morpholo- gies as explained above, as well as estimates for the N/O ratio throughout the disk. For simplicity, we assume that all nitrogen is the form of N2. This choice is justified since the presence of some NH3 only moderately changes 0.00.20.40.60.81.01.2C/O ratiopure icegasdust101100101102r [AU]0.00.20.40.60.81.0on H2O ice101100101102s [cm]0.00.10.20.30.40.50.60.7N/O ratiopure icegasdust101100101102r [AU]0.00.10.20.30.40.50.6on H2O ice101100101102s [cm] Theoretical models of the magnitude and role of N/O (and N/C) ratios in exoplanet atmospheres are needed in order to use these ratios as probes for a planet's forma- tion location. Models that explore the effect of varying the C/O ratio in exoplanet atmospheres exist in litera- ture, and they display a large and observable effect on gas giant envelope chemistry (Lodders 2009, Molli`ere et al. 2015). However, no similar model explorations exist for the effect of N/O and C/N/O ratios, and both are needed to exploit this potential constraint. Given the existence of such theoretical models, measurements of the N/O ra- tio in planetary envelopes may be possible to infer from atmospheric compositions of nitrogen versus carbon and oxygen bearing species. Nitrogen carriers have not been targeted so far due to lack of instrument sensitivity, but such observations and detections are likely in the near fu- ture with the advent of JWST (e.g., NH3, Greene et al. 2016). The N/O ratio enhancement is larger than that of the gas phase C/O ratio throughout most of the disk. Thus measurements of an enhanced C/O ratio in an exo- planet atmosphere could be corroborated (disproved) by measurements of enhanced (non-enhanced) N/O ratios. Moreover, Figure 6 shows that giant planets that have formed in situ between the H2O and CO snowlines are expected to present elevated both C/O and N/O ratios in their atmospheres, whereas planets between the CO and N2 snowlines will have a highly enhanced N/O ratio in their atmospheres, but not C/O. Fig. 6. -- Gas phase C/O (blue curve) and N/O (red curve) ratios divided by the average C/O and N/O ratio in a static disk, assum- ing CO and N2 are pure ices, and there is no CH4 or NH3. The dashed vertical lines mark the H2O, CO2, CO and N2 snowlines. The arrow indicates that the N/O ratio is enhanced by orders of magnitude compared to its average value between the CO and N2 snowlines. The gaseous N/O ratio is enhanced throughout most of the disk, and more enhanced than the C/O ratio. Due to disk dynamics and ice compositions, the lo- cations of the CO and N2 snowlines, and thus the disk regions with highly elevated gas phase N/O and C/O ratios, are uncertain and may span tens of AU. Both ice morphologies discussed in this study, pure and water dominated ices, are plausible in protoplanetary disks and depend on whether H2O and CO ices formed on similar timescales or successively (e.g., Garrod & Pauly 2011). Observations of protostellar cores show that a large frac- tion of CO is bound in a pure ice multilayer (Pontoppidan 7 et al. 2003), but theoretical models also suggest an icy mantle structure where CO resides on a H2O ice layer (e.g., Collings et al. 2003). One can also imagine a sce- nario where CO is in a water binding environment and N2 is not. This could be attributed to the fact that H2O may bind preferentially to CO than N2, since both H2O and CO are polar molecules while N2 is not. It is also possible for N2 ices to form later than CO (e.g., Pagani et al. 2012), and thus be deposited on the outer layers of the icy mantles which are typically water poor (e.g., Gar- rod & Pauly 2011). The impact of the ice environment on the snowline location is much smaller in the case of CO2 and NH3, as their binding energies and behavior are closer to that of H2O. No detailed measurements for the CH4 binding energy in a water environment exist so far, but due to its low desorption temperature a similar be- havior to that of CO and N2 would be expected. While the presence of some carbon in the form of CH4 only modestly affects our results, CH4 may become important in disks where a large fraction of the CO abundance has been converted into hydrocarbons (e.g., Du et al. 2015). Changes in stellar luminosity (e.g., Kennedy et al. 2006) and gas mass accretion rate (e.g., Chambers 2009), as well as the evolution of icy dust particles due to grain growth and fragmentation (e.g., Birnstiel et al. 2012), may introduce additional uncertainties in the snowline locations, and thus the C/N/O ratios. Moreover, the diffusion of vapor across the snowlines following the cold finger effect (Stevenson & Lunine 1988, Cyr et al. 1998) will change the shape of the C/O and N/O curves and therefore the magnitude of the C/N/O ratios between different snowlines. The effect of dynamical processes on snowline locations is discussed in more detail in Paper I, Section 5.2. Given the number of uncertainties in snow- line locations, detections of snowlines in a sample of disks at different evolutionary stages are needed to provide ob- servational constraints on the relative importance of ice compositions and disk dynamics in setting snowline loca- tions. The uncertainties in snowline locations caused by disk dynamics, ice compositions, and other effects out- lined above can be resolved in extreme cases, such as a detection of a CO snowline at a temperature correspond- ing to pure CO ice desorption in a static disk (e.g., Qi et al. 2013 at ∼ 17 K) or CO desorption from a water dominated ice in a dynamic disk. In intermediate cases it is more difficult to resolve the relative importance of ice compositions and disk dynamics. For example, the CO snowline in HD 163296 is at a higher temperature of ∼ 25 K (Qi et al. 2015), which could be caused ei- ther by CO being in a water dominated environment or by dynamical effects that push the CO snowline inward. Detections of multiple snowlines in the same disk could potentially break this degeneracy. Uncertainties in snowline locations of this magnitude also affect interpretations of Solar system observations. Recent measurements of nitrogen abundance in comet 67P/Churyumov-Gerasimenko found a N2/CO ratio ∼ 10−3 (Rubin et al. 2015). A low N2/CO ratio is consis- tent with comets having formed inside the N2 snowline where N2 is still in the gas phase. However, it is also possible that the measured N2 abundance in 67P may be due to post-formation processes such as radiogenic heating (Rubin et al. 2015), and thus may not reflect 101100101102r [AU]1.01.52.02.53.0H2OCO2CON2(C/O)gas / (C/O)avg(N/O)gas / (N/O)avg 8 the comet's primordial composition. Theoretical models suggest that Jupiter-family comets, such as 67P, origi- nate from the Kuiper belt (Duncan & Levison 1997; but see Rubin et al. 2015 for alternative formation scenarios for 67P). It is thus possible, in principle, to use measure- ments of the N2 abundance in Jupiter-family comets to determine where the N2 snowline was located in our Solar system. However due to the uncertainty in the calculated location of the N2 snowline (see Section 3.3), as well as the uncertainty of the formation zone of Jupiter-family comets (anywhere between 5 and >30 AU; Pontoppidan et al. 2014), more detailed modeling is needed. 5. SUMMARY In this paper we explore the role of icy grain com- positions and disk dynamics on the snowline locations of major volatile carrier molecules and the C/N/O ra- tios in protoplanetary disks. We enhance the coupled drift-desorption model developed in Piso et al. (2015) by adding more carbon- and nitrogen-bearing species into our framework, and by considering different binding ice environments. Our results can be summarized as follows: 1. Due to the high volatility of N2, the gaseous N/O ratio outside the H2O snowline is enhanced by a factor of two compared to its average value, by more than a factor of three between the CO2 and CO snowlines, and by many orders of magnitude between the CO and N2 snowlines due to the com- plete depletion of oxygen gas in this region. This enhancement is more pronounced than in the case of the gas-phase C/O ratio, which is increased by at most a factor of two compared to the stellar value. 2. The effect of CH4 and NH3 on the C/O and N/O ratios is small, even when we consider the maxi- mum observed CH4 and NH3 abundances in proto- stellar cores. In this scenario, the gas phase C/O ratio increases by 8% between the CO2 and CH4 snowlines, and the gaseous N/O ratio decreases by 18% between the NH3 and CO2 snowlines. In both cases, large gas phase C/O and N/O ratios in the outer disk are preserved. 3. Grain composition sensitively affects the CO and N2 snowline locations. If CO and N2 reside in wa- ter dominated rather than pure ices, their snow- lines move inward by up to ∼70 %. This effect is separate from that of radial drift and viscous gas accretion, which also cause an inward movement of the CO and N2 snowlines by up to ∼50 %. 4. The locations of the CO and N2 snowlines are un- certain when we consider both viscous versus static disks, and pure versus water dominated ices. The snowlines in a viscous disk with CO or N2 in a wa- ter environment are by up to a factor of ∼7 closer to the host star that in a static disk with CO or N2 as pure ices. Our results have direct consequences for the composi- tion of nascent giant planets. The considerable inward movement of the CO and N2 snowlines due to the ice grains being water dominated rather than pure ices im- plies than giant planets with high C/O and/or N/O ra- tios in their atmospheres may form closer in than pre- viously predicted by theoretical models. Moreover, our model shows that wide separation gas giants may have an excess of nitrogen in their envelopes, which may be used to trace their origins. We thank the anonymous referee for helpful comments and suggestions. This work is supported by a Simons Collaboration on the Origins of Life (SCOL) investiga- tor award to KIO. JP was supported by the Banneker institute. REFERENCES Ali-Dib, M., Mousis, O., Petit, J.-M., & Lunine, J. I. 2014, ApJ, Andrews, S. M., Wilner, D. J., Hughes, A. M., Qi, C., & Dullemond, C. P. 2010, ApJ, 723, 1241 Atreya, S. K., Donahue, T. M., & Kuhn, W. R. 1978, Science, 2009, ApJ, 690, 1497 785, 125 201, 611 Bergin, E. A., Blake, G. A., Ciesla, F., Hirschmann, M. M., & Li, J. 2015, Proceedings of the National Academy of Science, 112, 8965 Birnstiel, T., Klahr, H., & Ercolano, B. 2012, A&A, 539, A148 Bisschop, S. E., Fraser, H. J., Oberg, K. I., van Dishoeck, E. F., & Schlemmer, S. 2006, A&A, 449, 1297 Boogert, A. C. A., Gerakines, P. A., & Whittet, D. C. B. 2015, Garrod, R. T., & Pauly, T. 2011, ApJ, 735, 15 Greene, T. P., Line, M. R., Montero, C., et al. 2016, ApJ, 817, 17 Henning, T., & Semenov, D. 2013, Chemical Reviews, 113, 9016 Hollenbach, D., Kaufman, M. J., Bergin, E. A., & Melnick, G. J. Kennedy, G. M., Kenyon, S. J., & Bromley, B. C. 2006, ApJ, 650, Lectez, S., Simon, J.-M., Mousis, O., et al. 2015, ApJ, 805, L1 Lodders, K. 2003, ApJ, 591, 1220 -- . 2009, ArXiv e-prints, arXiv:0910.0811 Madhusudhan, N., Amin, M. A., & Kennedy, G. M. 2014, ApJ, Mandell, A. M., Bast, J., van Dishoeck, E. F., et al. 2012, ApJ, L139 794, L12 747, 92 788, L24 ARA&A, 53, 541 718, 1100 Bottinelli, S., Boogert, A. C. A., Bouwman, J., et al. 2010, ApJ, Mandt, K. E., Mousis, O., Lunine, J., & Gautier, D. 2014, ApJ, Chambers, J. E. 2009, ApJ, 705, 1206 Chiang, E., & Youdin, A. N. 2010, Annual Review of Earth and Planetary Sciences, 38, 493 Collings, M. P., Anderson, M. A., Chen, R., et al. 2004, MNRAS, 354, 1133 Collings, M. P., Dever, J. W., Fraser, H. J., & McCoustra, M. R. S. 2003, Ap&SS, 285, 633 Cyr, K. E., Sears, W. D., & Lunine, J. I. 1998, Icarus, 135, 537 Du, F., Bergin, E. A., & Hogerheijde, M. R. 2015, ArXiv e-prints, L33 Mart´ın-Dom´enech, R., Munoz Caro, G. M., Bueno, J., & Goesmann, F. 2014, A&A, 564, A8 Molli`ere, P., van Boekel, R., Dullemond, C., Henning, T., & Mordasini, C. 2015, ApJ, 813, 47 Mousis, O., Fletcher, L. N., Lebreton, J.-P., et al. 2014, Planet. Space Sci., 104, 29 Mousis, O., Lunine, J. I., Luspay-Kuti, A., et al. 2016, ApJ, 819, Mumma, M. J., & Charnley, S. B. 2011, ARA&A, 49, 471 Noble, J. A., Congiu, E., Dulieu, F., & Fraser, H. J. 2012, arXiv:1506.03510 Duncan, M. J., & Levison, H. F. 1997, Science, 276, 1670 Fayolle, E. C., Balfe, J., Loomis, R., et al. 2016, ApJ, 816, L28 Fraser, H. J., Collings, M. P., McCoustra, M. R. S., & Williams, D. A. 2001, MNRAS, 327, 1165 Garrod, R. T., & Herbst, E. 2006, A&A, 457, 927 MNRAS, 421, 768 ApJ, 678, 1032 -- . 2011a, ApJ, 740, 109 Oberg, K. I., Boogert, A. C. A., Pontoppidan, K. M., et al. 2008, L16 621, L33 760, L17 408, 981 9 Rodgers, S. D., & Charnley, S. B. 2002, MNRAS, 330, 660 Rubin, M., Altwegg, K., Balsiger, H., et al. 2015, Science, 348, 232 Schegerer, A. A., & Wolf, S. 2010, A&A, 517, A87 Shakura, N. I., & Sunyaev, R. A. 1973, A&A, 24, 337 Stevenson, D. J., & Lunine, J. I. 1988, Icarus, 75, 146 van Dishoeck, E. F. 2006, Proceedings of the National Academy van Dishoeck, E. F., Bergin, E. A., Lis, D. C., & Lunine, J. I. 2014, Protostars and Planets VI, 835 Visser, R., van Dishoeck, E. F., Doty, S. D., & Dullemond, C. P. 2009, A&A, 495, 881 Williams, D., & Herbst, E. 2002, Surface Science, 500, 823 Wyckoff, S., Tegler, S. C., & Engel, L. 1991, ApJ, 367, 641 Zhang, K., Pontoppidan, K. M., Salyk, C., & Blake, G. A. 2013, ApJ, 766, 82 Oberg, K. I., Murray-Clay, R., & Bergin, E. A. 2011b, ApJ, 743, Oberg, K. I., van Broekhuizen, F., Fraser, H. J., et al. 2005, ApJ, Owen, T., Mahaffy, P. R., Niemann, H. B., Atreya, S., & Wong, M. 2001, ApJ, 553, L77 Pagani, L., Bourgoin, A., & Lique, F. 2012, A&A, 548, L4 P´erez, L. M., Carpenter, J. M., Chandler, C. J., et al. 2012, ApJ, of Science, 103, 12249 Piso, A.-M. A., Oberg, K. I., Birnstiel, T., & Murray-Clay, R. A. 2015, ApJ, 815, 109 Pontoppidan, K. M., Salyk, C., Bergin, E. A., et al. 2014, Protostars and Planets VI, 363 Pontoppidan, K. M., Fraser, H. J., Dartois, E., et al. 2003, A&A, Qi, C., Oberg, K. I., Andrews, S. M., et al. 2015, ApJ, 813, 128 Qi, C., Oberg, K. I., Wilner, D. J., et al. 2013, Science, 341, 630
1202.2218
2
1202
2012-02-16T21:19:41
Spectral properties of oscillatory and non-oscillatory {\alpha}^2-dynamos
[ "astro-ph.EP", "astro-ph.SR" ]
The eigenvalues and eigenfunctions of a linear {\alpha}^{2}-dynamo have been computed for different spatial distributions of an isotropic \alpha-effect. Oscillatory solutions are obtained when \alpha exhibits a sign change in the radial direction. The time-dependent solutions arise at so called exceptional points where two stationary modes merge and continue as an oscillatory eigenfunction with conjugate complex eigenvalues. The close proximity of oscillatory and non-oscillatory solutions may serve as the basic ingredient for reversal models that describe abrupt polarity switches of a dipole induced by noise. Whereas the presence of an inner core with different magnetic diffusivity has remarkable little impact on the character of the dominating dynamo eigenmodes, the introduction of equatorial symmetry breaking considerably changes the geometric character of the solutions. Around the dynamo threshold the leading modes correspond to hemispherical dynamos even when the symmetry breaking is small. This behavior can be explained by the approximate dipole-quadrupole degeneration for the unperturbed problem. More complicated scenarios may occur in case of more realistic anisotropies of \alpha- and \beta-effect or through non-linearities caused by the back-reaction of the magnetic field (magnetic quenching).
astro-ph.EP
astro-ph
June 16, 2018 3:22 Geophysical and Astrophysical Fluid Dynamics gieseckeoscalpha2 Geophysical and Astrophysical Fluid Dynamics Vol. 00, No. 00, January 2008, 1 -- 11 2 1 0 2 b e F 6 1 . ] P E h p - o r t s a [ 2 v 8 1 2 2 . 2 0 2 1 : v i X r a Spectral properties of oscillatory and non-oscillatory α2-dynamos. A. Giesecke, F. Stefani, G. Gerbeth Helmholtz-Zentrum Dresden-Rossendorf, Dresden, Germany (Received 00 Month 200x; in final form 00 Month 200x ) The eigenvalues and eigenfunctions of a linear α2-dynamo have been computed for different spatial distributions of an isotropic α-effect. Oscillatory solutions are obtained when α exhibits a sign change in the radial direction. The time-dependent solutions arise at so called exceptional points where two stationary modes merge and continue as an oscillatory eigenfunction with conjugate complex eigenvalues. The close proximity of oscillatory and non-oscillatory solutions may serve as the basic ingredient for reversal models that describe abrupt polarity switches of a dipole induced by noise. Whereas the presence of an inner core with different magnetic diffusivity has remarkable little impact on the character of the dominating dynamo eigenmodes, the introduction of equatorial symmetry breaking considerably changes the geometric character of the solutions. Around the dynamo threshold the leading modes correspond to hemispherical dynamos even when the symmetry breaking is small. This behavior can be explained by the approximate dipole-quadrupole degeneration for the unperturbed problem. More complicated scenarios may occur in case of more realistic anisotropies of α- and β-effect or through non-linearities caused by the back-reaction of the magnetic field (magnetic quenching). Keywords: Dynamo, alpha-effect, mean-field-theory, oscillatory fields, reversal 1 Introduction Oscillatory or reversing magnetic fields driven by a flow of an electrically conducting fluid are a well known astrophysical phenomenon. Thus the solar magnetic field regularly oscillates on a characteristic 22 yrs time scale whereas the dominating dipole component of the Earth's magnetic field irregularly changes its orientation every few 100 kyrs, conducting a so called reversal. Sign changes of the magnetic field also have been observed in the Cadarache von-K´arm´an-Sodium (VKS) dynamo, and depending on the difference of the rotation rates of the two flow driving impellers various regimes with oscillatory and/or chaotic behavior can be obtained (Berhanu et al. 2007). Whereas the time scale of the solar cycle can be reproduced using simple αΩ-dynamo models, an explanation of the observed equatorial migration requires further assumptions, e.g. meridional flow or negative radial shear at the surface. Regarding the geodynamo, three dimensional simulations of the magnetohydrodynamic equations as well as mean field models have been able to reproduce essential features of the Earth's magnetic field (Hoyng et al. 2002, Christensen 2011), but others, for example the very nature of the reversal mechanism or the large variation of reversal rates (McFadden and Merrill 1986) still are unsolved issues. A prominent feature of this paleomagnetic reversal frequency distribution is the occurrence of a few very long periods (>∼ 20Myrs) during which no reversal occurred at all (so called superchrons; Harland et al., 1982). Another surprising but less known property are deviations of the distribution of inter-reversal time periods from an ideal Poisson distribution (Carbone et al. 2006, Sorriso-Valvo et al. 2007) indicating long term correlations in the reversal trigger mechanism(s). In order to reliably disentangle internal reversal trigger mechanisms (e.g. intrinsic changes of the fluid flow pattern in the liquid outer core) from external sources (e.g. precession or changes in the heat flux through the core mantle boundary) numerical simulations are required that cover sufficient long time periods to allow for a large number of reversal events. In addition, numerical models should also be capable to incorporate the mechanisms that trigger individual reversals and reproduce general reversal ∗Corresponding author. Email: [email protected] Geophysical and Astrophysical Fluid Dynamics ISSN: 0309-1929 print/ISSN 1029-0419 online c(cid:13) 2008 Taylor & Francis DOI: 10.1080/0309192YYxxxxxxxx June 16, 2018 3:22 Geophysical and Astrophysical Fluid Dynamics gieseckeoscalpha2 2 characteristics such as duration and field geometry during the actual reversal. Mean field models are ideally suited to execute long term simulations of the (mean field) induction equation because they are computational cheap. The price for this capability is a strong simplification by parameterizing induction effects of the turbulent small scale flow e.g. in terms of the α-effect (Krause and Radler 1980). Hence, any internal changes of the small scale turbulence (e.g. changes in statistical properties of the convection caused by gradual growth of the solid inner core) are suspected to encroach/expand into the comparably simple mean field coefficients. Nevertheless, mean field modelling of reversing magnetic fields has been remarkably successful (Hoyng et al. 2002, Stefani and Gerbeth 2005, Fischer et al. 2009). The present work is based on the idea that a polarity change is part of an oscillation of the dominant dipole mode (Stefani et al. 2007). A consequential model for irregularly occurring reversals requires a proximity of an oscillatory and a non-oscillatory branch so that a single transition between both states might be induced by noise. This is substantiated by the fact that polarity reversals in numerical dynamos are generally found in intermediate parameter regions between stable (dipolar) dynamos with small fluctuations and highly fluctuating and unstable dynamos (Olson and Christensen 2006). The mean flow in the Earth's fluid core most probably is weak so that usual geodynamo models are based on an α2 mechanism where an α-effect (re-)generates the poloidal field from the poloidal field and vice versa. A general requirement for the occurrence of oscillating eigenmodes in α2-dynamos is ∇α 6= 0 (Radler and Brauer 1987). However, for a long time it was assumed that the leading eigenmode in simple α2-dynamos is non-oscillatory and dominant oscillatory solutions are a curiosity that requires a rather elaborate configuration (Rudiger et al. 2003). Recently, it has been discovered that oscillatory α2-dynamos are quite common when the radial profile of a spherically symmetric, isotropic α exhibits a sign change (Stefani and Gerbeth 2003) and a stringent mathematical treatment of this model yields very general conditions for the occurrence of oscillating solutions (Gunther et al. 2010). Meanwhile oscillating solutions have also been found in direct numerical simulations of thermal convection in a spherical shell (Schrinner et al. 2011) or in a spherical wedge geometry with random helical forcing (Mitra et al. 2010). Such oscillating α2-dynamos might provide an alternative approach for the solar dynamo, e.g. the model of Mitra et al. (2010) exhibits equatorward migration without the requirement of meridional circulation or negative radial shear (which in the sun only occurs in a rather thin layer located close to the surface). Here, we examine the behavior of axisymmetric eigenmodes generated in a two-dimensional kinematic α2- dynamo with isotropic α-effect. The focus of our examinations is on the spectrum of eigenvalues in terms of growth rates and oscillation frequencies for two different radial profiles of the α-effect. For this purpose we pick up the model of Giesecke et al. (2005a) where a sinusoidal radial α distribution is assumed. We show that for this α-distribution the leading modes around the dynamo threshold are oscillating independent of inner core conductivity or latitudinal perturbations that would provide an equatorial symmetry breaking. In extension to Giesecke et al. (2005a), where only the two dominating oscillating modes around the onset of dynamo action were discussed, here we investigate the emergence/disappearance of oscillating modes at so called exceptional points in more detail and briefly identify the fundamental differences in the radial structure of oscillatory and non-oscillatory eigenmodes. The work aims at building a bridge between original mean field models that have been developed in the sixties of the last Century, and recent results from observational and numerical approaches demonstrating the valuable impact of mean field theory for the development of a geodynamo reversal model. 2 Equations and Method The basic assumption of the mean field approach is a separation of the velocity field u and the magnetic field B which are split into mean parts hui, hBi and fluctuating parts u′, B′ according to B = hBi + B ′ and u = hui + u′, (1) where h·i represents an appropriate space- and time average so that the Reynolds averaging rules apply. Then the temporal development of the mean magnetic field hBi is described by the mean field induction June 16, 2018 3:22 Geophysical and Astrophysical Fluid Dynamics gieseckeoscalpha2 equation ∂t hBi = ∇ × (hui × hBi + E − η∇ × hBi) 3 (2) with the turbulent electromotive force (EMF) E = hu′ × B ′i that describes the average induction action of (unresolved) small scale flow perturbations. E results from the interaction of the fluctuating velocity field u′ with the mean magnetic field hBi and is formally represented by a linear functional of hui, hBi and the statistical properties of u′ (Krause and Radler 1980). Assuming that the mean quantities vary only slightly around a certain space-time point, only contribution from a certain neighborhood must be taken into account and the EMF can be written as a Taylor expansion: Ei = (cid:10)u′ × B ′(cid:11)i = αij hBji + βijk ∂ hBki ∂xj + ... (3) where the coefficients αij and βijk depend on the properties of the turbulent fluctuations u′. In general, the turbulence is anisotropic so that α- and β-effect are described by tensors of 2nd and 3rd rank, respectively. Here we restrict ourselves to the simplest case of isotropic turbulence so that α and β are given by αij = α0δij , βijk = −βǫijk (4) (5) so that α and β are prescribed by scalar quantities. Note that for a non-vanishing α-effect the turbulence additionally must be non-mirrorsymmetric. With the definition of the turbulent diffusivity ηT = η + β (6) and vanishing mean flow hui = 0 the mean field induction equation simplifies to an equation describing the temporal development of an α2-dynamo: ∂tB = ∇ × (αB − ηT∇ × B) (7) (the brackets around the mean field B from now on are dropped for simplicity). Equation (7) is linear in B and the ansatz B = B(r)eγt leads to a linear eigenvalue problem MB = γB (8) with the matrix M containing the dynamo operator from the right hand side of (7), and the eigenvalue γ = σ+iν (growth rate σ and frequency ν). For uniform distributions of α and ηT (semi-) analytic solutions are known (e.g. Krause and Radler, 1980). For more complex flows and spatial distributions of α and η a numerical solution of the eigenvalue problem is required (e.g. Roberts and Stix 1972, Dudley and James 1989). Here we solve (8) in a sphere surrounded by a non-conducting vacuum applying the method pre- sented in Schrinner et al. (2010). The approach is based on the biorthogonality of the electric current j = µ−1 ∇ × B and the vector potential A and explicitly utilizes an expansion of the magnetic field 0 B into (analytical known) free decay modes. The method allows a very fast computation of eigenvalues and eigenfunctions for nearly arbitrary spatial distributions of α and/or ηT. Further we restrict our ex- amination to the axisymmetric field which is dominating for isotropic α and β. Higher azimuthal modes will be important in case of anisotropic α coefficient (Tilgner 2004), although in case of anisotropic ηT axisymmetric modes become again dominant (Elstner and Rudiger 2007). June 16, 2018 3:22 Geophysical and Astrophysical Fluid Dynamics gieseckeoscalpha2 4 (colour online) Central panel: growth rates for an α2 dynamo with α = α0 cos ϑ. Note the transition between different types of Figure 1. degeneration when going from α0 = 0 (free decay) to α0 ≫ αcrit. The small panels surrounding the central figure present the geometric structure of the eigenmodes for α0 = 0 (left hand side) and for α0 = 15 (right hand side). The colored contours denote the toroidal part and the arrows denote the poloidal part. Note the separation of toroidal and poloidal components in the free decay case (α0 = 0, left). 3 Results We start with a full sphere (embedded in vacuum) where α is created from turbulent motions in a convective layer subject to rotation. Then the EMF contains a term parallel to the mean magnetic field E ∝ (g·Ω)B where g points in the radial direction and Ω is parallel to the rotation axis (z-axis) so that (g·Ω) ∝ cos ϑ. Thus, the α-effect is maximum at the poles and vanishes at the equator: α = α0 cos ϑ. (9) This α model has been examined e.g. in Roberts (1972). Here we revive these results and additionally present eigenvalues and eigenfunctions of higher order modes. The resulting spectrum of the leading dy- namo eigenmodes in terms of the growth rates versus the amplitude α0 is shown in figure 1. All eigenvalues are real, i.e. all eigenfunctions are non-oscillatory modes. Due to the strict antisymmetry of α with respect to the equator the parity of the eigenmodes remains preserved for increasing α. That is, the α-term in the induction equation couples toroidal and poloidal modes in a way that the dynamo eigenmode remains antisymmetric (in the following we call this "dipolar-like") or symmetric (in the following we call this "quadrupolar-like") with respect to the equator. At α0 = 0 (free decay) an exact degeneration between "consecutive" modes is obtained (as indicated on the left side of figure 1). The degenerated eigenfunctions are always a pair of a purely toroidal and a purely poloidal mode which differ in multipolar June 16, 2018 3:22 Geophysical and Astrophysical Fluid Dynamics gieseckeoscalpha2 5 0 = 7.645) is close to the value for the quadrupolar mode (αcrit degree by ∆l = 1. The degeneration vanishes for α0 > 0 and around the dynamo threshold a slight predominance of the dipole mode occurs. The critical α for the onset of dynamo action of the dipole mode (αcrit 0 = 7.813). Both values agree with the results obtained by Roberts (1972) (αcrit 0 = 7.803 for the quadrupole mode). For α ≫ αcrit all eigenfunctions again approach a twofold degenerated state that consists of a pair of degenerated modes with similar geometric structure, but opposite equatorial symmetry (see right side in figure 1). The approximate degeneration between dipolar-like and quadrupolar-like modes is the result of a symmetry in the equations describing the α2-dynamo which is exact in case of perfect conducting boundary conditions. In that case dipolar eigenfunctions and quadrupolar eigenfunctions are adjoints of each other and thus have the same (conjugate) eigenspectrum (Proctor 1977). The degeneration is approximately retained in case of insulating boundary conditions so that the eigenvalues remain close to each other (Dobler and Radler 1998). 0 = 7.637 for the dipole mode and αcrit Radial dependence of α Next, we introduce a radial dependence for α taken from the model presented in Giesecke et al. (2005a). This model utilizes an idealized parameterization in the radial direction that is based on local simulations of rotating magnetoconvection assuming conditions roughly suitable for the geodynamo (weak stratification, fast rotation, strong magnetic field; Giesecke et al., 2005b). Qualitatively these simulations show that on the northern hemisphere α is negative in the lower part of the liquid outer core and positive in the upper part, whereby the zero occurs roughly in the middle of the convective instable layer. The idealized radial α-profile that incorporates these properties is given by α(r) = α0 cos ϑ sin(cid:18)2π (r − Rin) (Rout − Rin)(cid:19) , (10) where the outer radius of the spherical domain is fixed to Rout = 1 and Rin represents the radius of an inner core (see figure 2). A change in the radial profile was also considered in the first planetary mean field models (Steenbeck and Krause 1969). However, in that work α was related to the radial derivative of the turbulence intensity u2 rms and the authors exclusively looked for non-oscillating solutions. Figure 2. sponding to the actual size of the Earth's inner core. (colour online) Radial and latitudinal distribution of the α-effect as given by (10). Left: Rin = 0; Right: Rin = 0.35 corre- Here, we start assuming no inner core, i.e. Rin = 0. The spectrum for the leading axisymmetric modes is shown in figure 3. In contrast to the simple α-profile examined in the previous section, oscillatory solutions (indicated by dotted curves) are obtained for this new radial α profile. In particular, around the dynamo threshold the dominating eigenmodes are oscillatory dipolar-like (blue curve) and quadrupolar-like (red curve) solutions. The oscillating solutions appear/disappear at so called exceptional points (EP) where two stationary modes coalesce and continue as two oscillatory modes with conjugate complex eigenvalues, i.e. with the same growth rate and positive and negative frequency, respectively (the right panel of figure 3 June 16, 2018 3:22 Geophysical and Astrophysical Fluid Dynamics gieseckeoscalpha2 6 shows only the positive frequency branch). Regarding the appearance of oscillatory modes at α = 0 (dotted Figure 3. Left panel: growth rates versus α0. The leading eigenmodes are numbered in the order of their appearance at α0 = 0. Right panel: oscillation frequency versus α0 (only the positive frequencies for the three leading time dependent modes are shown). curves on the left panel of figure 3) it is necessary to stress that exactly at α = 0 (i.e. for free decay) these modes are stationary and degenerated, i.e. α = 0 is an exceptional point for these modes. The spectrum is far more complex than in models without any radial structure of the α-effect and which is apparent by the confusing structure of the eigenvalues with a variety of level crossings and/or exceptional points where oscillatory solutions constitute or vanish. However, for α ≫ αcrit predominantly stationary modes are observed. Decomposition in free decay modes The change of the temporal character of the dynamo eigenmodes from non-oscillatory to oscillatory behavior is attended by a change in the radial structure. This can be seen by means of a decomposition of a dynamo eigenmode in terms of the leading free decay modes (which also represent the basis utilized in the numerical scheme). The dominating contributions in dependence on the magnitude of the α-effect are shown in figure 4. For both symmetry classes the oscillatory regime Figure 4. Contribution of characteristic free decay modes to the first dipolar-like dynamo eigenmode (left panel, eigenmode 1 in figure 3) and the first quadrupolar-like dynamo eigenmode (right panel, eigenmode 2 & 3 in figure 3). is characterized by domination of modes with radial wavenumber n = 1 whereas stationary solutions are June 16, 2018 3:22 Geophysical and Astrophysical Fluid Dynamics gieseckeoscalpha2 characterized by a domination of (n = 0) modes1. Another feature shown in the decomposition of the dynamo eigenmode is the increment of the multipolar degree of the dominant contribution with increasing α. 7 Inner core The influence of an inner core with finite electrical conductivity had been examined by Hollerbach and Jones (1993, 1995) and Wicht (2002) with surprisingly opposite conclusions. Whereas the latter only found little influence of finite inner core conductivity on temporal behavior of the dipole field, Hollerbach and Jones (1993, 1995) concluded from simulations that the finite conductivity of an inner core has a stabilizing impact and reversals could only occur if the field in the fluid outer core exhibits a large and long lasting fluctuation that allows the field to reverse throughout the inner core as well. Figure 5(a) shows that the spectrum of the eigenmodes is only slightly changed when an inner core is considered with a radius Rin = 0.35 and a uniform diffusivity within inner and outer core (ηT = 1). The α-effect is prescribed by the distribution (10) with α = 0 for r < Rin. The results for the same size of the inner core, but with a reduced diffusivity for r < Rin, is shown in figures 5(b) and (c). The introduction of an inner core without changing the diffusivity distribution has remarkable little influence on the spectrum (figure 5a) whereas a reduced inner core diffusivity for sufficiently small α results in a split-up of the oscillatory modes into two stationary branches (figure 5b and c). However, around the onset of dynamo Figure 5. Growth rates versus α0 for a modified α-profile given by (10). An inner core is considered with an (earth-like) radius Rin = 0.35 with a magnetic diffusivity ηcore = 1, 0.5, 0.1 (from left to right). Solid black curves represent non-oscillatory solutions and grey/dotted curves show oscillatory solutions. action the typical pattern of the leading dynamo eigenmodes is hardly changed with increasing inner core conductivity. For example, the reduction from ηcore = 1 to ηcore = 0.1 leads only to a small decrease of the critical α-magnitude from αcrit ≈ 16.9 to αcrit ≈ 16.5. Furthermore, the temporal behavior of the leading modes is not changed within this regime, i.e. the leading modes remain oscillating around the onset of dynamo action independent of the inner core conductivity. In particular for α ≫ αcrit no changes can be observed in the behavior of the leading modes. The main impact of an enhanced inner core conductivity is manifested in the cancellation of the degeneration of the eigenfunction for α <∼ αcrit. Consequently, the spectrum becomes more complex than in case of no inner core or with uniform conductivity. Equatorial symmetry breaking A coupling between dipole and quadrupole is the basis of the low di- mensional reversal model of P´etr´elis and Fauve (2008), P´etr´elis et al. (2009) where the authors conclude that the reversal rate is constrained by breaking of the equatorial symmetry. With this motivation in the background, we investigate the influence of equatorial symmetry breaking of the α-effect. In the model an additional term is added proportional to cos(2ϑ) leading to α = α0 sin(cid:18) 2π (r − Rin) Rout − Rin (cid:19) (cos ϑ + b cos(2ϑ)) (11) 1n characterizes the degree of a spherical Bessel function of a particular free decay mode and determines the radial behavior. June 16, 2018 3:22 Geophysical and Astrophysical Fluid Dynamics gieseckeoscalpha2 8 where b ≪ 1 and again Rin = 0.35. The main impact of the parameter b is a coupling of dipolar-like and quadrupolar-like modes so that a distinction in terms of parity with respect to the equator is no longer possible. The resulting growth rates for various values of the parameter b are shown in figure 6 whereby α is restricted to values around the dynamo threshold. Without symmetry breaking two oscillatory eigenfunc- Figure 6. Extraction of growth rates vs. α0 for dynamo models with equatorial symmetry breaking demonstrating the impact of symmetry breaking on the degeneration of the eigenfunctions. Solid curves denote stationary solutions and dotted curves denote oscillatory solutions. b = 0.00, 0.01, 0.05, 0.10. The contour plots on the right hand side show the pattern of the (hemispherical) eigenmodes for case (b) at α0 = 17 (colour online). tions that correspond to an oscillatory dipolar-like mode and an oscillatory quadrupolar-like mode cross the dynamo threshold around α0 ≈ 16.9. The coupling of dipolar-like and quadrupolar-like modes results in a number of significant changes in the eigenvalues and geometric structure. Here we restrict the dis- cussion to the behavior around the dynamo threshold of the leading eigenmode. First of all, we observe a reduction of the critical α with increasing b which is attended by a breakup of the degeneration of the leading eigenmodes. However, the dominating eigenmodes retain their oscillatory property but due to the coupling of both symmetry classes a distinction into dipolar-like and quadrupolar-like eigenmodes is no longer possible. Compared to the undisturbed case, a considerable change in the field structure occurs for sufficiently strong α. Around the dynamo threshold (and above) the coupling of both classes results in the occurrence of hemispherical dynamos that oscillate in time. Since dipolar-like and quadrupolar-like modes contribute roughly the same amount to the new coupled mode (recall that the growth rates of both types are roughly equal in the case without any perturbation) the coupled dynamo mode is concentrated within one hemisphere. Since dipolar and quadrupolar modes have a very similar structure, their contribution cancel out in one hemisphere so that magnetic energy is concentrated in the remaining hemisphere. June 16, 2018 3:22 Geophysical and Astrophysical Fluid Dynamics gieseckeoscalpha2 9 4 Conclusions We have re-confirmed in a simple mean-field dynamo model that oscillatory α2-dynamos are possible when the radial profile of α exhibits a more complex structure. Our model is based on a particular radial profile of the α-effect that behaves ∝ sin(r). The choice of this profile is not arbitrary but is motivated by simulations of rotating magnetoconvection (Giesecke et al. 2005b) and quasi-linear computations performed by Soward (1974). On a first glance the occurrence of oscillatory solutions seems quite robust. Around the onset of dynamo action, growing oscillatory solutions dominate independently of core conductivity (and/or core size) or equatorial symmetry breaking. More important for a disappearance of oscillatory solutions might be deviations from the ideal sin-profile such as shifts of the zero crossing (Giesecke et al. 2005a) which has not been examined here. The eigenfunctions of the α2-model exhibit a couple of remarkable properties that characterize the geometric structure of the eigenfields. In the oscillatory regime the eigenfunctions are dominated by con- tributions with higher radial wavenumber whereas the steady solutions essentially are determined by eigenfunctions with sparse radial structure. Hence, it is suggestive to look for an indication of a similar contribution of higher radial modes during an reversal in three-dimensional MHD simulations of the geody- namo or in field reconstructions from paleomagnetic observations as e.g. executed by Leonhardt and Fabian (2007). For α-distributions with perfect equatorial (anti-)symmetry α ∝ cos ϑ the dipolar and quadrupolar modes remain separated since no interaction between these modes is possible. The eigenmodes with differ- ent equatorial symmetry exhibit an approximate degeneration which would be exact in case of perfectly conducting boundary conditions. The corresponding proximity of growth rates for dipole and quadrupole has a dramatic impact when a small perturbation is considered that breaks the equatorial symmetry. In the vicinity of the dynamo threshold the resulting coupling between dipolar- and quadrupolar-like modes leads to hemispherical dynamo action. Indeed, the spatial reconstruction of the last reversal (the Matuyama- Brunhes transition ∼ 780 kyrs ago) shows a growing contribution of the quadrupolar component during the actual reversal (Leonhardt and Fabian 2007) but there are no hints for hemispherical dynamo action taken place in the Earth's core. Nevertheless, hemispherical dynamo action might have been the reason for the non-uniform crust magnetization as a result of the ancient martian dynamo (Landeau and Aubert 2011). Our results differ from the achievement of Gallet and P´etr´elis (2009) who examined a kinematic α2 model using a strongly localized α-effect concentrated in two thin shells. In their model it is the coupling between dipole and quadrupole induced by equatorial symmetry breaking which leads to an oscillating eigenmode, whereas in our model the coupling between different radial modes (induced by the radial variation of α) is responsible for the oscillatory behavior. The simplicity of both models, however, prevents a robust conclusion which specific behavior indeed is realized in the geodynamo. Regarding the Earth's magnetic field, a reason for the suppression of the quadrupole might arise from anisotropies of the turbulent flow essentially caused by the fast rotation of the Earth. These anisotropies resulting from the fast rotation of the Earth which suppresses variations/fluctuations along the rotation axis should be reflected in more realistic models of the α-tensor as well as in the diffusivity tensor. Using more realistic anisotropic structure of the α-tensor (but isotropic η) results in domination of non-axisymmetric (i.e. m = 1) modes (Giesecke et al. 2005a). This domination can be circumvented by assuming anisotropic description for η as well (Tilgner 2004). Note that non-axisymmetric modes indeed seem to be important in the reconstruction of the field pattern during the last reversal (the Matuyama-Brunhes transition) where the most important field contribution is determined by a m = 2 contribution (Leonhardt and Fabian 2007). A substantial limitation of the presented results is the linear character of the kinematic models, in particular regarding the proximity of the eigenvalues of leading dipolar and quadrupolar eigenmodes. Thus a robust prediction on the behavior in a saturated state is difficult and more elaborated mean field models will require the consideration of backreaction of the magnetic field by virtue of α-quenching. Qualitatively, we expect a similar behavior as it has been described by Stefani and Gerbeth (2005) in a one-dimensional non-linear model of a reversing α2-dynamo. Assuming that the magnetic quenching of the α-effect determines the instantaneous growth rate the authors showed that the corresponding eigenmode is June 16, 2018 3:22 Geophysical and Astrophysical Fluid Dynamics gieseckeoscalpha2 10 REFERENCES inevitably driven towards the oscillatory branch. This shift is attended by self accelerating field decay and the emerging local maximum of the growth rate in connection with noise provide the essential preconditions to switch from a stationary branch to an oscillatory branch and vice versa. Acknowledgments Financial support from Deutsche Forschungsgemeinschaft (DFG) in frame of the Collaborative Research Center (SFB) 609 is gratefully acknowledged. REFERENCES Berhanu et al., M., Magnetic field reversals in an experimental turbulent dynamo. Europhys. Lett. 2007, 77, 59001 -- +. Carbone, V., Sorriso-Valvo, L., Vecchio, A., Lepreti, F., Veltri, P., Harabaglia, P. and Guerra, I., Clustering of polarity reversals of the geomagnetic field. Phys. Rev. Lett. 2006, 96, 128501 -- +. Christensen, U.R., Geodynamo models: Tools for understanding properties of Earth's magnetic field. Phys. Earth Planet. Int. 2011, 187, 157 -- 169. Dobler, W. and Radler, K.H., An integral equation approach to kinematic dynamo models. Geophys. Astrophys. Fluid Dyn. 1998, 89, 45 -- 74. Dudley, M.L. and James, R.W., Time-dependent kinematic dynamos with stationary flows. Phil. Trans. R. Soc. Lond. A 1989, 425, 407 -- 429. Elstner, D. and Rudiger, G., How can α2-dynamos generate axisymmetric magnetic fields?. Astron. Nachr. 2007, 328, 1130 -- 1132. Fischer, M., Gerbeth, G., Giesecke, A. and Stefani, F., Inferring basic parameters of the geodynamo from sequences of polarity reversals. Inverse Probl. 2009, 25, 065011. Gallet, B. and P´etr´elis, F., From reversing to hemispherical dynamos. Phys. Rev. E 2009, 80, 035302 -- +. Giesecke, A., Rudiger, G. and Elstner, D., Oscillating α2-dynamos and the reversal phenomenon of the global geodynamo. Astron. Nachr. 2005a, 326, 693 -- 700. Giesecke, A., Ziegler, U. and Rudiger, G., Geodynamo α-effect derived from box simulations of rotating magnetoconvection. Phys. Earth Planet. Int. 2005b, 152, 90 -- 102. Gunther, U., Langer, H. and Tretter, C., On the spectrum of the magnetohydrodynamic mean-field α2- dynamo operator. SIAM J. Math. Anal. 2010, 42, 1413 -- 1447. Harland, W., Cox, A., Liewellyn, P., Pickton, C., Smith, A. and Walters, R., A geological time scale, 1982 (Cambridge University Press). Hollerbach, R. and Jones, C.A., Influence of the Earth's inner core on geomagnetic fluctuations and reversals. Nature 1993, 365, 541 -- 543. Hollerbach, R. and Jones, C.A., On the magnetically stabilizing role of the Earth's inner core. Phys. Earth Planet. Int. 1995, 87, 171 -- 181. Hoyng, P., Schmitt, D. and Ossendrijver, M.A.J.H., A theoretical analysis of the observed variability of the geomagnetic dipole field. Phys. Earth Planet. Int. 2002, 130, 143 -- 157. Krause, F. and Radler, K.H., Mean-field magnetohydrodynamics and dynamo theory, 1980 (Oxford: Perg- amon Press). Landeau, M. and Aubert, J., Equatorially asymmetric convection inducing a hemispherical magnetic field in rotating spheres and implications for the past martian dynamo. Phys. Earth Planet. Int. 2011, 185, 61 -- 73. Leonhardt, R. and Fabian, K., Paleomagnetic reconstruction of the global geomagnetic field evolution during the Matuyama/Brunhes transition: Iterative Bayesian inversion and independent verification. Earth Planet. Sci. Lett. 2007, 253, 172 -- 195. McFadden, P.L. and Merrill, R.T., Geodynamo energy source constraints from palaeomagnetic data. Phys. Earth Planet. Int. 1986, 43, 22 -- 33. June 16, 2018 3:22 Geophysical and Astrophysical Fluid Dynamics gieseckeoscalpha2 REFERENCES 11 Mitra, D., Tavakol, R., Kapyla, P.J. and Brandenburg, A., Oscillatory Migrating Magnetic Fields in Helical Turbulence in Spherical Domains. Astrophys. J. Lett. 2010, 719, L1 -- L4. Olson, P. and Christensen, U.R., Dipole moment scaling for convection-driven planetary dynamos. Earth Planet. Sci. Lett. 2006, 250, 561 -- 571. P´etr´elis, F. and Fauve, S., Chaotic dynamics of the magnetic field generated by dynamo action in a turbulent flow. J. Phys. Condens. Matter 2008, 20, 4203. P´etr´elis, F., Fauve, S., Dormy, E. and Valet, J.P., Simple mechanism for reversals of Earth's magnetic field. Phys. Rev. Lett. 2009, 102, 144503 -- +. Proctor, M.R.E., On the eigenvalues of kinematic alpha-effect dynamos. Astron. Nachr. 1977, 298, 19 -- 25. Radler, K.H. and Brauer, H.J., On the oscillatory behaviour of kinematic mean-field dynamos. Astron. Nachr. 1987, 308, 101 -- 109. Roberts, P.H., Kinematic dynamo models. Phil. Trans. R. Soc. Lond. A 1972, 272, 663 -- 698. Roberts, P.H. and Stix, M., AC-effect dynamos, by the Bullard-Gellman formalism. Astron. Astrophys. 1972, 18, 453. Rudiger, G., Elstner, D. and Ossendrijver, M., Do spherical α2-dynamos oscillate?. Astron. Astrophys. 2003, 406, 15 -- 21. Schrinner, M., Petitdemange, L. and Dormy, E., Oscillatory dynamos and their induction mechanisms. Astron. Astrophys. 2011, 530, A140. Schrinner, M., Schmitt, D., Jiang, J. and Hoyng, P., An efficient method for computing the eigenfunctions of the dynamo equation. Astron. Astrophys. 2010, 519, A80+. Sorriso-Valvo, L., Stefani, F., Carbone, V., Nigro, G., Lepreti, F., Vecchio, A. and Veltri, P., A statistical analysis of polarity reversals of the geomagnetic field. Phys. Earth Planet. Int. 2007, 164, 197 -- 207. Soward, A.M., A convection-driven dynamo: I. The weak field case. Phil. Trans. R. Soc. Lond. A 1974, 275, 611 -- 646. Steenbeck, M. and Krause, F., On the Dynamo Theory of Stellar and Planetary Magnetic Fields. II. DC Dynamos of Planetary Type. Astronomische Nachrichten 1969, 291, 271 -- 286. Stefani, F. and Gerbeth, G., Oscillatory mean-field dynamos with a spherically symmetric, isotropic helical turbulence parameter α. Phys. Rev. E 2003, 67, 027302 -- +. Stefani, F. and Gerbeth, G., Asymmetric polarity reversals, bimodal field distribution, and coherence resonance in a spherically symmetric mean-field dynamo model. Phys. Rev. Lett. 2005, 94, 184506 -- +. Stefani, F., Xu, M., Sorriso-Valvo, L., Gerbeth, G. and Gunther, U., Oscillation or rotation: a comparison of two simple reversal models. Geophys. Astrophys. Fluid Dyn. 2007, 101, 227 -- 248. Tilgner, A., Small scale kinematic dynamos: beyond the α-effect. Geophys. Astrophys. Fluid Dyn. 2004, 98, 225 -- 234. Wicht, J., Inner-core conductivity in numerical dynamo simulations. Phys. Earth Planet. Int. 2002, 132, 281 -- 302.
1708.04600
4
1708
2019-11-28T12:16:21
Anomalous light curves of young tilted exorings
[ "astro-ph.EP" ]
Despite the success of discovering transiting exoplanets, several recently observed objects (e.g. KIC-8462852, J1407 and PDS-110) exhibit unconventional transit signals, whose appropriate interpretation in terms of a spherical single body has been challenging, if not impossible. In the aforementioned examples the presence of a ring-like structure has been proposed for explaining the unusual data. Thus, in this paper we delve into the dynamics of a tilted exoring disturbed by a third close companion, and the role that the Lidov-Kozai mechanism may have to explain irregular and anomalous transit signals of ringed planets, as well as the ring's early stages. To that end, we performed numerical simulations and semi-analytical calculations to assess the ring's dynamical and morphological properties, and their related transit observables. We found that tilted ringed structures undergo short-term changes in shape and orientation that are manifested as strong variations of transit depth and contact times, even between consecutive eclipses. Any detected anomaly in transit characteristics may lead to a miscalculation of the system's properties (planetary radius, semi-major axis, stellar density and others). Moreover, oscillating ring-like structures may account for the strangeness of some light-curve features in already known and future discovered exoplanets.
astro-ph.EP
astro-ph
MNRAS 000, 1 -- 5 (2017) Preprint 2 December 2019 Compiled using MNRAS LATEX style file v3.0 Anomalous light curves of young tilted exorings Mario Sucerquia(cid:63), J. A. Alvarado-Montes, Vanesa Ram´ırez and Jorge I. Zuluaga Solar, Earth and Planetary Physics Group (SEAP) Instituto de F´ısica - FCEN, Universidad de Antioquia, Colombia Calle 70 No. 52-21, Medell´ın, Colombia Accepted 2017 September 22. Received 2017 September 21; in original form 2017 July 11 ABSTRACT Despite the success of discovering transiting exoplanets, several recently observed ob- jects (e.g. KIC-8462852, J1407 and PDS-110) exhibit unconventional transit signals, whose appropriate interpretation in terms of a spherical single body has been chal- lenging, if not impossible. In the aforementioned examples the presence of a ring-like structure has been proposed for explaining the unusual data. Thus, in this paper we delve into the dynamics of a tilted exoring disturbed by a third close companion, and the role that the Lidov-Kozai mechanism may have to explain irregular and anoma- lous transit signals of ringed planets, as well as the ring's early stages. To that end, we performed numerical simulations and semi-analytical calculations to assess the ring's dynamical and morphological properties, and their related transit observables. We found that tilted ringed structures undergo short-term changes in shape and orienta- tion that are manifested as strong variations of transit depth and contact times, even between consecutive eclipses. Any detected anomaly in transit characteristics may lead to a miscalculation of the system's properties (planetary radius, semi-major axis, stellar density and others). Moreover, oscillating ring-like structures may account for the strangeness of some light-curve features in already known and future discovered exoplanets. Key words: Techniques: photometric -- Planets and satellites: dynamical evolution and stability and rings -- Stars: individual: KIC-8462852, J1407, PDS-110. 1 INTRODUCTION Transit light-curve signals encode valuable information about the star, the exoplanet and its surroundings. An ap- propriate interpretation can provide us useful information to understand and characterize the distribution of plane- tary and stellar properties. Contrary to the straightforward explanation of simple lightcurves (the most frequent ones), in recent years we have discovered complex transit signals that challenge the well-established transit theories. How- ever, given the complexity of these signals, new findings may emerge to reveal planetary processes that are absent in the Solar System. Mamajek et al. (2012) found a single transit which has been interpreted as a colossal ring around a sub-stellar ob- ject orbiting the star J1407. Similarly, Osborn et al. (2017), have proposed a ring-like structure as a feasible explanation for the two high-depth eclipses observed for PDS-110 system. Recently, Boyajian et al. (2016) reported an odd brightness stellar variability of KIC 8462852, including dramatic and (cid:63) E-mail: [email protected] © 2017 The Authors irregular dims without any analogous case in exoplanet ob- servations. The latest object (that has been dubbed for short "Tabby's star") is the most intriguing case. Feasible expla- nations reported in literature, argue the possible detection of a younger star with coalescing material orbiting around it (Lisse, Sitko & Marengo 2015), a cloud of disintegrating comets (Boyajian et al. 2016; Bodman & Quillen 2016) and a planetary debris field (Lisse et al. 2015; Boyajian et al. 2016), as well as the combined effect of a huge planetary ring and co-orbital trojan asteroids (Ballesteros 2017). Be- yond these well-known cases, there are lots of anomalous detections in photometric surveys (Burke et al. 2014, and references therein), that are awaiting for a proper confirma- tion and explanation. The discovery of rings around exoplanets (hereafter ex- orings) is one of the expected outcomes of the forthcoming exoplanetary research. Although no exoring has been discov- ered so far (at least around a Jupiter-like planet), theoreti- cal investigations (Schlichting & Chang 2011; Hedman 2015) have showed that hypothetical 'warm rings', i.e. rings made of refractory particles instead of volatile ones, could exist around the most abundant Jupiter-sized planets, namely 2 Sucerquia et al. close-in giants. The detection and characterization of a pop- ulation of exorings around extrasolar giants is awaiting the development of novel and fast analysis techniques (Zuluaga et al. 2015), and more precise photometry provided by future missions such as JWST and PLATO. Rings are complex structures and their transit signals will not be as simple as those of single planets. The most important dynamical mechanism affecting tilted rings is the so-called Lidov-Kozai mechanism (hereafter LKM) (Lidov 1962; Kozai 1962). LKM has been successfully applied to explain unconventional planetary system architectures, such as planets in highly eccentric orbits (Takeda & Rasio 2005), formation of hot Jupiters (Lithwick & Naoz 2011), among others. We hypothesize that LKM may be responsible of significant changes in the spatial configuration of exorings around close-in giant planets. These changes may occur in time-scales short enough to be detected in present and fu- ture lightcurves of ringed planets. Therefore, we propose that several anomalous transit signals could be explained as a consequence of the ring's complex dynamics. This paper is organized as follows: Section 2 describes the behavior of rings under the LKM from an analytic ap- proach. In Section 3 we show the outcomes of numerical simulations, performed to asses the evolution of a tilted ring under the perturbing forces of the host star. Section 4 dis- cusses the effect of ring evolution on the lightcurves. Finally, we devote Section 5 to summarize and present the most re- markable conclusions of this work. 2 THE LIDOV-KOZAI MECHANISM Particles in an initially tilted ring around a close-in planet are strongly affected by perturbations of the host star. When the initial inclination of the ring particles' orbits, as mea- sured from the planet's orbit, is large enough (i.e. io >∼ 39.2◦, see a depiction of the system in Figure 1), a cyclic exchange of angular momentum between the particles and the dis- turber arises. As a result, both the particle's orbital incli- nation and eccentricity undergo strong periodic oscillations. This is the Lidov-Kozai Mechanism. The maximum eccen- tricity reached by a particle emax depends only on io (see e.g. Ford, Kozinsky & Rasio 2000 and references therein): emax (cid:39) (cid:113) 1 − (5/3) cos2 io. (1) When particle orbits become more eccentric, the shape and ring's size ∆ are also affected. Moreover, if io is large enough, particles in the inner part of the ring may collide with the planet and the ring is eroded over time. The total depleting of ring's particles is avoided only if io is lower than an incli- nation threshold ith, which depends on planetary radius Rp and particle initial semi-major axis ar (see Figure 2). The value of ith is computed from emax, by imposing that the pe- riapsis distance at maximum eccentricity qr = ar (1 − emax) be greater than Rp, or equivalently: 2 ar Rp − 1 cos2 ith (cid:39) 3 5 , ar (2) here ar/Rp is bounded between fi and fe, the inner and outer relative-radius of the ring, respectively. The initial value of fe is assumed smaller than or equal to the Roche radius RRoche. However, when the outermost particles reach an ec- centricity emax, their apoapsis Q = fe (1+emax) become larger (cid:18) Rp (cid:19)2(cid:18) (cid:19) than that limit. Particles reaching those regions may coa- lesce to eventually drive the formation of moonlets (Crida & Charnoz 2012). This process could also contribute to the loss of ring's particles. The above mechanisms limit the range of initial sizes and inclinations that rings should have to survive LKM with- out loosing a significant fraction of their mass. In Figure 2 we illustrate how the limits of those properties are deter- mined and used for the numerical experiments described in the following section. 3 NUMERICAL SIMULATIONS To test our hypothesis we performed a set of numerical ex- periments using HNBODY (Rauch & Hamilton 2012), where a collection of 104 massless and collisionless particles, ini- tially distributed across a tilted ring, evolves under the gravitational forces of both a close-in Neptune-like planet (Mp = 10−4 M(cid:12), Rp = 2.5 × 104 km) and a Solar-mass star. The planet is located in a circular orbit with ap = 0.1 au. The ring has an initial inclination io = 50◦ and inner ( fi) and outer ( fe) relative-radii of 2.11 and 2.81, respec- tively. The latter value corresponds to the Roche radius as computed for loose-bound bodies made of refractory ma- terial, i.e. densities about 2, 000 kg/m3. If we also assume a ring's total mass about 1019 kg (that of Saturn's rings), each particle will have a mass of about 1015 kg and a size of ∼ 5 km. Particles' initial positions are randomly generated us- ing a Poisson Disc Sampling algorithm to avoid oversampled regions (see e.g. Cook 1986 and references therein), and the system was allowed to evolve during 300 planetary orbital periods. Since our goal here is to estimate the time-scales and other broader effects of a massive oscillating disk in the light curve of a transiting planet, we have neglected the effect that self-gravitation, collisional damping and viscosity have on the ring's evolution (see eg. Hyodo & Charnoz 2017 and references there in). A more accurate approach should take into account these factors. Also, self-gravitation and colli- sional damping will act on reducing the dispersion of par- ticles and circularizing their orbits. Viscosity (which arises from translational, collisional and gravitational interchange of angular momentum among ring particles) will contribute to spread out the disk, increasing the mass-loss rate, and consequently reducing the ring's optical depth and damping its oscillations (Daisaka et al. 2001). As a side note, it is worth to notice that including the gravitational interaction among ring particles will also drive the formation of ringlets and wakes. However, since the dy- namical time-scales of these processes (several years) is of the order of ring oscillation damping, their role on the light- curve evolution could be negligible. On the other hand, the interaction among ring particles could contribute to self- organize the oscillating ring. This could actually enhance the ring's signal, but also modify the oscillation's time-scales predicted here. A graphical illustration of the results of our numeri- cal experiments are presented in Figure 3. Spatial config- uration of the resulting structure as viewed from different projections at the time of planetary transits are depicted in 25 snapshots (250 planetary periods). As it is evident in the face-on projection (left panel), the ring's width changes significantly whilst the whole structure oscillates. From a MNRAS 000, 1 -- 5 (2017) Anomalous light curves of young exorings 3 Figure 1. Schematic representation of a ringed planet when passing in front of its host star. The initial inclination of the ring plane is io . The unitary vector n points in the direction of the average angular momentum of the ring particles. α, β and γ are the angles between n and the coordinate axis x, y and z, respectively. The x-axis points in the direction of the observer. Figure 2. Left panel: minimum periapsis q and maximum apoapsis Q reached by particles in the inner (in subscript) and outer edge (out subscript) of rings with different initial inclinations io, at emax. Black and red vertical dotted lines represent the initial inclination io used in our simulations and the threshold inclination ith (see Equation 2) respectively. Right panel: original (blue strip) and estimated ring width (area with dots) for io = 40◦, 50◦ at emax. cursory comparison between the first and last snapshot, we realize that ∆ 'expands' by a factor of ∼3, which is notice- able in the left panel (lower plot) of Figure 4, that gives a better insight about the broadness evolution. To follow the changes in the ring's orientation, we computed the ringlets' total angular momentum. This quantity gives us the direc- tion of the plane vector n (see Figure 1). The orientation of the ring (plotted in the upper panel of Figure 4) is calcu- lated for the entire lifespan of the system, by means of the director cosines or director angles α, β and γ also depicted in the left panel (upper plot) of Figure 1. 4 EFFECT IN LIGHTCURVES Numerical simulations suggest that a tilted ring subject to LKM, evolves towards a toroidal-like structure. Computing the transit signal of such a complex body is not trivial. Fur- thermore, we have noticed in our simulations that in spite of the three dimensional dispersion, the vertical density of par- ticles is strongly peaked around the instantaneous Laplace plane of the ring. As a first approximation we will model the transit signal of the oscillating rings as that of a unique disc- like structure, with internal and external radii fi, fe, width ∆ and projected inclination iR = 90◦ − α (see the evolution of these parameters in the left panel of Figure 4). In or- der to take into account the variation in vertical dispersion, we computed the expected transit depth assuming different values of the effective opacity τ. To estimate the transit's observables, namely the ob- served minimum flux Fobs, the transit depth δ = ARp/A(cid:63) (with ARp and A(cid:63) are the effective projected ring-planet and MNRAS 000, 1 -- 5 (2017) 4 Sucerquia et al. Figure 3. Snapshots of the evolution of a tilted ring for 250 transits. The ring's face-on, edge-on and on-sky projections are shown in the left, middle and right panel, respectively. The sizes are on scale. stellar area), and the equivalent planetary radius Rp,obs (cid:39) √ δ, we use Eqs. (1), (2) and (3) in Zuluaga et al. (2015). The results are presented in right panel of Figure 4. As expected, significant oscillations arise both in Fobs and Rp,obs, especially in the case of dense rings (higher values of τ). In that case the transit of the oscillating rings is manifested by a maximum stellar dimming of ∼ 5 per cent, correspond- ing to a planet's observed radius ∼ 3.5 times larger than a planet without rings. It is worth to notice that the effective planetary radius estimated from the lightcurve, is consistent with those values predicted in the analytical results of Fig- ure 2, giving some confidence on the adopted approximation to compute the transit's observables. The transit signal is mainly determined by the evolu- tion of the projected inclination α. However, variations in the spatial extent and width of the ring create additional dips and peaks in the lightcurve. It is important to stress that the curves in the rightmost column of Figure 4 only represent the flux and effective radius at the eclipse's cen- tre, and not actual lightcurves. Moreover, the exact value of those maxima depends on the way that the optical depth changes in an oscillating ring. Transit contact times also depend on the evolution of ∆ and α (which is noticeable for different ring's projections shown in Figure 3). In a real case these variations give rise to uncertainties in the planetary orbital period, as well as in the stellar density Zuluaga et al. (2015). 5 SUMMARY AND DISCUSSION The recent discovery of anomalous stellar lightcurves, have inspired the search for astrophysical systems and mecha- nisms which can drive non-trivial transit signals. In some cases it has become common to appeal to transient phenom- ena without any parallel in the Solar System. In this work we studied a dynamical process, namely the Lidov-Kozai mech- anism (LKM), capable of producing a quasi-periodic alter- ation of the transit's observable features of a young ringed close-in planet. For a given set of ring's initial properties (io, fi and fe) LKM determines the evolution and final fate of a ring through three mechanisms: 1) depletion of material of the ring through collision of particles with the planet at their periapsides, 2) drifting of ring's particles beyond the Roche limit, potentially driving the coalescence of material and the formation of moonlets, and 3) alteration of the spatial dis- tribution and thereby ring's geometrical parameters. Any of these mechanisms will be quite notorious in the correspond- ing planetary lightcurves. LKM and its light-curve signatures are transient in na- ture. The light-curve depth substantially depends on the ring projected inclination α, which as shown in the left panel (up- per plot) of Figure 4 tends to damp (driven by the damping of the physical ring inclination β). In our case the time re- quired for the amplitude of β to decay to about 40 per cent (t40) is only ∼ 9 yr. Therefore, the probability of observing a system undergoing this process is very small. However, as LKM oscillation period strongly depends on the distance to the perturbing object (e.g. the host star), the time-scales for outermost planets could be relatively large. For instance, if we assume the same planetary mass but use distances of 0.2, 0.3 and 0.5 au, t40 results in ∼ 20, 35 and 85 yr, respectively, increasing the chances to detect the predicted anomalies. LKM could also explain the absence of any ring signa- ture among the already close-in planets surveyed by Heising, Marcy & Schlichting (2015) and Aizawa et al. (2017). De- pending on the ring's initial mass, LKM may entirely oblit- erate an initial light ring. On the other hand, if rings are massive enough, the induced eccentric orbits enhances the collisional rate of particles, which may contribute to circu- larize and make more coplanar the particles' orbits (see dis- cussion in section 5.4 by Hyodo & Charnoz 2017). The latter mechanism, coupled with the mass-loss processes described before, may lead to final low-inclined and less-massive rings as those observed in the Solar System. The results presented in this work may also shed light on the possible fate (and observational signatures) of irreg- ular moons wiped out by tides (i.e. within the Roche limit). As shown here, the observation of anomalous lightcurves of close-in massive planets could be an evidence of these kinds of processes. Therefore, including the effect of LKM when preparing present and future searches for exomoons and ex- orings around close-in transiting planets, could be an inter- MNRAS 000, 1 -- 5 (2017) Anomalous light curves of young exorings 5 Figure 4. Left panel: upper plot, evolution of the director angles (α, β and γ) of the total angular momentum of ring particles, n; lower plot shows the evolution of the ring effective extent, fi (yellow) and fe (green), and width (black solid line). Right panel: upper plot, effective planet size as obtained from eclipse maximum depth, and minimum stellar flux (lower plot) as computed assuming different effective optical depths τ. esting task to pursue, if not mandatory. To notice here, our results are restricted to close-in planets. However, LKM may also operate if the external perturber, instead of the host star, is a massive moon or even a close planetary compan- ion. In those cases the anomalous transit signals described here could also be observed in ringed planets located at con- siderably larger distances from their stars. Our simulations assume a dense ring without any sub- structure (ringlets, wakes or gaps). For instance, gaps could produce remarkable signatures in the lightcurve (as observed for the cases of J1407b and PDS-110), but LKM oscilla- tions tends to erase any trace of them due to the signifi- cant induced changes in the orbits' eccentricity. That said, it could be interesting to study the effect that LKM has on the evolution of any ring substructure, but they have much larger formation time-scales than the processes considered here (Daisaka et al. 2001). The gaps' formation could require the presence of moons that accrete or disperse the mate- rial in time-scales of many orbital periods. The ringlets' for- mation will require several diffusion time-scales (Tremaine 2003). In this sense, our results are restricted to young discs where some of these effects are just happening or about to arise. Nevertheless, our outcomes (Figure 3) exhibit some void and overdensed regions, which might be confounded in lightcurves with the aforementioned substructures, although these features do not remain over time. Finally, in the light of our results, it could also be possible to speculate about the nature of the hypothetical object orbiting KIC-8462852. Provided enough information one may fit the observable signatures to a large oscillating disc made of circumplanetary debris subject to LKM. We speculate that if consecutive observations evidence some sig- natures of damped oscillations in the observed transit depth, we could be witnessing for the first time the disruption of a moon and the birth of a new ringed exoplanet. However, the confirmation or disprove of this hypothesis is well beyond the scope of this paper. MNRAS 000, 1 -- 5 (2017) ACKNOWLEDGEMENTS M.S. is supported by Doctoral Program of Colciencias and the CODI/UdeA. J.I.Z. is supported by Vicerrector´ıa de Do- cencia U de A. To Nadia Silva for the proofreading. REFERENCES Aizawa M., Uehara S., Masuda K., Kawahara H., Suto Y., 2017, AJ, 153, 193 Ballesteros F. J. A.-M. P. F.-S. A. M. V. J., 2017, Monthly Notices of the Royal Astronomical Society: Letters, p. slx105 Bodman E. H. L., Quillen A., 2016, ApJ, 819, L34 Boyajian T. S., et al., 2016, MNRAS, 457, 3988 Burke C. J., et al., 2014, ApJS, 210, 19 Cook R. L., 1986, ACM Transactions on Graphics (TOG), 5, 51 Crida A., Charnoz S., 2012, Science, 338, 1196 Daisaka H., Tanaka H., Ida S., 2001, Icarus, 154, 296 Ford E. B., Kozinsky B., Rasio F. A., 2000, ApJ, 535, 385 Hedman M. M., 2015, ApJ, 801, L33 Heising M. Z., Marcy G. W., Schlichting H. E., 2015, ApJ, 814, 81 Hyodo R., Charnoz S., 2017, The Astronomical Journal, 154, 8pp Kozai Y., 1962, AJ, 67, 591 Lidov M. L., 1962, Planet. Space Sci., 9, 719 Lisse C. M., Sitko M. L., Marengo M., 2015, ApJ, 815, L27 Lithwick Y., Naoz S., 2011, ApJ, 742, 94 Mamajek E. E., Quillen A. C., Pecaut M. J., Moolekamp F., Scott E. L., Kenworthy M. A., Collier Cameron A., Parley N. R., 2012, AJ, 143, 72 Osborn H. P., et al., 2017, Monthly Notices of the Royal Astro- nomical Society, 471, 740 Rauch K. P., Hamilton D. P., 2012, ASCL (ascl:1201.010) Schlichting H. E., Chang P., 2011, ApJ, 734, 117 Takeda G., Rasio F. A., 2005, ApJ, 627, 1001 Tremaine S., 2003, The Astronomical Journal, 125, 894 Zuluaga J. I., Kipping D. M., Sucerquia M., Alvarado J. A., 2015, ApJ, 803, L14 This paper has been typeset from a TEX/LATEX file prepared by the author.
1006.4443
1
1006
2010-06-23T08:21:20
Homogeneous studies of transiting extrasolar planets. III. Additional planets and stellar models
[ "astro-ph.EP", "astro-ph.SR" ]
I derive the physical properties of 30 transiting extrasolar planetary systems using a homogeneous analysis of published data. The light curves are modelled with the JKTEBOP code, with attention paid to limb darkening and eccentricity. The light from some systems is contaminated by faint nearby stars, which if ignored will systematically bias the results. I show that this must be accounted for using external measurements of the amount of contaminating light. A contamination of 5% is enough to make the measurement of a planetary radius 2% too low. The physical properties of the 30 transiting systems are obtained by interpolating in stellar model predictions to find the best match to their measured quantities. The error budgets are used to compile a list of systems which would benefit from additional observations. The systematic errors arising from the inclusion of stellar models are assessed by using five different theoretical models. This model dependence sets a lower limit on the accuracy of measurements of the system properties, and at worst is 1% for the stellar mass. The correlations of planetary surface gravity and mass with orbital period have significance levels of only 3.1 sigma and 2.3 sigma. The division of planets into two classes based on Safronov number is increasingly blurred. Most of the objects studied here would benefit from more photometry and/or spectroscopy, as well as a better understanding of low-mass stars.
astro-ph.EP
astro-ph
Mon. Not. R. Astron. Soc. 000, 000–000 (0000) Printed 23 April 2018 (MN LATEX style file v2.2) Homogeneous studies of transiting extrasolar planets. III. Additional planets and stellar models John Southworth⋆ Astrophysics Group, Keele University, Staffordshire, ST5 5BG, UK 23 April 2018 ABSTRACT I derive the physical properties of thirty transiting extrasolar planetary systems using a homo- geneous analysis of published data. The light curves are modelled with the JKTEBOP code, with special attention paid to the treatment of limb darkening, orbital eccentricity, and error analysis. The light from some systems is contaminated by faint nearby stars, which if ignored will systematically bias the results. I show that it is not realistically possible to account for this using only transit light curves: light curve solutions must be constrained by measure- ments of the amount of contaminating light. A contamination of 5% is enough to make the measurement of a planetary radius 2% too low. The physical properties of the thirty transiting systems are obtained by interpolating in tabulated predictions from theoretical stellar models to find the best match to the light curve parameters and the measured stellar velocity amplitude, temperature and metal abundance. Statistical errors are propagated by a perturbation analysis which constructs complete error budgets for each output parameter. These error budgets are used to compile a list of systems which would benefit from additional photometric or spectroscopic measurements. The systematic errors arising from the inclusion of stellar models are assessed by using five independent sets of theoretical predictions for low-mass stars. This model dependence sets a lower limit on the accuracy of measurements of the physical properties of the systems, ranging from 1% for the stellar mass to 0.6% for the mass of the planet and 0.3% for other quantities. The stellar density and the planetary surface gravity and equilibrium temperature are not affected by this model dependence. An external test on these systematic errors is performed by comparing the two discovery papers of the WASP-11 / HAT-P-10 system: these two studies differ in their assessment of the ratio of the radii of the components and the effective temperature of the star. I find that the correlations of planetary surface gravity and mass with orbital period have significance levels of only 3.1σ and 2.3σ, respectively. The significance of the latter has not increased with the addition of new data since Paper II. The division of planets into two classes based on Safronov number is increasingly blurred. Most of the objects studied here would benefit from improved photometric and spectroscopic observations, as well as improvements in our understanding of low-mass stars and their effective temperature scale. Key words: stars: planetary systems - stars: binaries: eclipsing - stars: binaries: spectro- scopic - stars: fundamental parameters 1 INTRODUCTION The study of extrasolar planets is scientifically and culturally im- portant, and after a late start (Mayor & Queloz 1995) the number of known planets is escalating rapidly1. Transiting planets are the crown jewels of this population as, with the exception of our own Solar system, they are the only planets whose masses and radii are directly measurable. In addition to this, it is possible to put con- ⋆ E-mail: [email protected] 1 see the Extrasolar Planets Encyclopaedia, http://exoplanet.eu/ c(cid:13) 0000 RAS straints on the properties of their atmospheres, in which much in- teresting physics occurs, through measurements of the depths of transits and occultations at different wavelengths. Whilst transiting extrasolar planets (TEPs) offer unique sci- entific possibilities, their study involves several complications. The most significant is that it is not in general possible to measure the mass and radius of a planet through basic observations alone. Ad- ditional constraints are needed, and are usually provided by forcing 2 John Southworth the properties of the host stars to match theoretical expectations2. This introduces not only a model dependence (i.e. systematic er- ror), but also the possibility of inconsistent results if different the- oretical predictions are used for some TEPs. Systematic errors can blur any distinctions between planets, making it hard to pick out discrete groups of TEPs from the varied general population. This systematic error cannot be abolished, but it can at least be standardised. In this series of papers I am analysing the known tran- siting systems using rigorously homogeneous methods, with the aim of removing the systematic differences in measurements of the physical properties of TEPs. The resulting physical properties are therefore statistically compatible, and any structure in distributions of parameters is maximised. In Paper I (Southworth 2008) I analysed the light curves of the fourteen transiting systems for which high-precision photome- try was then available, paying careful attention to the role of limb darkening and to the estimation of comprehensive errorbars. Pa- per II (Southworth 2009) used these results plus the predictions of three different theoretical stellar models to measure the physical properties of the fourteen TEPs. In this work I broaden the analysis to thirty TEPs and five sets of theoretical stellar models, resulting in improved statistics and better systematic error estimates. There are a few homogeneous analyses of transiting systems available in the literature. A good analysis of 23 systems was pre- sented by Torres et al. (2008, hereafter TWH08), but these authors tried only two different theoretical model sets and did not assign systematic errors to their results. Analogously, such work would also benefit from homogeneous analysis of the spectra of the host stars in order to put their effective temperature and chemical abun- dance measurements on a consistent scale. Steps towards this goal were pioneered by Valenti & Fischer (2005) and are being contin- ued by Ammler-von Eiff et al. (2009) and Ghezzi et al. (2010), but a homogeneous study of the host stars of all known TEPs is not currently available. In Sect. 2 I present the methods used to analyse the light curves of the thirty TEPs included in this work. Sect. 3 discusses the five theoretical stellar model sets and their application to determining the physical properties of the TEPs. Sect. 4 presents the new results for these objects, Sect. 5 discusses the influence of systematic er- rors due to the use of theoretical models, and in Sect. 6 I summarise the physical properties of the known TEPs and explore correlations between various parameters. Those readers interested in the gen- eral properties of TEPs rather than specific systems can skip Sect. 4 without problem. 2 LIGHT CURVE ANALYSIS: JKTEBOP I have modelled the light curves of each TEP using the JKTE- BOP3 code (Southworth et al. 2004a,b). JKTEBOP grew out of the original EBOP program written for eclipsing binary star systems (Popper & Etzel 1981; Etzel 1981) and implementing the NDE model (Nelson & Davis 1972). JKTEBOP uses biaxial spheroids to model the component stars (or star and planet) so allows for depar- tures from sphericity. The shapes of the components are governed by the mass ratio, q, although the results in this work are all ex- tremely insensitive to the value of this parameter. The main parameters of a JKTEBOP fit are the orbital inclina- tion, i, and the fractional radii of the two stars4, rA and rb. The fractional radii are defined as Rb a RA a (1) rA = rb = where RA and Rb are the stellar and planetary radii and a is the orbital semimajor axis. rA and rb correspond to radii of spheres of the same volume as the biaxial spheroids. In JKTEBOP the frac- tional radii are re-parameterised as their sum and ratio: rA + rb k = rb rA = Rb RA (2) because these are only weakly correlated with each other. In general the orbital period, Porb, is taken from the literature and the time of transit midpoint, T0, is included as a fitted parameter in each JKTEBOP run. 2.1 Treatment of limb darkening The limb darkening (LD) of the star is an important 'nuisance pa- rameter' affecting transit light curves which can be parametrised using any of five LD laws in JKTEBOP. Wherever possible the LD coefficients are included as fitted parameters, but when there is in- sufficient information for this the coefficients are fixed at theoreti- cal values. For each light curve I have obtained solutions with each of the five LD laws (see Paper I for their definition) and with both LD coefficients fixed, with the linear coefficient fitted and the non- linear coefficient fixed (hereafter referred to as 'LD fit/fix'), and with both coefficients fitted. Theoretical LD coefficients have been taken from Van Hamme (1993), Claret (2000, 2004a) and Claret & Hauschildt (2003). The tabulated values have been bilinearly interpolated, using the JKTLD code5, to the known effective temperature (Teff ) and surface gravity (log g) of the star. I find that there is usually a spread of 0.1–0.2 in the theoretical LD coefficients for cool stars, so when the nonlinear LD coefficient is not included as a fitted parameter it is perturbed in the Monte Carlo simulations by ±0.1 on a flat distribution to account for this. The dependence on theoretical calculations in this case is still exceptionally small, because the linear and nonlinear coefficients of the LD laws are highly correlated with each other (Southworth et al. 2007a). Once solutions have been obtained for the five LD laws, the fi- nal result is calculated by taking the weighted means of the param- eter values for the four two-parameter LD laws (i.e. the linear law is not used). The parameter errorbars are taken to be the largest of the individual errorbars (see below) plus a contribution to account for scatter of the parameter values from different LD law solutions. 2.2 Error analysis For each solution I run 1000 Monte Carlo simulations (Southworth et al. 2004c, 2005) to provide robust estimates of the 1σ statistical errorbars. The starting parameter values are perturbed 2 In principle, astrometric observations could either replace radial velocity measurements, or augment them and thus provide the missing constraint, but this has not yet been achieved in practise. 3 JKTEBOP is written in FORTRAN77 and the source code is available at http://www.astro.keele.ac.uk/∼jkt/codes/jktebop.html 4 Throughout this work stellar parameters are indicated by a subscripted 'A' and planet parameters by a subscripted 'b', to conform to IAU nomen- clature. 5 JKTLD is written in FORTRAN77 and the source code is available at http://www.astro.keele.ac.uk/∼jkt/codes/jktld.html c(cid:13) 0000 RAS, MNRAS 000, 000–000 Homogeneous studies of transiting extrasolar planets. III. 3 Monte Carlo errorbars are multiplied bypχ 2 for each simulation to avoid sticking artificially close to the origi- nal best fit. If the reduced χ2 of the fit, χ 2 ν , is greater than unity, the ν to account for this. Monte Carlo simulations do not fully account for the presence of correlated ('red') noise, which is an unavoidable reality in high- precision light curves of bright stars. I therefore also run a resid- ual permutation (or "prayer bead") algorithm (Jenkins et al. 2002) with the quadratic LD law. If there is significant correlated noise the residual-permutation errorbars will exceed the Monte-Carlo er- rorbars. I then take the larger of the two error estimates to represent the final errorbars of the photometric parameters. 2.3 Orbital eccentricity Some TEPs have a non-circular orbit which must be accounted for in the light curve analysis. Orbital eccentricity is very difficult to detect from the shape of a transit light curve (Kipping 2008) but can have a significant effect on the resulting parameters (for an example see Sect. 4.13). Non-circular orbits normally become apparent from radial velocity (RV) measurements of the parent stars. These RVs can then be used to determine the eccentricity, e, and the longitude of periastron, ω, of the binary orbit, JKTEBOP has been modified to account for orbital eccentricity by including the possibility of specifying values for either e and ω or the combinations e cos ω and e sin ω. These values and their uncertainties are then simply treated as extra observations, and e and ω (or their combinations) are included as fitted parameters. In this way the uncertainties in e and ω are correctly propagated into the errorbars in the other photometric parameters. I prefer to work with e cos ω and e sin ω rather than e and ω, because the latter two quantities are strongly correlated with each other (e.g. Bruntt et al. 2006). 2.4 Contaminating light It is possible for additional light to contaminate photometric obser- vations of transiting planets. Any extra light from nearby faint stars will dilute the transit depth, causing a systematic error in the light curve parameters. This idea is becoming more important for several reasons. Firstly, the CoRoT satellite has a large point spread func- tion (PSF) which usually contains a number of stars aside from the one hosting a transiting planet. Secondly, observations using tele- scope defocussing are potentially more susceptible to contaminat- ing light. Thirdly, Daemgen et al. (2009) have detected faint com- panions to three transiting systems (TrES-2, TrES-4 and WASP-2) from ground-based high-resolution observations of fourteen TEPs obtained with a lucky-imaging camera. These companions could be bound to their respective transiting systems, or may just be aster- isms. Temporarily ignoring the orbital ephemeris (Porb and T0), there are three main observables in a transit shape: its depth, dura- tion, and the duration of totality (Paper I). From transit light curves we measure three quantities, which in the case of JKTEBOP are rA, rb and i. It is therefore expected to be impossible to fit directly for contaminating light, as this would require measuring four indepen- dent parameters using only three observables. This expectation will now be verified. c(cid:13) 0000 RAS, MNRAS 000, 000–000 Figure 1. Plot of the variation of parameters fitted to a set of synthetic transit light curves with 1 mmag of Gaussian noise added. The synthetic datasets were generated for a range of third light values but fitted with the assumption of L3 = 0. Dotted lines show the input parameter values for the synthetic light curve calculations. 2.4.1 The effect of third light I have explored the possibility of measuring contaminating light by simulating a set of light curves with reasonable parameters (rA + rb = 0.1, k = 0.1, i = 87◦). I added contaminating light (by convention referred to as 'third light' and expressed as a frac- tion of the total system light) by amounts ranging from L3 = 0 to 0.75 in steps of 0.05. These were transformed into typical good ground-based light curves by retaining approximately 400 points within each transit and adding a Gaussian scatter with standard de- viation 1 mmag. These synthetic light curves were then fitted with JKTEBOP under the assumption that L3 = 0. The results (Fig. 1) show that the presence of L3 results in systematic overestimates of rA and underestimates of rb and i. The bottom panel of Fig. 1 shows that the quality of the fit does not get worse as L3 increases. Unaccounted third light therefore biases the resulting parameters without being detectable through its impact on the quality of the fit. As a second test I modelled the same synthetic light curves again, this time fitting for third light. The resulting values of L3 are shown in Fig. 2 and are extremely scattered as well as biased to smaller values. JKTEBOP deliberately does not restrict photometric parameters to physically realistic values (e.g. L3 > 0), to avoid statistical biases in the uncertainties arising from Monte Carlo sim- ulations. Fig. 2 demonstrates that there is a very small amount of information on L3 in a good ground-based light curve, but that this 4 John Southworth Figure 2. Plot of fitted versus input values of third light for the same light curves as in Fig. 1, but with L3 included as a fitted parameter. The dotted line shows parity. Note the large scale on the y-axis. information is far too sparse to be useful. Fig. 3 shows the resulting values of the other main parameters: rb and i are biased towards lower values and there is no trend visible in the sizes of the residu- als. In reality a large value of L3 is not expected because such a bright star would show up in the spectroscopic observations of a transiting system. Fainter stars can be found via high- resolution imaging if they are slightly away from the planet host star (Daemgen et al. 2009). But it may never be possible to rule out the presence of a much fainter star (L3 < 5% depending on the quality of the spectroscopic observations) which almost exactly co- incide with the planet host. As a guide, 5% of third light can be compensated for by increasing rA by 1%, and decreasing rb by 2% and i by 0.1◦. It can therefore change the derived radius of the planet by several percent. 2.4.2 Accounting for third light The observations of Daemgen et al. (2009) make it possible to ac- count for third light when analysing transit light curves. They mea- sured magnitude differences (light ratios) in the SDSS i and z pass- bands. When necessary I have propagated the light ratios to other passbands by convolving synthetic spectra from ATLAS9 model at- mospheres (Kurucz 1979, 1993) with passband response functions made available by the Isaac Newton Group6. Armed with L3 values for the correct passbands, I have in- cluded these in the same way as e and ω. JKTEBOP was modified to accept measured L3 values as observations, and L3 was included as a fitted parameter. Note that several published studies have instead simply subtracted L3 from a light curve before analysis, which is statistically incorrect as it neglects the uncertainty in L3. Daemgen et al. (2009) surveyed fourteen transiting systems and detected faint companions to three of them. The companions Figure 3. Plot of the variation of parameters fitted to the same light curves as in Fig. 1, but in this case with L3 included as a fitted parameter. The dotted lines show parity. are within 0.7–1.6′′ of the transit host stars, and are fainter by gen- erally 4 mag in the i band. Daemgen et al. found that their pres- ence changed the physical properties of the TEPs by roughly 1σ. Of the three affected objects, TrES-2 and TrES-4 are analysed in the current work and WASP-2 is the subject of a separate publica- tion (Southworth et al. 2010). 3 INCORPORATING STELLAR MODELS: ABSDIM Analysis of a transit light curve gives the quantities rA, rb and i 7. From RV measurements of the parent star it is possible to obtain e, ω, and the velocity amplitude of the star, KA. With these ob- servables we remain unfortunately one piece of information short of being able to calculate the full physical properties of the system. An additional constraint is needed, and this is generally supplied by forcing the properties of the star to match the predictions of theoret- ical stellar evolutionary models. To guide this process we can use the spectroscopically measured Teff and metal abundance,(cid:2) Fe the star. In the current work I adopt the method outlined in Paper II, in which the variable governing the solution process is taken to be Kb, the orbital velocity amplitude of the planet. An initial value of Kb H(cid:3), of 6 http://catserver.ing.iac.es/filter/ 7 From transit light curves we also get Porb and T0. The uncertainty in Porb is generally negligible, and T0 does not enter the ABSDIM analysis. c(cid:13) 0000 RAS, MNRAS 000, 000–000 Homogeneous studies of transiting extrasolar planets. III. 5 Table 1. Physical ingredients and coverage of the stellar models used in this work. Yini is the primordial helium abundance, ∆Y /∆Z is the helium-to-metals enrichment ratio, Z⊙ is the solar metal abundance (fraction by mass) and αMLT is the mixing length parameter. Model set Reference Range in mass ( M⊙) Range in metal abundance (Z) Yini Claret Y2 Teramo VRSS DSEP Claret (2004b, 2005, 2006, 2007) Demarque et al. (2004) Pietrinferni et al. (2004) VandenBerg et al. (2006) Dotter et al. (2008) 0.2 to 1.5 0.4 to 5.2 0.5 to 10.0 0.4 to 4.0 0.1 to 5.0 0.01 to 0.05 0.00001 to 0.08 0.0001 to 0.04 0.005 to 0.050 0.000041 to 0.0404 0.24 0.23 0.245 0.23544 0.245 ∆Y ∆Z 2.0 2.0 ∼1.4 2.2 1.6 Z⊙ αMLT Notes Calculated on request 0.02 0.02 0.0198 0.188 0.0189 1.68 1.743 1.913 1.90 1.938 is defined, usually in the region of 150 km s−1, and the full physi- cal properties of the system are calculated using standard formulae (e.g. Hilditch 2001). Armed with the resulting stellar mass, MA, and (cid:2) Fe H(cid:3), I linearly interpolate within tabulated theoretical model results to find the predicted radius and Teff of the star. This process is iteratively repeated whilst varying Kb in order to minimise the figure of merit fom = (cid:18) r(obs) A − (R(pred) ) A σ(r(obs) A /a) 2 (cid:19) +(cid:18) T (obs) eff − T (pred) ) (cid:19) eff σ(T (obs) eff 2 (3) which results in the best-fitting system properties. In principle it is possible to also solve for the age of the system, but in practise the wide variety of evolutionary timescales of stars make this difficult. I therefore step through the possible ages of the star in 0.1 Gyr in- crements, starting at zero age and finishing when the star leaves the main sequence, in order to find the best overall solution. I do not make any attempt to match spectroscopically measured log g val- ues as they are usually much less reliable than the surface gravity of the star calculated from the MA and RA obtained above. The above procedure implicitly applies the strong constraint on stellar density obtained from the light curve analysis (Seager & Mall´en-Ornelas 2003). The uncertainties in the system properties are calculated by a perturbation analysis, in which each input parameter is modified by its 1σ uncertainty and new solutions specified. The uncertainty for each output parameter is then calculated by adding the uncertainties due to each input parameter in quadrature. This perturbation analy- sis has the advantage of yielding detailed error budgets, where the effect of the uncertainty of every input parameter on every output parameter is specified. These error budgets indicate what additional observations are the best for improving our understanding of a spe- cific TEP. 3.1 Which stellar models to use? As outlined above, the physical properties of TEPs are calculated by forcing the properties of the parent star to match theoretical ex- pectations. This dependence on theoretical predictions is a concern and will cause a systematic error. It is well known that whilst theo- retical models are pretty good at reproducing the actual properties of stars, the various model sets are not flawless and do not agree perfectly with each other. The existence of different theoretical model sets for low-mass stars opens the possibility of using several of them and explicitly deducing the systematic errors in TEP properties caused by their use. In Paper II six different sets of theoretical models were inves- tigated and three adopted for calculating the planet properties. The Siess and Cambridge-2007 models were in relatively poor agree- c(cid:13) 0000 RAS, MNRAS 000, 000–000 ment with other models, and the Cambridge-2000 models had a lower coverage of the relevant parameter space than other models. I was therefore left with only three different model sets, which was insufficient to define high-quality systematic error estimates. On top of this, the Padova models included heavy element abundances only up to Z = 0.03 so did not cover quite a few transiting systems. In the current paper I have therefore adopted the same solu- tion procedure as introduced in Paper II, but with a significantly re- vised database of theoretical model predictions and with one more change. Instead of using the Claret models to define my baseline solutions and two other model sets to obtain systematic error es- timates, I have used the unweighted mean and standard deviation of the results from all five model sets to describe the baseline solu- tions and systematic errors. The dependence of the final results on a single model set is therefore broken: all five model sets are treated equally and the choice of which sets of models to use becomes less important. Of the sets of theoretical models included in Paper II, only the Y2 models survive unchanged here (see Table 1). The Claret models have been supplemented by additional calculations for higher metal abundances of Z = 0.06 and 0.07. The third model set used here is Teramo8 (Pietrinferni et al. 2004), and I selected the ones with moderate convective core overshooting (for masses >1.1 M⊙), the standard mass loss law (η = 0.4) and normal el- emental abundances (scaled-solar, i.e. no enhancement of the α- elements). For the fourth model set I acquired the Victoria-Regina (VRSS) models9 (VandenBerg et al. 2006) with scaled-solar ele- mental abundances. In these models the convective core overshoot- ing parameter depends on mass and is empirically calibrated. The fifth and final model set (DSEP) comes from Dartmouth Stellar Evolution Database10 and again comprises the calculations with scaled-solar elemental abundances. I selected those models which follow the standard helium-to-metal enrichment law. The DSEP models include a contraction to the zero-age main sequence which can take tens of Myr. In Fig. 4 I compare the predictions in mass and radius of the five sets of stellar models, for an age of 0.5 Gyr and for the adopted solar chemical composition (which differs between models). The models have a fairly good agreement in the mass–radius plane, par- ticularly near 1.0 M⊙ as they are calibrated on the Sun, but a wider variety in the mass–Teff plane. Also shown in Fig. 4 are the masses, radii and Teff values of the sample of detached eclipsing binary star 8 Obtained from the BaSTI database on 17/11/2009: http://albione.oa-teramo.inaf.it/ 9 Obtained on 18/11/2009 from http://www.cadc-ccda.hia-iha. nrc-cnrc.gc.ca/cvo/community/VictoriaReginaModels/. 10 Obtained on 18/11/2009 from http://stellar.dartmouth.edu/ ∼models/index.html. 6 John Southworth Figure 4. Mass–radius (left) and mass–Teff (right) plots showing the predictions of the five sets of stellar models adopted in this work (solid lines). The predictions are for an age of 0.5 Gyr (to minimise the effects of evolution to and from the ZAMS) and for the solar chemical composition (which varies between models). The measured masses, radii and Teff values of a sample of detached eclipsing binaries (see Paper II) are shown for comparison, using blue errorbars. The Sun is indicated by the usual ⊙; note that the predictions of the models do not pass through the solar values on these plots as the Sun is much older than 0.5 Gyr. systems constructed in Paper II. It can be seen that the disagreement with the models is much larger than the errorbars. This conclusion holds for all chemical compositions for which the five sets of mod- els are available. Similarly, adopting an age either before or beyond the main sequence can provide an agreement for individual eclips- ing binaries, but not for all of them simultaneously. Fig. 4 illustrates what can be expected for the variation of sys- tematic errors with mass. The models agree with each other very well in some regions, so systematics will be minimised, and less well at lower and higher masses, when systematics will be larger. The agreement between models is clearly much better than with the properties of well-studied eclipsing binaries. This means that using the five model sets will lead to only a lower limit on the system- atic errors in the properties of TEPs. A probable upper limit to the systematic effects can be obtained by calculating solutions with an eclipsing binary mass–radius relation instead of a stellar model set; this is applied below and discussed further in Paper II. 3.2 Calculating the physical properties of transiting planets Using the method and theoretical stellar models outlined above, the mass, radius, surface gravity and mean density of the star (MA, RA, log gA, ρA) and of the planet (Mb, Rb, gb, ρb) can be calculated. For each TEP I have obtained results for each of the five stellar model sets and also using an empirical mass–radius relation de- fined by the eclipsing binaries. The final result for each parameter is the unweighted mean of the five stellar-model results. The statis- tical errorbar is taken to be the largest one from these five solutions and the systematic errorbar is taken to be the standard deviation of the parameter values from the five solutions. I include the final Kb value when possible to aid the comparison between different solu- tions for the same planet. Kb is the parameter through which all of the stellar model dependence enters. The surface gravity of the planet, gb, can be calculated from purely geometrical observed quantities (Southworth et al. 2007b) so has no systematic error. Similarly, the stellar density, ρA, is in- dependent of the stellar models (Seager & Mall´en-Ornelas 2003) if MA ≫ Mb is assumed. In addition to the above parameters, I have also calculated the equilibrium temperature (Teq) and Safronov (1972) number (Θ) of the planet. Teq is given by the equation Teq = Teff (cid:16) 1 − A 4F (cid:17)1/4(cid:16) RA 2a (cid:17)1/2 (4) where A is the Bond albedo and F is a heat redistribution factor. Because A and F are not known precisely I instead calculate a modified equilibrium temperature (T ′ eq) which equals Teq if A = 1 − 4F : eq = Teff (cid:16) RA T ′ = Teff (cid:16) rA 2a (cid:17)1/2 2 (cid:17)1/2 (5) The Safronov number is defined as the ratio of the escape velocity to the orbital velocity of the planet: Θ = 1 Vorb(cid:17)2 2 (cid:16) Vesc = (cid:16) a Rb(cid:17)(cid:16) Mb MA(cid:17) = 1 rb Mb MA (6) From Eq. 5 above it can be seen that T ′ eq depends only on the stellar Teff and the fractional radius obtained from the light curve analysis. T ′ eq therefore turns out (like gb) to be independent of the stellar models, but does have some systematic error as it is depen- dent on the effective temperature scale of low-mass (F, G and K) stars. c(cid:13) 0000 RAS, MNRAS 000, 000–000 Homogeneous studies of transiting extrasolar planets. III. 7 Table 2. Parameters from the light curve analyses presented here and in previous works, and used here to determine the physical properties of the TEPs. The orbital periods are taken from the literature, and the bracketed numbers represent the uncertainty in the preceding digits. Systems for which orbital eccentricity was accounted for are indicated with a ⋆ in the column marked "e?". System GJ 436 HAT-P-1 HAT-P-2 HD 149026 HD 189733 HD 209458 OGLE-TR-10 OGLE-TR-56 OGLE-TR-111 OGLE-TR-113 OGLE-TR-132 OGLE-TR-182 OGLE-TR-211 OGLE-TR-L9 TrES-1 TrES-2 TrES-3 TrES-4 WASP-1 WASP-2 WASP-3 WASP-4 WASP-5 WASP-10 WASP-18 XO-1 XO-2 XO-3 XO-4 XO-5 Orbital period (days) 2.64389524 (76) 4.4652934 (93) 5.6334729 (61) 2.8758911 (25) 2.21857578 (80) 3.52474859 (38) 3.101278 (4) 1.211909 (1) 4.0144479 (41) 1.4324757 (13) 1.689868 (3) 3.97910 (1) 3.67724 (3) 2.485533 (7) 3.0300728 (6) 2.4706101 (18) 1.3061864 (5) 3.553945 (75) 2.519961 (18) 2.15222144 (39) 1.846835 (2) 1.33823150 (61) 1.6284246 (13) 3.0927616 (10) 0.94145181 (44) 3.9415128 (28) 2.6158640 (16) 3.1915289 (32) 4.12502 (2) 4.187757 (11) e? Orbital inclination, Fractional stellar Fractional planetary Reference i (degrees) radius, rA radius, rb ⋆ ⋆ ⋆ ⋆ ⋆ 86.43 ± 0.18 86.25 ± 0.22 85.9 ± 1.5 88.0 ± 2.0 85.78 ± 0.25 86.590 ± 0.046 83.87 ± 0.69 79.8 ± 2.4 88.11 ± 0.66 87.7 ± 1.4 83.3 ± 2.4 84.3 ± 1.2 88.0 ± 2.0 82.07 ± 0.69 88.67 ± 0.71 83.80 ± 0.36 82.07 ± 0.17 81.53 ± 0.60 88.0 ± 2.0 84.81 ± 0.17 84.1 ± 1.3 89.0 ± 1.0 85.8 ± 1.1 88.81 ± 0.40 85.0 ± 2.1 89.06 ± 0.84 88.8 ± 1.2 83.89 ± 0.40 89.9+0.1 −3.9 87.04 ± 0.65 0.0731 ± 0.0027 0.0935 ± 0.0025 0.1247 ± 0.0106 0.140+0.012 −0.006 0.1113 ± 0.0031 0.11384 ± 0.00041 0.157 ± 0.009 0.245 ± 0.026 0.0842 ± 0.0038 0.1592 ± 0.0043 0.211 ± 0.020 0.137 ± 0.014 0.1422+0.0150 −0.0083 0.1731 ± 0.0083 0.0964 ± 0.0018 0.1282 ± 0.0035 0.1666+0.0017 −0.0015 0.1802 ± 0.0083 0.1737+0.0057 −0.0089 0.1238 ± 0.0018 0.201 ± 0.010 0.1825+0.0011 −0.0010 0.1847 ± 0.0061 0.0865 ± 0.0041 0.2795 ± 0.0084 0.0886 ± 0.0019 0.1237+0.0024 −0.0047 0.1447 ± 0.0046 0.1300+0.0283 −0.0051 0.1004 ± 0.0049 0.00605 ± 0.00023 0.01051 ± 0.00031 0.00847 ± 0.00082 0.0068+0.0011 −0.0008 0.0175 ± 0.0005 0.01389 ± 0.00006 0.0182 ± 0.0011 0.0241 ± 0.0034 0.01107 ± 0.00067 0.02331 ± 0.00089 0.0198 ± 0.0024 0.0135 ± 0.0013 0.01181+0.00146 −0.00083 0.01910 ± 0.00085 0.01331 ± 0.00035 0.01658 ± 0.00043 0.02731+0.00055 −0.00043 0.0174 ± 0.0012 0.0182+0.0007 −0.0011 0.01643 ± 0.00030 0.0218 ± 0.0011 0.02812+0.00022 −0.00014 0.02050 ± 0.00091 0.01349 ± 0.00065 0.0272 ± 0.0012 0.01166 ± 0.00035 0.01300+0.00033 −0.00070 0.01317 ± 0.00047 0.01124+0.00334 −0.00054 0.01054 ± 0.00073 Paper I This work This work Paper I Paper I Paper I Paper I Paper I Paper I This work Paper I This work This work This work Paper I This work This work This work Paper I Southworth et al. (2010) This work Southworth et al. (2009b) Southworth et al. (2009a) This work Southworth et al. (2009c) Paper I This work This work This work This work 4 RESULTS FOR INDIVIDUAL SYSTEMS 4.1 HAT-P-1 In this section I present the photometric (JKTEBOP) and absolute- dimensions (ABSDIM) analyses of a set of thirty TEPs based on high-quality data. In many cases I adopt the JKTEBOP results from Paper I or later works (Southworth et al. 2009a,b,c, 2010). The final JKTEBOP results of all TEPs are collected in Table 2, which also includes the orbital periods and indicates for which systems a non- circular orbit was adopted. The mass ratio of each TEP is required as an input parameter for the light curve analysis, but in all cases its effect on the solution is negligible. Representative values have been taken from the literature but will not be discussed further. The physical properties of all thirty TEPs are obtained or re- vised in the current work, using the new theoretical model sets H(cid:3) and KA values discussed in Sect. 3. This also requires Teff, (cid:2) Fe for each system. These are summarised in Table 3. The values are mostly unchanged for the fourteen TEPs studied in Paper II, but in a few cases improved values have become available and replace the previous entries. In Papers I and II the individual systems were tackled roughly in order of increasing complexity. The current work reverts to the more structured approach of attacking the TEPs in alphabetical order, beginning with those objects for which a light curve analysis is presented (Sects. 4.1 to 4.15, then moving on to those whose photometric parameters are adopted unchanged from Paper I (Sect. 4.16). c(cid:13) 0000 RAS, MNRAS 000, 000–000 HAT-P-1 was found to be a TEP by Bakos et al. (2007a) from data taken by the HAT survey (Bakos et al. 2002, 2004). Its low mass (0.5 MJup) and large radius (1.2 RJup) make it one of the least dense exoplanets known. Excellent light curves from the FLWO 1.2 m (z band), Lick 1.0 m Nickel (Z band) and Wise 1.0 m tele- scopes were presented by Winn et al. (2007c) and the first two of these were included in Paper I. Since then additional data from the Nickel (Z band) and the Hawaiian 2 m Magnum (V band) tele- scopes have been obtained by Johnson et al. (2008). In this work I have analysed the latter two datasets in order to refine the results from Paper I. In both cases I have adopted solutions with the lin- ear LD coefficient fitted and the nonlinear coefficient fixed but per- turbed in the Monte Carlo simulations ('LD fit/fix'). The residual permutation analyses indicate that correlated errors are important for both datasets. The final light curve parameters are the weighted means of those for the four studied datasets. The results agree well with each other except for k, for which χ 2 ν = 2.8. The errorbar for k has been multiplied by √2.8 to account for this. The light curve fits are plotted in Fig. 5 and summarised in Table A3. They are in good agreement with literature values. The physical properties of HAT-P-1 have been calculated us- ing the five different sets of stellar evolutionary models plus the 8 John Southworth Table 3. Measured quantities for the parent stars which were adopted in the analysis presented in this work. System Velocity amplitude ( m s−1) Teff (K) Reference GJ 436 HAT-P-1 HAT-P-2 HD 149026 HD 189733 HD 209458 OGLE-TR-10 OGLE-TR-56 OGLE-TR-111 OGLE-TR-113 OGLE-TR-132 OGLE-TR-182 OGLE-TR-211 OGLE-TR-L9 TrES-1 TrES-2 TrES-3 TrES-4 WASP-1 WASP-2 WASP-3 WASP-4 WASP-5 WASP-10 WASP-18 XO-1 XO-2 XO-3 XO-4 XO-5 18.34 ± 0.52 Maness et al. (2007) 59.3 ± 1.4 Johnson et al. (2008) 983.9 ± 17.2 P´al et al. (2010) Sato et al. (2005) 43.3 ± 1.2 200.56 ± 0.88 Boisse et al. (2009) 85.1 ± 1.0 Naef et al. (2004) 80 ± 17 Konacki et al. (2005) 212 ± 22 Bouchy et al. (2005a) Pont et al. (2004) 78 ± 14 267 ± 34 TWH08 167 ± 18 Moutou et al. (2004) 120 ± 17 82 ± 16 Udalski et al. (2008) 510 ± 170 Snellen et al. (2009) 115.2 ± 6.2 Alonso et al. (2004) 181.3 ± 2.6 O'Donovan et al. (2006) Pont et al. (2008) Sozzetti et al. (2009) 369 ± 11 97.4 ± 7.2 Mandushev et al. (2007) 111 ± 9 Wheatley et al. (2010) Pont et al. (2008) P´al et al. (2010) Sato et al. (2005) 3500 ± 100 Bean et al. (2006) 5975 ± 50 Bakos et al. (2007a) 6290 ± 60 6147 ± 50 5050 ± 50 Bouchy et al. (2005b) Santos et al. (2004) 6117 ± 50 Santos et al. (2006) 6075 ± 86 Santos et al. (2006) 6119 ± 62 5044 ± 83 Santos et al. (2006) 4804 ± 106 Santos et al. (2006) 6210 ± 59 Gillon et al. (2007) 5924 ± 64 6325 ± 91 Udalski et al. (2008) Snellen et al. (2009) 6933 ± 58 5226 ± 50 Santos et al. (2006) 5795 ± 73 Ammler-von Eiff et al. (2009) 5650 ± 75 6200 ± 75 6110 ± 50 5150 ± 80 Triaud et al. (2010) 6400 ± 100 Pollacco et al. (2008) 5500 ± 100 Gillon et al. (2009) 5700 ± 100 Gillon et al. (2009) 4675 ± 100 Christian et al. (2009) 6400 ± 100 Hellier et al. (2009) Sozzetti et al. (2009) Sozzetti et al. (2009) Stempels et al. (2007) 153.6 ± 3.0 Triaud et al. (2010) 290.5 ± 9.5 Tripathi et al. (2010) 242.1+2.8 Triaud et al. (2010) −3.1 268.7 ± 1.8 Triaud et al. (2010) 553.1 ± 7.5 1816.9 ± 2.0 Triaud et al. (2010) 116.0 ± 9.0 McCullough et al. (2006) 5750 ± 50 McCullough et al. (2006) Johnson et al. (2009b) 85 ± 8 Burke et al. (2007) Johns-Krull et al. (2008) 1488 ± 10 Winn et al. (2009) 163 ± 16 McCullough et al. (2008) 6397 ± 70 McCullough et al. (2008) 5340 ± 50 Burke et al. (2007) 6429 ± 75 144.9 ± 2.0 P´al et al. (2009) 5370 ± 70 P´al et al. (2009) Reference H(cid:3) (cid:2) Fe −0.03 ± 0.2 Bonfils et al. (2005) 0.13 ± 0.05 Bakos et al. (2007a) 0.14 ± 0.08 P´al et al. (2010) 0.36 ± 0.05 Sato et al. (2005) −0.03 ± 0.05 Bouchy et al. (2005b) 0.02 ± 0.05 Santos et al. (2004) 0.28 ± 0.10 Santos et al. (2006) 0.25 ± 0.08 Santos et al. (2006) 0.19 ± 0.07 Santos et al. (2006) 0.15 ± 0.10 Santos et al. (2006) 0.37 ± 0.07 Gillon et al. (2007) 0.37 ± 0.08 Pont et al. (2008) 0.11 ± 0.10 Udalski et al. (2008) −0.05 ± 0.20 Snellen et al. (2009) 0.06 ± 0.05 Santos et al. (2006) 0.06 ± 0.08 Ammler-von Eiff et al. (2009) −0.19 ± 0.08 Sozzetti et al. (2009) 0.14 ± 0.09 Sozzetti et al. (2009) 0.23 ± 0.08 Stempels et al. (2007) −0.08 ± 0.08 Triaud et al. (2010) 0.00 ± 0.20 Pollacco et al. (2008) −0.03 ± 0.09 Gillon et al. (2009) 0.09 ± 0.09 Gillon et al. (2009) 0.03 ± 0.20 Christian et al. (2009) 0.00 ± 0.09 Hellier et al. (2009) 0.02 ± 0.05 McCullough et al. (2006) 0.45 ± 0.05 Burke et al. (2007) −0.18 ± 0.05 Johns-Krull et al. (2008) −0.04 ± 0.05 McCullough et al. (2008) 0.05 ± 0.06 P´al et al. (2009) empirical mass–radius relation from Paper II. The individual solu- tions are given in Table A4 and then compared with literature val- ues, where a good agreement is found. 4.2 HAT-P-2 HAT-P-2 was discovered to be a TEP system by Bakos et al. (2007b), under the name HD 147506. It is a very bright system (V = 8.7) with a massive planet (Mb = 8.74 MJup) in a highly eccentric orbit. It has been found not to exhibit a spin–orbit mis- alignment (Winn et al. 2007a; Loeillet et al. 2008), in contrast to other massive TEPs on eccentric orbits (Johnson et al. 2009a). Good z-band light curves of HAT-P-2 have been published by Bakos et al. (2007b), covering one transit with the FLWO 1.2 m, and by P´al et al. (2010), covering another six transits with the FLWO 1.2 m and the Perkins telescopes. Here I analyse the FLWO datasets together, omitting the small amount of data taken on the night of 2007/03/18, as well as the Perkins data. One com- plication is the orbital eccentricity: this was accounted for us- ing the method discussed in Sect. 2.3 and adopting the constraints e cos ω = −0.5152 ± 0.0036 and e sin ω = −0.0441 ± 0.0084 (P´al et al. 2010). In both cases correlated errors were not important and the LD fit/fix solutions were adopted. The best fits are shown in Fig. 6. The two light curve solutions unfortunately do not agree very well (9.3σ discrepancy in k). I therefore adopt the FLWO 1.2 m results, as these are the much more extensive of the two sets of data and have full coverage of the transit phases. The FLWO results agree well with those of P´al et al. (2010), for which most of the data come from, but have a larger rA and rb than other literature values (Table A7). As expected given the light curve results, my ABSDIM analysis returns system properties in good agreement with those of P´al et al. (2010) but not with other literature studies (Table A8). The prime mover in the most recent solutions is rA, which has a strong effect on the density of the star and thus the other physical properties. The radius of the planet is uncertain by 10%, despite the existence of a high-quality light curve for HAT-P-2, because the transit depth is shallow (0.6%). An improved photometric study is warranted. 4.3 OGLE-TR-113 Like OGLE-TR-132 (studied in Paper I), OGLE-TR-113 was iden- tified as a possible planetary system by Udalski et al. (2002) and its nature was confirmed by Bouchy et al. (2004) using the OGLE light curve and new RV measurements. It was independently confirmed as a TEP by Konacki et al. (2004), also from the OGLE light curve and high-precision RVs, and an abundance analysis of the parent star has been presented by Santos et al. (2006). Whilst OGLE-TR- 113 exhibits a deep transit, its photometric tractability is hindered by the presence of a brighter star only 3′′ away. Apart from the OGLE discovery observations (Udalski et al. 2002), three photometric studies of OGLE-TR-113 have been pub- lished. Gillon et al. (2006) used the ESO New Technology Tele- scope (NTT) and SUSI2 imager to observe two transits in the R band, and obtained what is currently the best light curve of OGLE- c(cid:13) 0000 RAS, MNRAS 000, 000–000 Homogeneous studies of transiting extrasolar planets. III. 9 Figure 6. Phased z-band light curves of HAT-P-2 compared to the best fits found using JKTEBOP and the quadratic LD law. The best fits and residuals are offset for display purposes. The upper light curve is from the FLWO 1.2 m (Bakos et al. 2007b; P´al et al. 2010) and the lower is from the Perkins 1.8 m (P´al et al. 2010). Figure 5. Phased light curves of HAT-P-1 compared to the best fits found using JKTEBOP and the quadratic LD law in Paper I and in this work. The best fits and residuals are offset from unity and zero fluxes, respectively, for display purposes. The light curves are, from top to bottom, FLWO z- band (Winn et al. 2007c), Lick Z-band (Winn et al. 2007c), Magnum V - band (Johnson et al. 2008) and Nickel Z-band (Johnson et al. 2008). a few Gyr, and in several cases is up against the edge of the stellar model grid at 20 Gyr. Aside from that, the properties of the star and planet are rather well-determined but would benefit from an improved KA value as well as a better light curve. The agreement with literature studies is good, although it seems that in some cases the published errorbars are smaller than one would expect. TR-113. Snellen & Covino (2007) observed a K-band transit and an occultation of the system using NTT/SOFI, and detected the oc- cultation with a significance of 2.8σ. D´ıaz et al. (2007) obtained V -band photometry of one transit using a Very Large Telescope (VLT) and the VIMOS instrument; additional data taken in the I and Ks bands are unavailable and of lower quality. In this work I analyse the Gillon et al. observations, the Snellen & Covino transit light curve, and the V -band data obtained by D´ıaz et al. For the second of these three datasets I allowed for light from the planet with a surface brightness ratio of 0.07± 0.02. The surface brightness ratio is a parameter of the JKTEBOP model which is important for eclipsing binary systems but usually left at zero for transiting systems due to the faintness of the planet with respect to the star. The best fits are shown in Fig. 7 and given in Table A12. In all three cases correlated noise is not important. For the Snellen light curve I had to adopt solutions with both LD coeffi- cients fixed, but for the other two I was able to use the LD fit/fix so- lutions. The final results for the Gillon and Snellen data are in good agreement. The solution of the D´ıaz data prefers a rather higher i and lower rA and rb. I therefore combined the Gillon and Snellen data results to obtain the final light curve parameters. The resulting physical properties of OGLE-TR-113 are given in Table A13. The system age is constrained only to be more than c(cid:13) 0000 RAS, MNRAS 000, 000–000 4.4 OGLE-TR-182 OGLE-TR-182 is the sixth TEP discovered as a result of the OGLE search for light variability in selected fields in the Southern hemi- sphere. Its discovery and analysis was presented by Pont et al. (2008), which remains the only study of this object to date. OGLE- TR-182 is difficult because of its faintness (V = 16.8 and I = 15.9) and crowded field. Pont et al. (2008) obtained 24 RV mea- surements using VLT/FLAMES/UVES, and a light curve with VLT/FORS1. The VLT light curve is analysed here and is rather affected by correlated noise. Including the linear LD coefficients as a fitted parameter gives substantially better fits than with both LD coeffi- cients fixed, but the data cannot support the determination of both LD coefficients. I therefore adopt the LD fit/fix solutions (see Fig. 8 and Table A15), which are not in good agreement with Pont et al. (2008). Compared to these authors I find a solution with a lower i and a correspondingly larger star and planet. The physical properties of OGLE-TR-182 are summarised in Table A16 and point to a planet with a rather low density of 0.33 ρJup. However, my results are rather different to those of Pont et al. (2008), and are in poorer agreement with the measured spectroscopic Teff and (rather uncertain) log g measurement. My analysis procedure is more sensitive to the quality of the light curve 10 John Southworth Figure 7. Phased light curves of OGLE-TR-113 with the best fits found us- ing JKTEBOP and the quadratic LD law and residuals of the fits. From top to bottom the datasets are Gillon et al. (2006) (R-band), Snellen & Covino (2007) (Ks-band, binned by ×5 for display purposes) and D´ıaz et al. (2007) (V -band). Figure 8. Phased VLT light curve of OGLE-TR-182 from Pont et al. (2008) compared to the best fit found using JKTEBOP and the quadratic LD law. The residuals are offset from zero for display purposes. Figure 9. Phased VLT light curve of OGLE-TR-211 from Udalski et al. (2008) compared to the best fit found using JKTEBOP and the quadratic LD law. The residuals are offset from zero for display purposes. than the more 'global' approach taken by Pont et al., so is poten- tially more susceptible to correlated noise. This possibility should be investigated by acquiring a new light curve, under good seeing conditions to cope with the crowded field. 4.5 OGLE-TR-211 OGLE-TR-211 is the seventh TEP discovered using OGLE data (Udalski et al. 2008). Its relative faintness means that the available follow-up photometry and spectroscopy is not definitive. The par- ent star is more massive and also more evolved than the Sun, which results in the transit being rather shallow (Fig. 9). Here I analyse the VLT light curve presented by Udalski et al. (2008), ignoring the observational errors supplied with the data which are quite dis- cretised (the only values are 0.001, 0.002 and 0.003) and contribute to instability in the light curve solution. I am not able to get a determinate solution to the VLT data. Possible fits occupy a locus extending from a high i with low rA to a lower i with a large rA. I have therefore calculated solutions for a range of i values and retained only those which in the ABSDIM analysis result in a Teff within a conservative 3σ of the observed value. The observed stellar log g did not provide a useful constraint. Allowable solutions extend from i = 90◦ down to a sharp cut-off around i = 86.25◦ so I present solutions for i = 86, 88, and 90◦ in Table A17. For the final result I accept the LD fit/fix solutions for i = 88◦ but specify errors which account for both the Monte Carlo errorbars and the variation between the different solutions (Table A18). The correlated errors are again important, as can be seen in Fig. 9. The physical properties of OGLE-TR-211 are shown in Table A19 and are in reasonable agreement with those of Udalski et al. (2008) except for the planetary mass. Mb depends mainly on the measured KA, for which both studies have used the same value, so it is not clear why such a discrepancy should arise. Table A19 includes the first determinations of the age and density of the star, planetary equilibrium temperature (which is quite high −55 K) and Safronov number. The system age is at T ′ eq = 1686+90 c(cid:13) 0000 RAS, MNRAS 000, 000–000 Homogeneous studies of transiting extrasolar planets. III. 11 Figure 11. Phased light curve of the transits of TrES-2 from Holman et al. (2007) compared to the best fit found using JKTEBOP and the quadratic LD law with the linear LD coefficient included as a fitted parameter. The residuals of the fit are plotted at the bottom of the figure, offset from zero. my solution corresponds to a significantly higher i (3σ) and lower rA and rb (both 6σ). The ABSDIM results for OGLE-TR-L9 (Table A25) are in rea- sonable agreement with those of Snellen et al. (2009), which is sur- prising given the differences in the photometric parameters. One reason for this is that the ABSDIM solution is governed mainly by the observed Teff (which has a relative uncertainty of 0.8%) rather than by rA (5%). I find that OGLE-TR-L9 b has one of the high- eqs (2039 ± 51 K) of any known TEP. Aside from its faint- est T ′ ness, this planet is an excellent candidate for multicolour 'transmis- sion photometry' to detect variations in planetary radius with wave- length due to atmospheric opacity effects (Fortney et al. 2008). The system would benefit from additional spectroscopy to provide im- proved measurements of (cid:2) Fe H(cid:3) and KA. This subsection completes my analysis of the TEPs discovered from OGLE survey data. 4.7 TrES-2 TrES-2 was the second TEP discovered by the Trans-Atlantic Exo- planet Survey (O'Donovan et al. 2006) and is both larger and more massive than Jupiter. Excellent ground-based light curves of three transits were obtained and studied by Holman et al. (2007), in close collaboration with Sozzetti et al. (2007). The relatively low orbital inclination of this system (i = 83.8◦) help rA and rb to be deter- mined to a high accuracy. The Holman z-band light curve of TrES-2 was studied in Pa- per I, but is revisited here because Daemgen et al. (2009) have since found a fainter star separated from TrES-2 by 1.089 ± 0.008 arc- sec and with a magnitude difference of ∆z = 3.429 ± 0.010 (see Sect. 2.4). interest because Mislis & Schmitt (2009) found a possible decrease in the transit duration between their own and Holman's observations. This would most likely in- dicate that the orbital inclination is getting lower, which in turn points to the presence of a low-mass third body in the TrES-2 sys- tem. Scuderi et al. (2009) obtained new data which did not confirm TrES-2 is of additional Figure 10. Phased GROND light curves of OGLE-TR-L9 from Snellen et al. (2009) compared to the best fit found using JKTEBOP and the quadratic LD law. From top to bottom the light curves are g, r, i and z. The residuals are offset from zero to the base of the figure. relatively well determined (2.6+0.6 +0.4 −0.7 −0.3 Gyr) because the star has evolved away from the zero-age main sequence. OGLE-TR-211 would certainly benefit from additional spectroscopic and photo- metric observations. 4.6 OGLE-TR-L9 OGLE-TR-L9 was discovered within the OGLE-II survey data (Udalski et al. 1997) by Snellen et al. (2009), and is a relatively massive planet orbiting a rapidly-rotating (Vsini = 39 km s−1) F3 V star. High-quality follow-up light curves were obtained by Snellen et al. using the newly-commissioned GROND instrument (Greiner et al. 2008) on the 2.2 m telescope at ESO La Silla. GROND is a CCD imager which utilises dichroics to observe si- multaneously in seven passbands (SDSS griz and near-infrared JHK). In the case of OGLE-TR-L9 the JHK data were too noisy to be useful, but the griz data are of good quality (Fig. 10). The griz observations have been analysed here (Tables A20 to A23). The gri light curves are good enough to support LD fit/fix so- lutions but for the z data both LD coefficients were held fixed. The parameters for the four light curves were combined to obtain the final photometric results (Table A24). Their agreement with those of Snellen et al. (2009) is not good – i and rA are correlated and c(cid:13) 0000 RAS, MNRAS 000, 000–000 12 John Southworth this hypothesis, but Mislis et al. (2009) have since resurrected the changing i. In addition, Rabus et al. (2009) presented a transit tim- ing study of TrES-2 which found a small sinusoidal perturbation with a 0.2 day period, but with only moderate statistical signifi- cance. TrES-2 is in the field of view of the NASA Kepler satellite (Koch et al. 2010), so a light curve of remarkable quality should become available for future investigations of these possibilities. Here I reanalyse the z-band light curve of Holman et al. (2007), this time with the incorporation of a third light value of L3(z) = 0.0408 ± 0.0004. From the adopted Teff of TrES-2 A (Table 3) and the magnitude differences in i and z (Daemgen et al. 2009) I find that the fainter companion star has Teff = 4390± 70 K (consistent with the ∼K5 spectral type found by Daemgen et al. 2009), and is substantially further away than TrES-2 so is not phys- ically bound to the system. The results of the JKTEBOP analysis are given in Table A26 and I adopt the LD fit/fix results. Correlated noise is not impor- tant. Table A27 shows a comparison with the results from Paper I, in which the analysis did not account for the faint companion star: k has decreased by 1σ whilst rA + rb and rA become smaller by less than 0.5σ. The best fit is shown in Fig. 11. The physical properties of the TrES-2 system are summarised in Table A28. My results agree with literature studies within the er- rors, although MA and Mb are larger by roughly 1σ. This can be traced back to the slightly smaller rA found above. More precise spectroscopic Teff and (cid:2) Fe H(cid:3) values would allow improved system parameters to be obtained, as would better photometry. This is cur- rently being obtained by the Kepler satellite, and a stunning light curve of TrES-2 can already be inspected in Gilliland et al. (2010). 4.8 TrES-3 TrES-3 is one of the more massive TEPs (Mb = 1.91 MJup) and orbits a rather cool star (Teff = 5650 K) with an orbital period of only 1.3 d. It was identified as a TEP by O'Donovan et al. (2007) and a discovery-quality light curve has also been obtained by the SuperWASP survey (Collier Cameron et al. 2007). Follow-up tran- sit photometry has been presented by Sozzetti et al. (2009) and Gibson et al. (2009), and occultation (secondary eclipse) observa- tions have been secured by Winn et al. (2008a). A space-based light curve was obtained by the EPOXI mission (Ballard et al. 2009) but has not yet been published. The O'Donovan et al. (2007) B- and z-band photometry of TrES-3 is of good quality, and the transit observations of the follow- up papers (Sozzetti et al. 2009; Gibson et al. 2009) are excellent. This, plus the fact that TrES-3 has a relatively low orbital inclina- tion (82◦), means that the light curve parameters can be obtained to an unusually high precision. In this work I analyse the B- and z-band data from O'Donovan et al. (2007), the V gri-band obser- vations from Sozzetti et al. (2009), and the Liverpool Telescope (LT) RISE measurements from Gibson et al. (2009). The last of the datasets, which has a passband of approximately V +R, was sorted in phase and then binned down by a factor of 20 (from 11 350 to 568 datapoints) to ease the computational burden. For the solutions of the seven light curves I adopt those with both LD coefficients fixed for V , r and i, and the LD fit/fix solu- tions for the remainder (Fig. 12). Correlated noise is important for the V , g and LT datasets. For most of the datasets I find that the residual-permutation algorithm returns errorbars which are signifi- cantly asymmetric, with larger upper errorbars than lower errorbars for both rA and rb, but that the Monte Carlo errorbars are close to symmetric. This implies that the red noise in the light curves is Figure 12. Phased light curves of the transits of TrES-3 compared to the best fits found using JKTEBOP and the quadratic LD law. Successive datasets and residuals are offset in flux for display purposes. From top to bottom are the B and z data (O'Donovan et al. 2007), V gri observations (Sozzetti et al. 2009) and LT V +R measurements (Gibson et al. 2009). biasing the results towards larger component radii, and would not have come to light if correlated noise was accounted for simply by rescaling the observational errors. I accordingly end up with asymmetric errorbars for each light curve solution. The V and g solutions are quite uncertain and dis- crepant with the other results, so were rejected. To calculate the final photometric result I combined the solutions of the remaining five light curves by multiplying their probability density functions. The result is given in Table A36 and is found to be in good agree- ment with literature results. The physical properties of the TrES-3 system are summarised in Table A37, and agree well with published values. For all five stellar model grids the best solution was found for zero age, and there is a possibility that edge effects will cause the uncertainties c(cid:13) 0000 RAS, MNRAS 000, 000–000 Homogeneous studies of transiting extrasolar planets. III. 13 Figure 13. Phased light curves of the transits of TrES-4 compared to the best fits found using JKTEBOP and the quadratic LD law. The upper data are in the z band and the lower data are in the B band. The fits and residuals are offset in flux for display purposes. to be slightly underestimated for this object (for an example of this phenomenon see HD 189733 in Paper I). 4.9 TrES-4 The planetary system TrES-4 was discovered by Mandushev et al. (2007), and is noteworthy for having a very hot (T ′ eq = 1861 K) and low-density (0.15 ρJup) planet. Revised physical properties of the system have been presented by TWH08 and Sozzetti et al. (2009), and near-IR observations of the secondary eclipse show no evidence for orbital eccentricity (Knutson et al. 2009). Like TrES-2 and WASP-2, high-resolution imaging observa- tions by Daemgen et al. (2009) have detected a faint companion to TrES-4. It resides at an angular separation of 1.555 ± 0.005 arcsec and is fainter by ∆i = 4.560 ± 0.017 and ∆z = 4.232 ± 0.025 than TrES-4 A. These result in third light contributions of L3(i) = 0.0150 ± 0.0002 and L3(z) = 0.0199 ± 0.0005, which are taken into account in the light curve analysis as described in Sect. 2.4. In this work I study the high-precision B and z light curves presented in the discovery paper (Mandushev et al. 2007), and pro- vide the first results to fully account for the third light contribu- tion. Using ATLAS9 model spectra I have propagated the third light to the B-band, finding L3(B) = 0.0040 ± 0.0003 and a Teff = 4206 ± 78 K for the companion star. The fainter star is more than twice as distant as TrES-4, so is not bound to the plane- tary system. The B data are rather sparse and do not allow LD coefficients to be fitted for (Fig. 13). The z observations cover two transits so were studied with Porb left as a fitted parameter; the LD fit/fix solutions were adopted. Correlated errors are important for both datasets. The two light curve solutions agree reasonably well (1.5σ) c(cid:13) 0000 RAS, MNRAS 000, 000–000 Figure 14. Phased light curves of WASP-3 compared to the best fits found using JKTEBOP and the quadratic LD law. From top to bottom the light curves are LT V +R (Gibson et al. 2008), which have been binned by a factor of ten for analysis, Keele R and IAC I (Pollacco et al. 2008). The residuals are offset from zero to the base of the figure. within the rather large errorbars, and were combined by weighted mean to find the final parameter values. This solution corresponds to a smaller i, and larger star and planet than found previously (Ta- ble A40). From Fig. 1 we would expect that accounting for third light would lead to a smaller rA and larger rb and i, which is only partially concordant with the current situation. The variation in pa- rameter values must therefore be due to the different analysis meth- ods used. The physical properties of TrES-4 are given in Table A41 and conform to a more evolved star than found by previous studies, as expected for the larger rA found above. The properties of the planet agree well with literature values, but are quite uncertain. The AB- SDIM error budget shows that an improved light curve and addi- tional RV measurements would benefit this system. All four TEPs discovered by TrES have now been analysed in the current series of papers. 4.10 WASP-3 WASP-3 was identified as a possible TEP by Street et al. (2007) based on SuperWASP data (Pollacco et al. 2006). Confirmation of its planetary nature was provided by Pollacco et al. (2008), who presented four transit light curves obtained from various sources and found WASP-3 b to be one of the most strongly irradiated TEPs (T ′ eq = 2028 K). High-precision light curves from the LT have since been presented and analysed by Gibson et al. (2008). Here I 14 John Southworth study these data, plus the Keele R-band and IAU 80 cm I-band ob- servations from Pollacco et al. (2008). The other two light curves (IAC 80 cm V and SuperWASP) have either large systematics or a large scatter. I have binned consecutive sets of 10 datapoints of the LT light curve in order to limit CPU time; the sampling rate of the binned data is 30 s. For the LT data I adopt the LD fit/fix results and find signifi- cantly larger errorbars from the residual-permutation analysis; cor- related noise is clearly visible in these data in Fig. 14. The other two datasets also contain significant red noise, and the LD fixed so- lutions were adopted. The agreement between the three light curve solutions is excellent so they have been combined into a weighted mean (Table A45). Apart from a 2σ larger rb, the final values agree well with the studies of Pollacco et al. (2008) and Gibson et al. (2008) but not with the preliminary results given by Damasso et al. (2009). The physical properties of the WASP-3 system were originally calculated using KA = 251.2 ± 9.3 km s−1 (Pollacco et al. 2008). After this work was completed a revised KA of 276.0 ± 11.0 was presented by Simpson et al. (2009), who also found that the angle between the planetary orbit and the stellar spin was λ = 15+10 ◦. −9 Shortly before the current work was submitted a further study of WASP-3 was produced (Tripathi et al. 2010), containing new re- sults including KA = 290.5+9.8 −9.2 km s−1 and λ = 3.3+2.5 The third and most recent KA has been used to obtain the physical properties of the WASP-3 system (Table A46). I find that WASP-3 b has a rather larger mass and radius than most literature studies, except for that of Tripathi et al. (2010). More precise mea- ◦. −4.4 surements of Teff and(cid:2) Fe H(cid:3) for WASP-3 would be useful to improve our understanding of its physical properties. More extensive spec- troscopy would also be useful to pin down KA, for which a variety of measurements are currently available. 4.11 WASP-10 WASP-10 was found to be a TEP system by Christian et al. (2009), and is notable for containing a fairly massive (3.2 MJup) planet transiting a small (0.70 R⊙) and low-mass (0.75 M⊙) star. The transit events are a generous 3.0% deep, so photometric follow- up of this system is comparatively easy. Christian et al. (2009) ob- tained data from the 0.8 m Tenagra and 1.2 m Mercator telescopes, but unfortunately none of the datasets cover a full transit event. High-precision follow-up photometry of one complete transit of WASP-10 was obtained by Johnson et al. (2009b), using a novel orthogonal frame transfer CCD (OPTIC) to shape the point spread function and thus obtain a scatter of only 0.5 mmag per datapoint at a reasonable sampling rate. I analyse the OPTIC observations to obtain the photometric parameters, and the Mercator data to pro- vide a consistency check. One complication for WASP-10 is its eccentric orbit. This is handled in the way described in Sect. 2.3, using the con- straints e cos ω = −0.045 ± 0.02 and e sin ω = 0.023 ± 0.04 (Johnson et al. 2009b). For both light curves I find that corre- lated noise is unimportant and that the LD fit/fix solutions are best (Fig. 15). The two sets of results agree well (Table A49) and I adopt the OPTIC ones as final. The values agree well with those of Johnson et al. (2009b) but my errorbars are rather larger, due in part to the inclusion of several different LD laws rather than the reliance on only one. The agreement with Christian et al. (2009) is less good but still acceptable. The ABSDIM analysis returns results (Table A50) which are again in good agreement with those of Johnson et al. (2009b) but Figure 15. Phased light curves of WASP-10 compared to the best fit found using JKTEBOP and the quadratic LD law. The upper curve and residuals represent the Mercator data (Christian et al. 2009) and the lower curve and residuals are the OPTIC data (Johnson et al. 2009b). All offsets are additive in flux. with larger errorbars. The agreement between different model sets is unusually poor for WASP-10, resulting in systematic errorbars which are a significant fraction of the random errorbars and as large as the total errorbars given by Johnson et al. Table A50 provides the first measurement of Θ for WASP-10 and also corrects a calcula- tion error in the T ′ eq listed in one of the published studies of this system. An improved photometric study of WASP-10 would settle the existing disagreement on its k value, and the system would also H(cid:3) measurements. be favoured by more precise Teff and(cid:2) Fe After the above study of WASP-10 was completed, new datasets on this star were presented by Dittmann et al. (2010) and Krejcova et al. (2010) and a disagreement over the system proper- ties became manifest. The two new studies both prefer system prop- erties similar to those of Christian et al. (2009) but discrepant with the results – based on a much better light curve – of Johnson et al. (2009b). This will be revisited in the future, once the newer data and perhaps further observations become available. 4.12 XO-2 The second planet discovered by the XO project (McCullough et al. 2006) is noteworthy for having a parent star which is very metal- rich ((cid:2) Fe H(cid:3) = 0.45) and a member of a common proper mo- tion binary system (Burke et al. 2007). Good transit light curves from the Perkins telescope were published in the discovery paper (McCullough et al. 2006), and from the FLWO 1.2 m by the Tran- sit Light Curve Project (Fernandez et al. 2009). The system has also been observed as part of the NASA EPOXI mission (Ballard et al. 2009). In this work I analyse the Perkins and FLWO observations (Fig. 16). In the former case correlated noise is important and in c(cid:13) 0000 RAS, MNRAS 000, 000–000 Homogeneous studies of transiting extrasolar planets. III. 15 Figure 16. Phased light curves of XO-2 compared to the best fits found using JKTEBOP and the quadratic LD law. The upper light curve is R-band from the Perkins telescope (Burke et al. 2007) and the lower one is z-band from the FLWO 1.2 m telescope (Fernandez et al. 2009). The residuals are offset from zero to the base of the figure. the latter it is not. I adopt the LD fit/fix results for both datasets and combine their probability density functions to find the final re- sults. The solutions have i ∼ 90◦ and thus asymmetric errorbars. The agreement between the two light curves and versus published values is good (Table A53). The ABSDIM analysis is complicated by the high (cid:2) Fe ble A54). The results using the five different theoretical models are scattered, giving systematic errors which are larger than the statisti- cal ones for the two most-affected quantities, MA and a. However, the agreement with other studies is good. A more precise KA value would be profitable. H(cid:3) (Ta- 4.13 XO-3 XO-3 was discovered by Johns-Krull et al. (2008) to be a TEP which is so massive it is near the 13 MJup value which represents the minimum mass of a brown dwarf. Johns-Krull et al. (2008) pre- sented two alternative sets of physical properties for the system, the first of which put XO-3 b at 13.25 ± 0.64 MJup but yielded a relatively poor fit to the observed transit light curve. The sec- ond set ignored the spectroscopic measurement of log gA, yield- ing M2 = 12.03 ± 0.46 MJup and a much better fit to the light curve. The latter solution is preferable because light curve shapes are more reliable than spectroscopically-derived surface gravities for late-type dwarf stars. A follow-up study of XO-3 was presented by Winn et al. (2008b), based on high-quality new light curves and also ignoring the spectroscopic log gA. Additional observations have been pre- sented by Winn et al. (2009). In this work I analyse the z-band pho- tometry obtained by Winn et al. (2008b) using the FLWO 1.2 m, c(cid:13) 0000 RAS, MNRAS 000, 000–000 Figure 17. Phased light curves of XO-3 compared to the best fits found using JKTEBOP and the quadratic LD law. From top to bottom the light curves are KeplerCam z-band (Winn et al. 2008b), and KeplerCam r-band and FLWO I-band (Winn et al. 2009). The residuals are offset from zero. and the r-band FLWO 1.2 m and I-band Nickel datasets presented by Winn et al. (2009). The photometric observations presented by Johns-Krull et al. (2008) are not included because they comprise many datasets of only moderate quality. An important property of XO-3 is its substantial orbital ec- centricity (e = 0.28), which is a common feature of the more mas- sive planets (Southworth et al. 2009c) and might indicate that they are a different population of objects to their less massive cousins. XO-3 is also known to exhibit a large spin-orbit misalignment (H´ebrard et al. 2008; Winn et al. 2009) suggestive of dynamical evolution through gravitational interactions (e.g. Fabrycky & Winn 2009). In the following analysis I adopt the constraints e = 0.2884 ± 0.0035 and ω = 346.3 ± 1.3, taken from Winn et al. (2009). Compared to a solution assuming a circular orbit, rA and rb decrease by 0.0033 and 0.00029 respectively, which is less than 1σ in both cases. Correlated errors are unimportant for all three light curves, and in each case the best solutions are LD fit/fix. The z and I data agree well but the r results have a higher i and a 2.5σ lower rA (Table A58). The most extensive dataset is z, so I have combined the results from this and I and rejected the r results as discrepant. The best fits are plotted in Fig 17. My results agree well with those of Winn et al. (2009) and with the second of the two alternative solutions given by Johns-Krull et al. (2008). When determining the physical properties of the XO-3 sys- tem I adopted increased uncertainties of 75 K in Teff and 0.05 dex in (cid:2) Fe H(cid:3), to allow for the possibility of systematic errors in these 16 John Southworth Figure 18. Phased light curve of XO-4 from the Perkins telescope (McCullough et al. 2008). The blue line shows the best fit from JKTEBOP using the quadratic LD law. The residuals are offset from zero. values for low-mass stars such as XO-3 A, and to account for the low spectroscopic gravity value (3.95 ± 0.06 versus 4.23 ± 0.04) in the discovery paper (Johns-Krull et al. 2008). Like the JKTE- BOP outcome, the results of the ABSDIM analysis agree well with those of Winn et al. (2009) but not with the preferred solution of Johns-Krull et al. (2008). The mass of the planet is Mb = 11.8 ± 0.5 MJup, which is close to but below the 13 MJup dividing line be- tween planets and brown dwarfs. The VRSS model results disagree strongly with those of the other models, so were not included in the eq of XO-3 b is high at 1729± 34 K, making it final analysis. The T ′ an interesting object for the study of planetary atmospheres. Aside from the VRSS models, XO-3 is one of the best-measured TEPs. An improved spectroscopic study, incorporating the best log gA value given in Table A59, would be the best way of improving this under- standing even further. 4.14 XO-4 XO-4 was discovered to be a TEP by McCullough et al. (2008); the parent star is one of the hottest of the known planetary hosts. Here I analyse the R-band Perkins telescope light curve obtained by McCullough et al. (2008). Correlated errors are unimportant and the best solutions are LD fit/fix. The inclination is near 90◦, result- ing in asymmetric errorbars. My photometric solution (Fig. 18) is in excellent agreement with that of McCullough et al. (Table A61). The results of the ABSDIM analysis are given in Table A62, and include the first reported measurements of the T ′ eq, Safronov number, gb, ρA and ρb of the XO-4 system. The other output pa- rameters agree well with those of McCullough et al. (2008); Mb is 0.2 MJup smaller in my solution but this is within the errorbars. Improved photometric and RV observations of XO-4 would be ben- eficial. 4.15 XO-5 The discovery that XO-5 is a TEP system was presented by Burke et al. (2008) and extensive follow-up observations in the Figure 19. Phased light curves of XO-5 compared to the best fits found using JKTEBOP and the quadratic LD law. From top to bottom the light curves are Perkins R-band (Burke et al. 2008), and KeplerCam i-band and z-band (P´al et al. 2009). The residuals are offset from zero. context of the HAT consortium were presented by P´al et al. (2009). I adopt the KA value from P´al et al. (2009), which agrees with but is much more precise than that given by Burke et al. (2008). A com- parison of the Teff and(cid:2) Fe H(cid:3) given by the two studies shows a con- cerning disagreement, which may be due to the treatment of log gA. P´al et al. (2009) fix log gA at the transit-derived value during their spectral synthesis analysis, and also have spectra with a higher sig- nal to noise ratio, so I have preferred their spectroscopic results. Three high-quality light curves of XO-5 are available: R-band coverage of one full transit using the Perkins telescope (Burke et al. 2008), and i- and z-band observations (both covering one full and one partial transit) obtained with the FLWO 1.2 m telescope (P´al et al. 2009). I have analysed all three datasets, and the best fits are shown in Fig. 19. In the case of the R and z data correlated noise is important, and in all cases the best solutions are LD fit/fix. The resulting photometric parameters are in good agreement except for k, for which there is a large scatter of 5σ (Table A66). The three parameter sets have therefore been combined and the errorbar in k increased to account for this discrepancy. The physical properties of XO-5 are given in Table A67, and are in good agreement with established values. This system is a good candidate for improved photometric observations, which would allow to sort out the discrepancy in k and improve the preci- sion of the system properties. c(cid:13) 0000 RAS, MNRAS 000, 000–000 Homogeneous studies of transiting extrasolar planets. III. 17 4.16 TEPs studied in previous papers The preceding subsections have presented full studies of fifteen TEPs. In this subsection I apply my improved ABSDIM analysis to the other fifteen systems for which photometric results have already been calculated (Paper I; Southworth et al. 2009a,b,c, 2010). Their physical properties are collected and compared to literature results in Tables A68 to A81, and relevant points are discussed below. GJ 436. Attempts to obtain solutions for Teff = 3350 K failed. Inspection of Fig. 4 explains this: for a 0.5 M⊙ star we expect Teff values in the range 3500–4000 K (except for the Claret models which prefer 3100 K). I have therefore adopted the higher Teff of 3500 K from Bean et al. (2006), with an errorbar doubled to 100 K. The results with the Claret models are discrepant so were not incor- porated into the final solution. This does not mean that the Claret models are wrong – they are in fact closer to the measured prop- erties of eclipsing binaries (primarily CM Dra) in this mass regime – but that they depart from the consensus established by the other model sets. The lesson here is that we require an improved under- standing of low-mass stars to better measure the physical properties of the important GJ 436 system. The revised results for GJ 436 (Ta- ble A68) are slightly smaller than those found in Paper II, and are in good agreement with literature studies. HAT-P-1. I used a new KA value from Johnson et al. (2008). My ABSDIM solutions prefer the Teff value used in Paper II (Bakos et al. 2007a) to the higher one found by Ammler-von Eiff et al. (2009). HD 189733. A new KA value is available from Boisse et al. (2009), supported by the value given by Triaud et al. (2009). The ABSDIM solutions with different models converged on either a very young (∼1 Gyr) or old (9–13 Gyr) age, resulting in large system- atic errorbars for most output parameters. HD 189733 is an active star (Bouchy et al. 2005b) with starspots (Pont et al. 2007), a short rotation period of 12 d (Winn et al. 2007b), RV jitter (Boisse et al. 2009), and Ca H & K emission modulated on its rotation period (Moutou et al. 2007). These facts imply a young age (Skumanich 1972; Mamajek & Hillenbrand 2008) so the search for the best so- lution was restricted to ages below 5 Gyr. This resulted in much more consistent solutions, with ages of 0.1–2.9 Gyr, which were accepted as the final results (Table A70). HD 209458. I used the revised value of KA = 84.67 ± 0.70 given by TWH08. The revised results are in good agreement with those from Paper II. RA has been measured directly by interfer- ometric means (van Belle & von Braun 2009) to be 2σ larger than found here and in all other studies of this object given in Table A71. OGLE-TR-10. The light curve solution used in Paper II was unintentionally a preliminary rather than the final one from Paper I. This has been corrected here, resulting in a less dense planet. The properties of OGLE-TR-10 are rather uncertain and the system is badly in need of improved spectroscopy and photometry. OGLE-TR-132. I use the more precise KA value of 167 ± 18 km s−1 given by Moutou et al. (2004) which was overlooked in Papers I and II. OGLE-TR-132 b has a high T ′ eq of 2017 ± 97 K (Table 75). WASP-2. The results for this system are reproduced from the dedicated study of Southworth et al. (2010). WASP-4, WASP-5, WASP-18. I use the new and improved KA values for these three TEPs from Triaud et al. (2010). c(cid:13) 0000 RAS, MNRAS 000, 000–000 5 TRACKING THE SYSTEMATIC ERRORS IN THE PROPERTIES OF TRANSITING SYSTEMS A major result of the current work is a detailed understanding of where model-dependent systematic errors surface in the analysis of TEPs, and the importance of these systematics for the various physical properties which can be calculated. The approach used in this work means that all of the model dependence is combined into Kb (the velocity amplitude of the planet), making it an excellent tracer of systematic errors. For each TEP a separate value of Kb is found using each of the five sets of stellar models, as well as for the empirical mass–radius relation (Paper II). I have converted these into 'consensus values', hKbi, using the same algorithm as for the other measured physical properties: the unweighted mean of the values from the five different stellar model sets. Remember that hKbi should not be taken as an indicator of correctness, only of concordance. From Eqs. 4 to 12 in Paper II it can be seen that the component masses are most sensitive to systematic errors (MA ∝ K 3 b and Mb ∝ K 2 b ) and that their radii, gravities and densities are less so (all directly proportional to Kb). The semimajor axis is also rather model-dependent as the other quantities required to calculate it all have much smaller uncertainties than Kb does. A detailed exploration of model-dependence is presented in Fig. 20 as a function of Teff , (cid:2) Fe H(cid:3) and MA. The top panels in this Figure highlight the generally good agreement between different model sets. The mass–radius relation is in much poorer agreement and is biased to high values as it does not account for the effects of evolution through the main-sequence phase (particularly apparent for HAT-P-2, TrES-4 and XO-3). This bias to large Kb comes from trying to reproduce the low density of an evolved star, and pushes the mass and radius of the planet and star to high and incorrect values. Turning to the stellar model sets in Fig. 20, clear trends with models yielded larger Kb values on average, particularly for low respect to Teff ,(cid:2) Fe values of Teff and (cid:2) Fe H(cid:3) and MA can be seen in many cases. The Claret H(cid:3). This predilection for high values is bal- anced by other model sets: the Teramo and VRSS models both tend to produce lower Kb values and the DSEP models trend to low Kb for cooler and more metal-poor stars. Within this m´elange, the Y2 models are the closest to the consensus value and do not exhibit sig- nificant trends with the stellar properties. The Y2 models are thus the best choice to obtain quick results, although of course other model sets are required for the assessment of systematic errors. Figs. 21 and 22 compare the sizes of the random and system- atic errors in stellar mass, as a function of MA and of (cid:2) Fe spectively. MA was chosen for this comparison because it is one of the parameters more affected by systematics; it was also ex- pected that these systematic effects would be minimised in the re- H(cid:3), re- gion of 1 M⊙ and (cid:2) Fe H(cid:3) = 0 as all of the stellar model sets are calibrated on the Sun. Surprisingly, this does not turn out to be the case. There is a hard lower limit of 1% on the errors in MA arising from model-dependent systematics, but this lower limit is reached are clearly larger for lower-mass stars (<0.9 M⊙) and for a high at a wide range of masses and (cid:2) Fe metal abundance ((cid:2) Fe H(cid:3) values. The systematic errors H(cid:3) > 0.4). The hard lower limit is unavoid- able until stellar model sets are in much better agreement, and the mass–radius relation results hint that the real systematic errors are probably somewhat higher than this. The random errors in MA tend to decrease towards higher masses (temporarily ignoring the much fainter OGLE systems), as expected because more massive stars are intrinsically brighter and therefore easier to obtain good data for. 18 John Southworth Figure 20. Comparisons between the Kb values obtained using specific sets of stellar evolutionary models and the unweighted mean value, hKbi, for each TEP. From left to right the panels show results for the mass–radius relation and then the five stellar model sets. The top panels compare Kb to hKbi for each model set, with parity indicated by a dotted line. Lower panels show the difference (Kb − hKbi) as functions of effective temperature, metal abundance and stellar mass. Errorbars have been ignored for clarity; their median values are ±3 km s−1 (statistical) and ±1.4 km s−1 (systematic). 5.1 An external test: WASP-11 versus HAT-P-10 One of the best ways to investigate the presence of systematic errors is to compare two independent studies of the same object. Whilst multiple discovery and characterisation publications have been pre- sented for several TEPs (e.g. XO-1 and XO-5), successive papers on these objects have in each case been informed by the initial dis- covery papers. There are two exceptions: HD 80606 and WASP- 11 / HAT-P-10. In the case of HD 80606 three groups discovered its transiting nature essentially simultaneously, but their analyses were heavily dependent on the same published spectroscopic observa- tions. WASP-11 / HAT-P-10 is therefore the only TEP which was discovered and fully characterised by two groups working with- out knowledge of each others' analyses. The discovery papers were submitted within a week of each other, but the agreed name of the system is WASP-11 / HAT-P-10 because the WASP group submit- ted their paper first. In Table 4 I collect the properties of WASP-11 determined in the two discovery papers (West et al. 2009; Bakos et al. 2009). The values in general show a gratifying agreement, but two quantities stand out as being discrepant to some extent. The Teff values dis- agree by 1.6σ and the k values by 2.7σ. The divergent k values have a large knock-on effect on the planetary properties (Rb, gb and ρb). The possibility of systematic errors in Teff is a well-known phenomenon (see Sect. 3), so the 1.6σ disagreement is unsurpris- ing. Similarly, k measurements primarily depend on the observed transit depth (see Sect. 2) and can be affected by imperfect normali- sation as well as by astrophysical effects such as starspots. Discrep- ancies in k have previously been found in the well-studied systems HD 189733 and HD 209458 (Paper I), WASP-4 (Southworth et al. 2009b), HAT-P-1 and WASP-10 (Sections 4.1 and 4.11). This implies that a high-quality study of a TEP cannot rely on photometric coverage of only one transit, irrespective of its quality, but must include observations of two or preferably more in order to pick up on well-camouflaged systematic errors affecting light curves. It is important to realise that a systematic over- or under- estimation of a transit depth has a big effect on measurements of the planet properties, but cannot be identified in any dataset covering only one transit. c(cid:13) 0000 RAS, MNRAS 000, 000–000 Homogeneous studies of transiting extrasolar planets. III. 19 Table 5. Physical properties of the stellar components of the TEPs studied in this work. For each quantity the first uncertainty is derived from a propagation of all observational errors and the second uncertainty is an estimate of the systematic errors arising from the dependence on stellar theory. System Semimajor axis (AU) Mass ( M⊙) Radius ( R⊙) log gA [cm/s] Density ( ρ⊙) Age (Gyr) 4.92 ± 0.55 +0.00013 −0.00019 1.290 +0.120 −0.058 1.271 +0.043 −0.024 +0.012 −0.017 0.592+0.083 −0.129 1.98 ± 0.17 unconstrained 0.02887 ± 0.00089 ± 0.00035 0.459 ± 0.043 ± 0.017 0.454 ± 0.029 ± 0.005 4.787 ± 0.030 ± 0.005 GJ 436 +0.5 0.05535 ± 0.00057 ± 0.00042 1.134 ± 0.035 ± 0.026 1.112 ± 0.031 ± 0.008 4.400 ± 0.024 ± 0.003 0.824 ± 0.066 2.1 +1.4 HAT-P-1 −0.6 −1.2 +0.5 4.092 ± 0.074 ± 0.002 0.268 ± 0.070 2.6 +0.4 0.06740 ± 0.00074 ± 0.00034 1.279 ± 0.042 ± 0.019 1.68 ± 0.15 ± 0.01 HAT-P-2 −0.3 −0.7 +0.3 1.2 +1.1 4.321 +0.042 +0.001 +0.004 0.04288 +0.00048 HD 149026 −0.1 −1.5 −0.070 −0.002 −0.006 −0.00027 +1.5 1.4 +4.7 0.03142 ± 0.00038 ± 0.00036 0.840 ± 0.030 ± 0.029 0.752 ± 0.023 ± 0.009 4.610 ± 0.026 ± 0.005 HD 189733 −1.3 −1.4 +0.5 0.04747 ± 0.00046 ± 0.00031 1.148 ± 0.033 ± 0.022 1.162 ± 0.012 ± 0.008 4.368 ± 0.005 ± 0.003 0.733 ± 0.008 2.3 +0.9 HD 209458 −0.4 −0.7 +0.3 4.178 ± 0.053 ± 0.001 0.361 ± 0.063 3.1 +3.3 0.04516 ± 0.00099 ± 0.00015 1.277 ± 0.082 ± 0.013 1.52 ± 0.10 ± 0.00 OGLE-TR-10 −0.2 −0.7 1.7 +1.4 +0.3 OGLE-TR-56 4.331 ± 0.094 ± 0.002 0.02386 ± 0.00028 ± 0.00009 1.233 ± 0.043 ± 0.014 1.26 ± 0.14 ± 0.00 −0.2 −2.0 unconstrained OGLE-TR-111 0.04651 ± 0.00099 ± 0.00051 0.833 ± 0.054 ± 0.027 0.842 ± 0.042 ± 0.009 4.508 ± 0.044 ± 0.005 unconstrained OGLE-TR-113 0.02278 ± 0.00047 ± 0.00022 0.768 ± 0.048 ± 0.022 0.780 ± 0.029 ± 0.008 4.539 ± 0.028 ± 0.004 1.5 +4.2 +0.3 4.275 ± 0.083 ± 0.003 OGLE-TR-132 0.03029 ± 0.00062 ± 0.00018 1.297 ± 0.078 ± 0.023 1.37 ± 0.14 ± 0.01 −0.4 −1.5 4.3 +0.5 +1.4 OGLE-TR-182 0.05205 ± 0.00057 ± 0.00031 1.187 ± 0.039 ± 0.021 1.53 ± 0.17 ± 0.01 4.142 ± 0.089 ± 0.003 −1.3 −1.9 +0.4 2.6 +0.6 +0.002 4.170 +0.052 +0.01 OGLE-TR-211 0.05105 +0.00076 +0.00020 −0.3 −0.7 −0.085 −0.002 −0.01 −0.00073 −0.00021 +0.3 1.503 ± 0.083 ± 0.008 4.236 ± 0.043 ± 0.002 0.418 ± 0.061 1.0 +0.6 0.0404 ± 0.0011 ± 0.0002 OGLE-TR-L9 −0.2 −0.7 0.03946 ± 0.00039 ± 0.00060 0.892 ± 0.026 ± 0.041 0.818 ± 0.017 ± 0.013 4.563 ± 0.019 ± 0.007 1.632 ± 0.092 3.4 +3.4 +1.9 TrES-1 −2.9 −3.0 +0.7 0.03635 ± 0.00063 ± 0.00035 1.049 ± 0.054 ± 0.030 1.002 ± 0.029 ± 0.010 4.457 ± 0.027 ± 0.004 1.043 ± 0.088 2.5 +2.8 TrES-2 −0.8 −2.5 4.581 +0.007 0.1 +0.7 0.02283 +0.00012 +0.0 +0.004 TrES-3 −0.0 −0.0 −0.00025 −0.010 −0.006 3.981 ± 0.047 ± 0.002 0.182 ± 0.030 3.7 +1.6 +0.2 0.04965 ± 0.00087 ± 0.00028 1.292 ± 0.067 ± 0.022 1.92 ± 0.11 ± 0.01 TrES-4 −0.4 −1.4 +0.3 3.0 +0.7 4.207 +0.045 +0.005 0.03898 +0.00036 WASP-1 −0.3 −0.6 −0.028 −0.005 −0.00039 0.03033 ± 0.00060 ± 0.00043 0.803 ± 0.049 ± 0.034 0.807 ± 0.019 ± 0.011 4.529 ± 0.017 ± 0.006 1.527 ± 0.067 11.9 +8.1 +3.3 WASP-2 −2.5 −4.3 +0.4 1.377 ± 0.085 ± 0.009 4.262 ± 0.044 ± 0.003 0.484 ± 0.073 2.1 +1.5 WASP-3 0.03187 ± 0.00086 ± 0.00020 1.26 ± 0.10 ± 0.02 −1.2 −0.3 +2.1 7.0 +5.2 0.905 +0.021 +0.031 0.02307 +0.00053 WASP-4 −1.8 −4.5 −0.022 −0.026 −0.00055 0.02714 ± 0.00049 ± 0.00022 1.004 ± 0.055 ± 0.025 1.077 ± 0.042 ± 0.009 4.375 ± 0.030 ± 0.004 0.803 ± 0.080 7.0 +3.2 +1.5 WASP-5 −3.0 −1.5 unconstrained 0.0378 ± 0.0014 ± 0.0008 WASP-10 +0.6 0.02034 ± 0.00032 ± 0.00016 1.256 ± 0.059 ± 0.029 1.222 ± 0.042 ± 0.010 4.363 ± 0.027 ± 0.003 0.689 ± 0.062 0.5 +1.2 WASP-18 −0.4 −0.9 0.04944 ± 0.00062 ± 0.00050 1.037 ± 0.039 ± 0.031 0.942 ± 0.022 ± 0.009 4.506 ± 0.021 ± 0.004 1.242 ± 0.080 0.9 +2.4 +0.8 XO-1 −0.8 −0.9 +4.7 0.03647 +0.00059 1.9 +4.5 XO-2 −1.9 −1.9 −0.00058 0.04529 ± 0.00057 ± 0.00045 1.206 ± 0.046 ± 0.036 1.409 ± 0.054 ± 0.014 4.222 ± 0.027 ± 0.004 0.431 ± 0.041 3.0 +0.9 +0.5 XO-3 −0.4 −0.6 2.7 +0.6 +0.2 0.05475 +0.00094 +0.00022 XO-4 −0.5 −0.3 −0.00051 −0.00035 0.0494 ± 0.0011 ± 0.0002 XO-5 unconstrained +0.002 1.285 +0.068 −0.002 −0.036 0.914 ± 0.064 ± 0.010 1.065 ± 0.064 ± 0.004 4.344 ± 0.043 ± 0.002 0.62 ± 0.21 1.40 ± 0.19 1.62 ± 0.13 0.50 ± 0.15 0.33 ± 0.10 0.345+0.068 −0.090 0.752 ± 0.081 ± 0.048 0.703 ± 0.036 ± 0.015 4.620 ± 0.049 ± 0.009 1.312 +0.059 +0.015 −0.056 −0.016 1.42 ± 0.11 ± 0.02 0.359+0.046 −0.160 0.76 ± 0.11 0.929 +0.014 −0.030 +0.014 −0.021 0.970 +0.028 −0.035 +0.018 −0.022 1.530 +0.362 −0.069 +0.006 −0.010 +0.010 −0.009 4.485 +0.011 −0.012 +0.005 −0.004 1.243 +0.034 −0.037 +0.013 −0.014 1.455 +0.052 −0.079 +0.00069 −0.00081 0.946 +0.046 −0.046 +0.054 −0.062 +0.00026 −0.00022 0.914 +0.064 −0.066 4.440 +0.037 −0.021 +0.008 −0.010 1.037+0.128 −0.058 +0.002 −0.002 0.403+0.069 −0.037 +0.002 −0.003 1.700+0.047 −0.051 +0.00012 −0.00017 +0.00013 −0.00014 1.233+0.020 −0.022 2.16 ± 0.31 +0.016 −0.024 4.178 +0.034 −0.169 1.56 +0.18 −0.10 0.818 +0.011 −0.013 6 PHYSICAL PROPERTIES OF THE TRANSITING EXTRASOLAR PLANETARY SYSTEMS The main result of this work is the determination of the basic physi- cal properties of thirty TEPs using homogeneous methods. Detailed results for each TEP are given in the many Tables in the Appendix, and the final results for all systems have been summarised in Tables 5 and 6. The homogeneous nature of these numbers means they are well suited for comparing different TEPs, for planning follow-up observations, and for performing detailed statistical studies. The masses and radii of the stars considered in this work are plotted in Fig. 23 with both their random and systematic errorbars, and with the empirical mass–radius relation (Paper II) indicated. A striking aspect of Fig. 23 is that for MA < 1.2 M⊙ the stars gen- erally stick close to the main sequence, but beyond this point their distribution turns almost vertically upwards. The two exceptions are XO-5, which has a more evolved lower-mass star, and OGLE- TR-L9, which contains a more massive star (1.4 M⊙). That low-mass stars stick closely to the main sequence is to be expected due to their long evolutionary timescales. The shift to larger RA at 1.2 M⊙ is reasonable because a larger stellar radius makes it more likely that a given planet is transiting (see Paper II). The avoidance of more massive stars is likely due to difficulty of measuring RVs for these objects, making them lower-priority and more observationally expensive targets for TEP search consortia at Figure 21. Plot of the sizes of the statistical (red filled circles) and system- atic (blue crosses) errorbars for the TEPs studied in this work, versus the stellar masses. The errors are plotted as fractions of MA and the statistical and systematic errors for each TEP are connected by grey lines. c(cid:13) 0000 RAS, MNRAS 000, 000–000 20 John Southworth Table 6. Physical properties of the planetary components of the TEPs studied in this work. For each quantity the first uncertainty is derived from a propagation of all observational errors and the second uncertainty is an estimate of the systematic errors arising from the dependence on stellar theory. System Mass ( MJup) Radius ( RJup) gb ( m s−2) Density ( ρJup) eq (K) T ′ Θ +0.01 −0.01 0.75 +0.15 −0.15 0.356 +0.013 −0.011 0.0737 ± 0.0051 ± 0.0018 GJ 436 0.524 ± 0.016 ± 0.008 HAT-P-1 8.74 ± 0.25 ± 0.09 HAT-P-2 +0.002 HD 149026 −0.003 1.150 ± 0.028 ± 0.027 HD 189733 0.714 ± 0.014 ± 0.009 HD 209458 0.68 ± 0.15 ± 0.00 OGLE-TR-10 OGLE-TR-56 1.300 ± 0.080 ± 0.010 OGLE-TR-111 0.54 ± 0.10 ± 0.01 OGLE-TR-113 1.24 ± 0.17 ± 0.02 OGLE-TR-132 1.17 ± 0.14 ± 0.01 OGLE-TR-182 1.06 ± 0.15 ± 0.01 OGLE-TR-211 OGLE-TR-L9 TrES-1 TrES-2 TrES-3 TrES-4 WASP-1 WASP-2 WASP-3 WASP-4 WASP-5 WASP-10 WASP-18 XO-1 XO-2 XO-3 XO-4 XO-5 +0.006 −0.006 0.847 ± 0.038 ± 0.024 2.06 ± 0.13 ± 0.03 +0.028 −0.023 1.565 ± 0.058 ± 0.026 3.16 ± 0.23 ± 0.13 10.29 ± 0.33 ± 0.16 0.924 ± 0.075 ± 0.019 +0.021 −0.025 11.83 ± 0.31 ± 0.23 +0.012 −0.019 1.084 ± 0.055 ± 0.008 4.34 ± 1.47 ± 0.05 0.761 ± 0.045 ± 0.023 1.253 ± 0.047 ± 0.024 +0.020 −0.029 0.877 ± 0.072 ± 0.010 1.237 +0.059 −0.062 0.860 +0.072 −0.072 1.910 +0.060 −0.070 0.555 +0.062 −0.057 1.521 +0.160 −0.153 1.57 +0.72 −0.57 13.7 ± 1.1 8.77 ± 0.56 152 ± 30 23.7+6.8 −6.2 21.5 ± 1.2 9.30 ± 0.08 5.7 ± 1.4 22.3 ± 6.7 11.5 ± 2.5 25.0 ± 3.7 18.5 ± 5.0 12.1 ± 2.9 11.6+2.9 −3.3 41 ± 14 1.51 ± 0.19 ± 0.02 0.290 ± 0.027 ± 0.002 5.1 ± 1.5 ± 0.0 +0.01 −0.01 0.365 ± 0.018 ± 0.004 1.217 ± 0.038 ± 0.009 1.19 ± 0.12 ± 0.01 0.610 +0.099 +0.002 −0.072 −0.003 1.151 ± 0.036 ± 0.013 0.755 ± 0.066 ± 0.009 1.380 ± 0.015 ± 0.009 0.272 ± 0.004 ± 0.002 1.72 ± 0.11 ± 0.01 0.134 ± 0.038 ± 0.000 1.20 ± 0.17 ± 0.00 0.75 ± 0.34 ± 0.34 1.077 ± 0.072 ± 0.012 0.43 ± 0.11 ± 0.00 1.111 ± 0.049 ± 0.011 0.91 ± 0.16 ± 0.01 1.25 ± 0.16 ± 0.01 0.59 ± 0.24 ± 0.00 1.47 ± 0.14 ± 0.01 0.332 ± 0.111 ± 0.002 1.262 +0.158 +0.005 0.372 +0.117 +0.002 −0.005 −0.091 −0.001 −0.132 1.614 ± 0.083 ± 0.009 1.03 ± 0.37 ± 0.01 1.099 ± 0.031 ± 0.017 0.573 ± 0.056 ± 0.009 1.261 ± 0.039 ± 0.012 0.625 ± 0.051 ± 0.006 +0.007 0.860 +0.050 +0.007 1.305 +0.027 −0.057 −0.004 −0.025 −0.010 1.81 ± 0.15 ± 0.01 0.148 ± 0.039 ± 0.001 1.484 +0.059 +0.005 +0.001 0.263 +0.058 −0.091 −0.006 −0.001 −0.036 1.043 ± 0.029 ± 0.015 19.32 ± 0.80 0.747 ± 0.047 ± 0.010 0.67 ± 0.11 ± 0.00 1.454 ± 0.083 ± 0.009 +0.005 +0.015 1.357 +0.033 0.495 +0.015 −0.018 −0.033 −0.013 −0.006 0.99 ± 0.14 ± 0.01 1.164 ± 0.056 ± 0.009 2.60 ± 0.39 ± 0.05 1.067 ± 0.064 ± 0.022 6.64 ± 0.90 ± 0.05 1.158 ± 0.054 ± 0.009 0.526 ± 0.063 ± 0.005 1.206 ± 0.039 ± 0.012 +0.013 0.569 +0.119 +0.019 0.992 +0.034 −0.072 −0.011 −0.057 −0.022 6.08 ± 0.67 ± 0.06 1.248 ± 0.047 ± 0.012 +0.01 −0.00 0.84 ± 0.18 ± 0.00 24.2 ± 2.6 16.65+0.26 −0.33 28.7 ± 2.6 68.9 ± 6.7 190 ± 16 15.8 ± 1.5 14.0+2.1 −1.5 188 ± 13 22.8+3.2 −9.5 22.7 ± 3.2 15.6 ± 1.2 19.5 ± 1.1 27.8+1.2 −1.4 6.7 ± 1.2 9.7+1.5 −1.1 1.089 ± 0.082 ± 0.004 0.71 +0.13 −0.40 1.29 +0.38 −0.06 +0.01 −0.01 0.0253 ± 0.0015 ± 0.0003 669 ± 22 0.0419 ± 0.0016 ± 0.0003 1291 ± 20 0.771 ± 0.077 ± 0.004 1516 ± 66 0.0393 +0.0054 +0.0002 1626+69 −0.0056 −0.0001 −37 0.0747 ± 0.0024 ± 0.0009 1191 ± 20 0.0427 ± 0.0005 ± 0.0003 1459 ± 12 0.0279 ± 0.0062 ± 0.0001 1702 ± 54 2140 ± 120 0.0418 ± 0.0065 ± 0.0002 0.056 ± 0.011 ± 0.001 1034 ± 28 0.0664 ± 0.0090 ± 0.0007 1355 ± 35 0.0436 ± 0.0072 ± 0.0003 2017 ± 97 0.0628 ± 0.0109 ± 0.0004 1550 ± 81 0.0460 +0.0097 +0.0002 1686+90 −0.0002 −0.0104 −55 0.153 ± 0.052 ± 0.001 2039 ± 51 0.0612 ± 0.0037 ± 0.0009 1147 ± 15 0.0688 ± 0.0024 ± 0.0007 1467 ± 27 +0.0006 0.0719 +0.0026 1630+23 −0.0026 −0.0004 −22 0.0373 ± 0.0042 ± 0.0002 1861 ± 54 0.0363 +0.0038 +0.0001 1800+32 −0.0033 −49 −0.0001 0.0613 ± 0.0021 ± 0.0009 1281 ± 21 0.0715 ± 0.0048 ± 0.0004 2028 ± 59 +0.0004 1661+30 0.0460 +0.0012 −0.0013 −30 −0.0005 0.0726 ± 0.0035 ± 0.0006 1732 ± 41 0.298 ± 0.019 ± 0.006 972 ± 31 0.288 ± 0.014 ± 0.002 2392 ± 51 0.0730 ± 0.0062 ± 0.0007 1210 ± 16 +0.0010 0.0432 +0.0049 1328+17 −0.0045 −0.0008 −28 0.711 ± 0.027 ± 0.007 1729 ± 34 0.1006 +0.0112 1630+169 +0.0006 −0.0253 −36 −0.0004 0.1075 ± 0.0082 ± 0.0004 1203 ± 33 Table 4. Physical properties of the WASP-11 / HAT-P-14 system deter- mined by the two discovery papers. Aside from the orbital period, quantities without uncertainties were calculated from the results given in the respec- tive papers. Parameter West et al. (2009) Bakos et al. (2009) Porb (d) Teff (K) Vsini ( km s−1) H(cid:3) (dex) H(cid:3),(cid:2) Fe (cid:2) M KA ( m s−1) e a (AU) rA rb k i (◦) MA ( M⊙) RA ( R⊙) log gA [cgs] Mb ( MJup) Rb ( RJup) gb ( m s−2) ρb ( ρJup) eq (K) T ′ 3.722465 4800 ± 100 < 6 0.0 ± 0.2 82.1 ± 7.4 0 fixed 0.043 ± 0.002 0.0801 0.01011 0.1273+0.0011 −0.0008 89.8+0.2 −0.8 0.77+0.10 −0.08 0.74+0.04 −0.03 4.45 ± 0.2 0.53 ± 0.07 0.91+0.06 −0.03 14.5+1.4 −1.6 0.69+0.07 −0.11 960 ± 70 3.7224747 4980 ± 60 0.5 ± 0.2 0.13 ± 0.08 74.5 ± 1.8 0 fixed 0.0435 ± 0.0006 0.0842 ± 0.0019 0.01104 0.1313 ± 0.0010 88.6+0.5 −0.4 0.83 ± 0.03 0.79 ± 0.02 4.56 ± 0.02 0.460 ± 0.028 1.005+0.032 −0.027 12.0 ± 0.8 0.479 ± 0.042 1020 ± 17 Figure 22. Same as Fig. 21, except that the error sizes are plotted versus stellar(cid:2) Fe H(cid:3) rather than mass. the follow-up stage. There may be a large and presently neglected population of TEPs around more massive stars11. 11 Whilst RV surveys mostly concentrate on F, G and K stars, they have c(cid:13) 0000 RAS, MNRAS 000, 000–000 Homogeneous studies of transiting extrasolar planets. III. 21 Figure 23. Plot of the masses versus the radii of the stars in the thirty TEPs studied in this work. The statistical uncertainties are shown by black open diamonds and the systematic uncertainties by red filled diamonds. The em- pirical mass–radius relation from Paper I is shown with a blue line. Figure 25. Mass–radius plot for the known transiting extrasolar planets. Those objects studied in this work are shown with (blue) filled circles and numbers taken from the literature with (red) crosses. The four gas giant planets in our Solar System are denoted with (black) filled circles. Dotted lines show loci where density equals 1.0, 0.5 and 0.25 ρJup. rors are much less important for these objects. The two outliers with very low densities are OGLE-TR-10 and TrES-4. 6.1 Physical properties of all known TEPs The physical properties of the thirty TEPs studied in this work have been augmented with the literature values for the other 41 known systems (as of 31/05/2010) to yield a larger but inhomogeneous sample. Fig. 25 shows their masses and radii on logarithmic axes, alongside those of the four Solar system gas giants. The dominant population of known TEPs continues to reside in a cloud of points centred roughly on 1.0 MJup and 1.2 RJup. There is a clear tail of high-mass planets, culminating in the brown dwarf CoRoT-3, as well as an increasing number of systems of much smaller mass and radius, typified by GJ 1214 and the first possible rocky exoplanet, CoRoT-7. Correlations have previously been noticed between Porb and gb (Southworth et al. 2007b) and Porb and Mb (Mazeh et al. 2005). The relevant plots are shown in Figs. 26 and 27. In both cases there are ten planets whose properties put them outside the range of Porb shown in these Figures. Neglecting this population of mas- sive planets (which may be a fundamentally different population to the lower mass ones; Fabrycky & Winn 2009; Southworth et al. 2009c), the rank correlation test of Spearman (1904) returns a prob- ability of 99.80% (3.1σ) that the Porb–gb correlation is real and 97.6% (2.3σ) that the Porb–Mb correlation is real. The correspond- ing figures from Paper II, based on 44 TEPs, are 98.3% and 97.5%. Two concerns with these correlations are apparent. Firstly, they are insignificant if the longer-period planets are not rejected12. Sec- ondly, the period–mass correlation has not increased in significance with the addition of 21 new TEPs since Paper II. Hansen & Barman (2007) divided up eighteen of the twenty Figure 24. Plot of the masses versus the radii of the planets in the thirty TEPs studied in this work. The statistical uncertainties are shown by black open diamonds and the systematic uncertainties by red filled diamonds. Blue dotted lines show where density is 1.0, 0.5 and 0.25 ρJup. Another notable feature of Fig. 23 is that the systematic errors (red filled diamonds in the Figure) are not negligible compared to the random errors, and in some cases (HD 189733, TrES-1, TrES-3, XO-2) are of a similar size or even larger. A similar diagram for the planets (Fig. 24) shows a much larger scatter in the planet properties, but also that systematic er- not been neglecting ones more massive than this; see Galland et al. (2005), Johnson et al. (2007) Lagrange et al. (2009) and Bowler et al. (2010). 12 Merrill's theorem should be borne in mind: "When one throws away discrepant results, the remainder are found to agree well." c(cid:13) 0000 RAS, MNRAS 000, 000–000 22 John Southworth Figure 26. Plot of the orbital periods versus the surface gravities of known TEPs. The objects studied in this work are shown with (blue) filled circles and other objects with (red) open circles. Figure 27. Same as Fig. 26, but for planetary mass instead of gb. TEPs then known into two classes based on their position in a di- agram of Θ versus Teq. An updated version of the diagram can be seen in Fig. 28, and agrees with previous conclusions (Paper II) that the division between the classes was blurred into insignifi- cance. A dotted line at Θ = 0.055 has been drawn to show the ex- pected boundaries between Class I (Θ ≈ 0.07 ± 0.01) and Class II (Θ ≈ 0.04 ± 0.01). There are several other equally possible lines which could be drawn through this diagram, which does not in- spire confidence in their reality. It is probable that the addition of new TEPs in the future will fill out this diagram with objects drawn from a single distribution. Figure 28. Plot of equilibrium temperature versus Safronov number for the full sample of planets. Objects shown with (blue) circles were studied in this work; those which are just (red) errorbars were not. The dotted line shows the separation between Class A and Class B proposed by Hansen & Barman (2007). 7 SUMMARY Measurements of the physical properties of many transiting extra- solar planets are available in the literature, but are not directly com- parable as they have been derived by different researchers using a variety of methods. In this series of papers I aim to derive the physical properties of the known TEPs using a rigorous and ho- mogeneous approach, resulting in quantities which are suitable for statistical analysis. Each TEP is tackled in two steps, the first being analysis of its transit light curves and the second being the determi- nation of its physical properties. The transit light curves are studied using the JKTEBOP code, which models binary systems using biaxial spheroids. I pay spe- cial attention to the limb darkening of the parent star; solutions are obtained for every light curve using each of five different limb dark- ening laws. Orbital eccentricity is accounted for by including e and ω as fitted parameters constrained by values and uncertainties mea- sured from radial velocity observations of the parent star. Parame- ter errors are assessed using Monte Carlo simulations, a residual- permutation algorithm, and the inter-agreement between solutions with different limb darkening laws. In this paper I present analyses of published transit light curves of fifteen TEPs (Table 2). Three of these (TrES-2, TrES-4 and WASP-2) have faint companion stars within 1.6′′ on the sky (Daemgen et al. 2009). If neglected, this 'third light' dilutes the transit depth and causes systematic errors in parameters measured from the light curves. As a guideline, L3 = 5% causes the radius of a representative planet to be underestimated by 2%. I show that it is not possible to detect third light in synthetic light curves represen- tative of good ground-based observations. A third light value must instead be measured and used to constrain the light curve solutions. In the second stage of the analysis of each TEP, I augment the light curve results with measurements of the velocity amplitude, Teff and (cid:2) Fe H(cid:3) for its parent star. I then interpolate within tabu- lated predictions from theoretical stellar models to find the over- all best-fitting physical properties of the star and the planet. This c(cid:13) 0000 RAS, MNRAS 000, 000–000 Homogeneous studies of transiting extrasolar planets. III. 23 Table 7. Summary of which types of additional observations would be use- ful for the thirty TEPs studied in this series of papers. ⋆ denotes where additional data would be useful, and ⋆⋆ indicates where it would be useful but difficult to either obtain or interpret. System Photometric observations Radial velocities Spectral synthesis GJ 436 HAT-P-1 HAT-P-2 HD 149026 HD 189733 HD 209458 OGLE-TR-10 OGLE-TR-56 OGLE-TR-111 OGLE-TR-113 OGLE-TR-132 OGLE-TR-182 OGLE-TR-211 OGLE-TR-L9 TrES-1 TrES-2 TrES-3 TrES-4 WASP-1 WASP-2 WASP-3 WASP-4 WASP-5 WASP-10 WASP-18 XO-1 XO-2 XO-3 XO-4 XO-5 ⋆ ⋆⋆ ⋆ ⋆⋆ ⋆⋆ ⋆⋆ ⋆⋆ ⋆⋆ ⋆⋆ ⋆ ⋆ ⋆ ⋆ ⋆ ⋆ ⋆ ⋆ ⋆ ⋆ ⋆⋆ ⋆ ⋆ ⋆ ⋆ ⋆⋆ ⋆⋆ ⋆⋆ ⋆⋆ ⋆⋆ ⋆⋆ ⋆⋆ ⋆⋆ ⋆ ⋆ ⋆ ⋆ ⋆ ⋆ ⋆ model dependence causes systematic errors in the resulting quanti- ties, which I assess by comparing the individual results found using five independent sets of theoretical model predictions. This leads also to 'consensus values' for the physical properties; the Yonsei- Yale models are overall the closest to this consensus. Measurements of three physical properties are exempt from this model dependence: the planetary surface gravity can be deter- mined from only observed quantities; the density of the star has only a negligible dependence on stellar models; and the planet's equilibrium temperature does not depend on stellar models but does rest on the effective temperature scale of low-mass (F, G and K) stars. The model dependence specifies a minimum level of uncer- tainty for other physical properties, and this ranges from 1% for MA to 0.6% for Mb and 0.3% for RA, Rb and a. Reducing these minimum levels will require improvements in our understanding of low-mass stars. An external test of systematic errors was ob- tained by comparing the discovery papers of the WASP-11 / HAT- P-10 system. The agreement between the two studies is good for most parameters, but less so for the transit depth and the stellar Teff . Tables 5 and 6 summarise the physical properties of thirty TEPs: the fifteen with light curve analyses in the current work plus another fifteen with photometric analyses in Paper I and later works. The statistical errors in these quantities are calculated using a perturbation analysis which returns complete error budgets for c(cid:13) 0000 RAS, MNRAS 000, 000–000 each output parameter. These error budgets indicate which type of follow-up observations are the best way of improving the parameter measurements for each TEP. In most the quality of the light curve is the primary limitation on measurements of the properties of the planet; many systems would also benefit from additional spectro- scopic observations. This is summarised in Table 7 for the benefit of follow-up programmes. The measured properties of the 30 TEPs were then reinforced with those of the other 35 known objects and plotted in several pa- rameter diagrams. I find that the correlation between orbital period and planetary surface gravity is significant at the 3.1σ level once clear outliers are thrown out – this significance has increased with the addition of new systems since Paper II. The period vs. plane- tary mass correlation has a significance of only 2.3σ, and this not increased since Paper II. Similarly, the distinction between Class I and Class II planets in the T ′ eq vs. Θ diagram is blurring into irrel- evance. In this work I have now treated 30 TEPs, including all cur- rently known from the OGLE, TrES and XO surveys. The discov- ery rate of transiting planets is expected to continue rising, as the ground-based surveys HAT and WASP mature and new areas of pa- rameter space are probed by the space missions CoRoT and Kepler. A homogeneous study of the atmospheric parameters of all TEP host stars would be very complementary to this work. Whilst the light curves obtained by CoRoT and Kepler are of extremely high quality, it should not be forgotten that extensive velocity observa- tions are also needed to study transiting systems in detail. Precise physical properties of transiting planets are a fundamental require- ment of constraining the atmospheric characteristics, formation and evolution of planetary systems. ACKNOWLEDGMENTS I acknowledge financial support from STFC in the form of a post- doctoral research assistant position, and a timely and useful report from the anonymous referee. I am grateful to Barry Smalley, An- drew Collier Cameron, Pierre Maxted and Aaron Dotter for discus- sions, and to Adriano Pietrinferni, Emma Nasi and Antonio Claret for calculating stellar model sets for me. I thank Ignas Snellen, G´asp´ar Bakos, Neale Gibson and Peter McCullough for sending me their data and to the NSTeD and CDS websites for making many other datasets available. The following internet-based resources were used in research for this paper: the ESO Digitized Sky Survey; the NASA Astrophysics Data System; the SIMBAD database oper- ated at CDS, Strasbourg, France; the arχiv scientific paper preprint service operated by Cornell University; and the BaSTI web tools. REFERENCES Alonso, R., et al., 2004, ApJ, 613, L153 Ammler-von Eiff, M., Santos, N. C., Sousa, S. G., Fernandes, J., Guillot, T., Israelian, G., Mayor, M., Melo, C., 2009, A&A, 507, 523 Anderson, D. R., et al., 2008, MNRAS, 387, L4 Bakos, G., Noyes, R. W., Kov´acs, G., Stanek, K. Z., Sasselov, D. D., Domsa, I., 2004, PASP, 116, 266 Bakos, G. ´A., L´az´ar, J., Papp, I., S´ari, P., Green, E. M., 2002, PASP, 114, 974 Bakos, G. ´A., et al., 2007a, ApJ, 656, 552 Bakos, G. ´A., et al., 2007b, ApJ, 670, 826 24 John Southworth Bakos, G. ´A., et al., 2009, apJ, 696, 1950 Ballard, S., et al., 2009, in Transiting planets, vol. 253 of IAU Symposium, p. 470 Bean, J. L., Benedict, G. F., Endl, M., 2006, ApJ, 653, L65 Boisse, I., et al., 2009, A&A, 495, 959 Bonfils, X., Delfosse, X., Udry, S., Santos, N. C., Forveille, T., S´egransan, D., 2005, A&A, 442, 635 Bouchy, F., Pont, F., Santos, N. C., Melo, C., Mayor, M., Queloz, D., Udry, S., 2004, A&A, 421, L13 Bouchy, F., Pont, F., Melo, C., Santos, N. C., Mayor, M., Queloz, D., Udry, S., 2005a, A&A, 431, 1105 Bouchy, F., et al., 2005b, A&A, 444, L15 Bowler, B. P., et al., 2010, ApJ, 709, 396 Bruntt, H., Southworth, J., Torres, G., Penny, A. J., Clausen, J. V., Buzasi, D. L., 2006, A&A, 456, 651 Burke, C. J., et al., 2007, ApJ, 671, 2115 Burke, C. J., et al., 2008, ApJ, 686, 1331 Christian, D. J., et al., 2009, MNRAS, 392, 1585 Claret, A., 2000, A&A, 363, 1081 Claret, A., 2004a, A&A, 428, 1001 Claret, A., 2004b, A&A, 424, 919 Claret, A., 2005, A&A, 440, 647 Claret, A., 2006, A&A, 453, 769 Claret, A., 2007, A&A, 467, 1389 Claret, A., Hauschildt, P. H., 2003, A&A, 412, 241 Collier Cameron, A., et al., 2007, MNRAS, 380, 1230 Daemgen, S., Hormuth, F., Brandner, W., Bergfors, C., Janson, M., Hippler, S., Henning, T., 2009, A&A, 498, 567 Damasso, M., Calcidese, P., Bernagozzi, A., Bertolini, E., Gi- acobbe, P., Lattanzi, M. G., Smart, R., Sozzetti, A., 2009, in "Pathways towards habitable planets" conference, in press, arXiv:0911.3587 Demarque, P., Woo, J.-H., Kim, Y.-C., Yi, S. K., 2004, ApJS, 155, 667 D´ıaz, R. F., et al., 2007, ApJ, 660, 850 Dittmann, J. A., Close, L. M., Scuderi, L. J., Morris, M. D., 2010, ApJ, submitted, arXiv:1003.1762 Dotter, A., Chaboyer, B., Jevremovi´c, D., Kostov, V., Baron, E., Ferguson, J. W., 2008, ApJS, 178, 89 Etzel, P. B., 1981, in Carling, E. B., Kopal, Z., eds., Photomet- ric and Spectroscopic Binary Systems, NATO ASI Ser. C., 69, Kluwer, Dordrecht, p. 111 Fabrycky, D. C., Winn, J. N., 2009, ApJ, 696, 1230 Fernandez, J. M., Holman, M. J., Winn, J. N., Torres, G., Shporer, A., Mazeh, T., Esquerdo, G. A., Everett, M. E., 2009, AJ, 137, 4911 Fortney, J. J., Lodders, K., Marley, M. S., Freedman, R. S., 2008, ApJ, 678, 1419 Galland, F., Lagrange, A., Udry, S., Chelli, A., Pepe, F., Queloz, D., Beuzit, J., Mayor, M., 2005, A&A, 443, 337 Ghezzi, L., Cunha, K., de Ara´ujo, F. X., Smith, V. V., de la Reza, R., Schuler, S., 2010, in K. Cunha, M. Spite, & B. Barbuy, ed., IAU Symposium, vol. 265, p. 432 Gibson, N. P., et al., 2008, A&A, 492, 603 Gibson, N. P., et al., 2009, ApJ, 700, 1078 et Gilliland, R. L., arXiv:1001.0142 Gillon, M., Pont, F., Moutou, C., Bouchy, F., Courbin, F., Sohy, S., Magain, P., 2006, A&A, 459, 249 Gillon, M., et al., 2007, A&A, 466, 743 Gillon, M., et al., 2009, A&A, 496, 259 Greiner, J., et al., 2008, PASP, 120, 405 al., 2010, ApJ Letters, in press, Hansen, B. M. S., Barman, T., 2007, ApJ, 671, 861 H´ebrard, G., et al., 2008, A&A, 488, 763 Hellier, C., et al., 2009, Nature, 460, 1098 Hilditch, R. W., 2001, An Introduction to Close Binary Stars, Cambridge University Press, Cambridge, UK Holman, M. J., et al., 2007, ApJ, 664, 1185 Jenkins, J. M., Caldwell, D. A., Borucki, W. J., 2002, ApJ, 564, 495 Johns-Krull, C. M., et al., 2008, ApJ, 677, 657 Johnson, J. A., Winn, J. N., Albrecht, S., Howard, A. W., Marcy, G. W., Gazak, J. Z., 2009a, PASP, 121, 1104 Johnson, J. A., Winn, J. N., Cabrera, N. E., Carter, J. A., 2009b, ApJ, 692, L100 Johnson, J. A., et al., 2007, ApJ, 665, 785 Johnson, J. A., et al., 2008, ApJ, 686, 649 Kipping, D. M., 2008, MNRAS, 389, 1383 Knutson, H. A., Charbonneau, D., Burrows, A., O'Donovan, F. T., Mandushev, G., 2009, ApJ, 691, 866 Koch, D. G., et al., 2010, ApJ, in press, arXiv:1001.0268 Konacki, M., Torres, G., Sasselov, D. D., Jha, S., 2005, ApJ, 624, 372 Konacki, M., et al., 2004, ApJ, 609, L37 Krejcova, T., Budaj, J., Krushevska, V., 2010, Communications in Asteroseismology, submitted, arXiv:1003.1301 Kurucz, R., 1993, ATLAS9 stellar atmosphere programs and 2 km/s grid. Kurucz CD-ROM No. 13 Kurucz, R. L., 1979, ApJS, 40, 1 Lagrange, A., Desort, M., Galland, F., Udry, S., Mayor, M., 2009, A&A, 495, 335 Loeillet, B., et al., 2008, A&A, 481, 529 Mamajek, E. E., Hillenbrand, L. A., 2008, ApJ, 687, 1264 Mandushev, G., et al., 2007, ApJ, 667, L195 Maness, H. L., Marcy, G. W., Ford, E. B., Hauschildt, P. H., Shreve, A. T., Basri, G. B., Butler, R. P., Vogt, S. S., 2007, PASP, 119, 90 Mayor, M., Queloz, D., 1995, Nature, 378, 355 Mazeh, T., Zucker, S., Pont, F., 2005, MNRAS, 356, 955 McCullough, P. R., et al., 2006, ApJ, 648, 1228 P. R., McCullough, arXiv:0805.2921 Mislis, D., Schmitt, J. H. M. M., 2009, A&A, 500, L45 Mislis, D., Schroter, S., Schmitt, J. H. M. M., Cordes, O., Reif, K., 2009, A&A, in press, arXiv:0912.4428 Moutou, C., Pont, F., Bouchy, F., Mayor, M., 2004, A&A, 424, L31 Moutou, C., et al., 2007, A&A, 473, 651 Naef, D., Mayor, M., Beuzit, J. L., Perrier, C., Queloz, D., Sivan, J. P., Udry, S., 2004, A&A, 414, 351 Nelson, B., Davis, W. D., 1972, ApJ, 174, 617 O'Donovan, F. T., et al., 2006, ApJ, 651, L61 O'Donovan, F. T., et al., 2007, ApJ, 663, L37 P´al, A., et al., 2009, ApJ, 700, 783 P´al, A., et al., 2010, MNRAS, 401, 2665 Pietrinferni, A., Cassisi, S., Salaris, M., Castelli, F., 2004, ApJ, 612, 168 Pollacco, D., et al., 2008, MNRAS, 385, 1576 Pollacco, D. L., et al., 2006, PASP, 118, 1407 Pont, F., Bouchy, F., Queloz, D., Santos, N. C., Melo, C., Mayor, M., Udry, S., 2004, A&A, 426, L15 Pont, F., et al., 2007, A&A, 476, 1347 Pont, F., et al., 2008, A&A, 487, 749 Popper, D. M., Etzel, P. B., 1981, AJ, 86, 102 2008, ApJ, submitted, et al., c(cid:13) 0000 RAS, MNRAS 000, 000–000 Homogeneous studies of transiting extrasolar planets. III. 25 Winn, J. N., et al., 2007b, AJ, 133, 1828 Winn, J. N., et al., 2007c, AJ, 134, 1707 Winn, J. N., et al., 2008b, ApJ, 683, 1076 Winn, J. N., et al., 2009, ApJ, 700, 302 al., submitted, 2009, MNRAS, Rabus, M., Deeg, H. J., Alonso, R., Belmonte, J. A., Almenara, J. M., 2009, A&A, 508, 1011 Safronov, V. S., 1972, Evolution of the Protoplanetary Cloud and Formation of the Earth and Planets (Jerusalem: Israel Program for Scientific Translation) Santos, N. C., Israelian, G., Mayor, M., 2004, A&A, 415, 1153 Santos, N. C., et al., 2006, A&A, 458, 997 Sato, B., et al., 2005, ApJ, 633, 465 Scuderi, L. J., Dittmann, J. A., Males, J. R., Green, E. M., Close, L. M., 2009, ApJ submitted, arXiv:0907.1685 Seager, S., Mall´en-Ornelas, G., 2003, ApJ, 585, 1038 Simpson, E. K., et arXiv:0912.3643 Skumanich, A., 1972, ApJ, 171, 565 Snellen, I. A. G., Covino, E., 2007, MNRAS, 375, 307 Snellen, I. A. G., et al., 2009, A&A, 497, 545 Southworth, J., 2008, MNRAS, 386, 1644 (Paper I) Southworth, J., 2009, MNRAS, 394, 272 (Paper II) Southworth, J., Maxted, P. F. L., Smalley, B., 2004a, MNRAS, 349, 547 Southworth, J., Maxted, P. F. L., Smalley, B., 2004b, MNRAS, 351, 1277 Southworth, J., Zucker, S., Maxted, P. F. L., Smalley, B., 2004c, MNRAS, 355, 986 Southworth, J., Smalley, B., Maxted, P. F. L., Claret, A., Etzel, P. B., 2005, MNRAS, 363, 529 Southworth, J., Bruntt, H., Buzasi, D. L., 2007a, A&A, 467, 1215 Southworth, J., Wheatley, P. J., Sams, G., 2007b, MNRAS, 379, L11 Southworth, J., et al., 2009a, MNRAS, 396, 1023 Southworth, J., et al., 2009b, MNRAS, 399, 287 Southworth, J., et al., 2009c, ApJ, 707, 167 Southworth, J., et al., 2010, MNRAS submitted Sozzetti, A., Torres, G., Charbonneau, D., Latham, D. W., Hol- man, M. J., Winn, J. N., Laird, J. B., O'Donovan, F. T., 2007, ApJ, 664, 1190 Sozzetti, A., et al., 2009, ApJ, 691, 1145 Spearman, C., 1904, American Journal of Psychology, 72 Stempels, H. C., Collier Cameron, A., Hebb, L., Smalley, B., Frandsen, S., 2007, MNRAS, 379, 773 Street, R. A., et al., 2007, MNRAS, 379, 816 Torres, G., Winn, J. N., Holman, M. J., 2008, ApJ, 677, 1324 Triaud, A. H. M. J., et al., 2009, A&A, 506, 377 Triaud, A. H. M. J., et al., 2010, A&A, submitted Tripathi, A., et al., 2010, ApJ, in press, arXiv:1004.0692 Udalski, A., Kubiak, M., Szyma´nski, M., 1997, AcA, 47, 319 Udalski, A., Szewczyk, O., Zebrun, K., Pietrzynski, G., Szyman- ski, M., Kubiak, M., Soszynski, I., Wyrzykowski, L., 2002, Acta Astronomica, 52, 317 Udalski, A., et al., 2008, A&A, 482, 299 Valenti, J. A., Fischer, D. A., 2005, ApJS, 159, 141 van Belle, G. T., von Braun, K., 2009, ApJ, 694, 1085 Van Hamme, W., 1993, AJ, 106, 2096 VandenBerg, D. A., Bergbusch, P. A., Dowler, P. D., 2006, ApJS, 162, 375 West, R. G., et al., 2009, A&A, 502, 395 Wheatley, arXiv:1004.0836 Winn, J. N., Holman, M. J., Shporer, A., Fern´andez, J., Mazeh, T., Latham, D. W., Charbonneau, D., Everett, M. E., 2008a, AJ, 136, 267 Winn, J. N., et al., 2007a, ApJ, 665, L167 submitted, P. J., et al., 2010, ApJ, c(cid:13) 0000 RAS, MNRAS 000, 000–000
1505.02941
1
1505
2015-05-12T10:00:23
New Paradigms For Asteroid Formation
[ "astro-ph.EP" ]
Asteroids and meteorites provide key evidence on the formation of planetesimals in the Solar System. Asteroids are traditionally thought to form in a bottom-up process by coagulation within a population of initially km-scale planetesimals. However, new models challenge this idea by demonstrating that asteroids of sizes from 100 to 1000 km can form directly from the gravitational collapse of small particles which have organised themselves in dense filaments and clusters in the turbulent gas. Particles concentrate passively between eddies down to the smallest scales of the turbulent gas flow and inside large-scale pressure bumps and vortices. The streaming instability causes particles to take an active role in the concentration, by piling up in dense filaments whose friction on the gas reduces the radial drift compared to that of isolated particles. In this chapter we review new paradigms for asteroid formation and compare critically against the observed properties of asteroids as well as constraints from meteorites. Chondrules of typical sizes from 0.1 to 1 mm are ubiquitous in primitive meteorites and likely represent the primary building blocks of asteroids. Chondrule-sized particles are nevertheless tightly coupled to the gas via friction and are therefore hard to concentrate in large amounts in the turbulent gas. We review recent progress on understanding the incorporation of chondrules into the asteroids, including layered accretion models where chondrules are accreted onto asteroids over millions of years. We highlight in the end ten unsolved questions in asteroid formation where we expect that progress will be made over the next decade.
astro-ph.EP
astro-ph
New Paradigms For Asteroid Formation Anders Johansen Lund University Emmanuel Jacquet Canadian Institute for Theoretical Astrophysics, University of Toronto Jeffrey N. Cuzzi NASA Ames Research Center Alessandro Morbidelli University of Nice-Sophia Antipolis, CNRS Mus´eum National d’Histoire Naturelle, Institut Universitaire de France Matthieu Gounelle Asteroids and meteorites provide key evidence on the formation of planetesimals in the Solar System. Asteroids are traditionally thought to form in a bottom-up process by coagulation within a population of initially km-scale planetesimals. However, new models challenge this idea by demonstrating that asteroids of sizes from 100 to 1000 km can form directly from the gravitational collapse of small particles which have organised themselves in dense filaments and clusters in the turbulent gas. Particles concentrate passively between eddies down to the smallest scales of the turbulent gas flow and inside large-scale pressure bumps and vortices. The streaming instability causes particles to take an active role in the concentration, by piling up in dense filaments whose friction on the gas reduces the radial drift compared to that of isolated particles. In this chapter we review new paradigms for asteroid formation and compare critically against the observed properties of asteroids as well as constraints from meteorites. Chondrules of typical sizes from 0.1 to 1 mm are ubiquitous in primitive meteorites and likely represent the primary building blocks of asteroids. Chondrule-sized particles are nevertheless tightly coupled to the gas via friction and are therefore hard to concentrate in large amounts in the turbulent gas. We review recent progress on understanding the incorporation of chondrules into the asteroids, including layered accretion models where chondrules are accreted onto asteroids over millions of years. We highlight in the end ten unsolved questions in asteroid formation where we expect that progress will be made over the next decade. 1. INTRODUCTION The Solar System contains large populations of pris- tine planetesimals that have remained relatively unchanged since their formation. Our proximity to the asteroid belt provides astronomers, planetary scientists and cosmo- chemists access to extremely detailed data about asteroid compositions, sizes and dynamics. Planetesimals are the building blocks of both terrestrial planets and the cores of the giant planets, as well as the super-Earths (with various degrees of gaseous envelopes) which are now known to or- bit around a high fraction of solar-type stars (Fressin et al., 2013). The formation of planetesimals is thus a key step to- wards the assembly of planetary systems, but many aspects of the planetesimal formation process remain obscure. Recent progress in understanding planetesimal forma- tion was triggered by two important realisations. The first is that macroscopic dust particles (mm or larger) have poor sticking properties. Laboratory experiments and coagula- tion models show that it is difficult to form planetesimals by direct sticking of silicate particles, most importantly be- cause particle growth stalls at millimeter sizes where the particles bounce off each other rather than stick (Guttler et al., 2010). While some particle growth is still possible at relatively high collision speeds, due to the net transfer of mass from small impactors onto large targets (Wurm et al., 2005; Windmark et al., 2012a), the resulting growth rate is too low to compete with the radial drift of the particles. The second important realisation is that particles are con- centrated to very high densities in the turbulent gas flow. This idea is not new – Whipple (1972) already proposed that large-scale pressure bumps can trap particles, as their radial drift speed vanishes in the pressure bump where the radial pressure gradient is zero. However, the advent of su- percomputing led to the discovery and exploration of a large 1 number of particle concentration mechanisms. Large-scale axisymmetric pressure bumps, akin to those envisioned by Whipple (1972), have been shown to arise spontaneously in simulations of protoplanetary disk turbulence driven by the magnetorotational instability (Johansen et al., 2009a; Simon et al., 2012). Particle densities reach high enough values inside these pressure bumps to trigger gravitational collapse to form planetesimals with sizes up to several 1000 km (Johansen et al., 2007, 2011; Kato et al., 2012). The baroclinic instability, which operates in the absence of cou- pling between gas and magnetic field, leads to the forma- tion of slowly overturning large-scale vortices (Klahr and Bodenheimer, 2003) which can act as dust traps in a similar way as pressure bumps (Barge and Sommeria, 1995). In the streaming instability scenario the particles play an active role in the concentration (Youdin and Goodman, 2005). The relative motion between gas and particles is sub- ject to a linear instability whereby axisymmetric filaments of a slightly increased particle density accelerate the gas towards the Keplerian speed and hence experience reduced radial drift. This leads to a run-away pile up of fast-drifting, isolated particles in these filaments (Johansen and Youdin, 2007). The densities achieved can be as high as 10,000 times the local gas density (Bai and Stone, 2010; Johansen et al., 2012), leading to the formation of planetesimals with characteristic diameters of 100-200 km for particle column densities relevant for the asteroid belt, on a time-scale of just a few local orbital periods. A concern about large-scale particle concentration mod- els is that typically very large particles are needed for op- timal concentration (at least dm in size when the models are applied to the asteroid belt). Chondrules of typical sizes from 0.1 to 1 mm are ubiquitous in primitive mete- orites, but such small particles are very hard to concentrate in vortices and pressure bumps or through the streaming instability. One line of particle concentration models has nevertheless been successful in concentrating chondrules. Swiftly rotating low pressure vortex tubes expel particles with short friction times (Squires and Eaton, 1990, 1991; Wang and Maxey, 1993). This was proposed to explain the characteristic sizes and narrow size ranges of chondrules observed in different chondrites (Cuzzi et al., 2001) and lead to the formation of 100-km-scale asteroids from rare high- density concentrations (Cuzzi et al., 2008, 2010). However, the evaluation of the probability for such high-density con- centrations to occur over sufficiently large scales depends on scaling computer simulations to the very large separa- tions between the energy injection scale and the dissipation scale relevant for protoplanetary disks; Pan et al. (2011) found that the particle clustering gets less contribution from the addition of consecutively larger scales than originally thought in the model of Cuzzi et al. (2008, 2010). Therefore the incorporation of chondrules into the aster- oids is still an unsolved problem in asteroid formation. This is one of the main motivations for this review. We refer the readers to several other recent reviews on the formation of planetesimals which provide a broader scope of the topic beyond asteroid formation (e.g. Cuzzi and Weidenschilling, 2006; Chiang and Youdin, 2010; Johansen et al., 2014). The review is organised as follows. The first two sec- tions discuss the constraints on asteroid formation from the study of meteorites (Section 2) and asteroids (Section 3). In Section 4 we review laboratory experiments and computer simulations of dust coagulation to illustrate the formidable barriers which exist to planetesimal formation by direct sticking. The turbulent concentration model, in which chondrule-sized particles are concentrated at the smallest scales of the turbulent gas flow, is discussed in Section 5. The following Section 6 is devoted to particle concentration in large-scale pressure bumps and through streaming insta- bilities. In Section 7 we discuss layered accretion models where the chondrules are accreted onto planetesimals over millions of years. Finally, in Section 8 we pose ten open questions in the formation of asteroids on which we expect major progress in the next decade. 2. CONSTRAINTS FROM METEORITES Meteorites provide a direct view of the solid material from which the asteroids accumulated, while the crystalli- sation ages of the component particles and the degree of heating and differentiation of the parent bodies give impor- tant information about the time-scales for planetesimal for- mation in the Solar System. Meteorites may be broadly classified in two categories (Weisberg et al., 2006): primitive meteorites (also known as chondrites) and differentiated meteorites. Chondrites, which make up 85% of the observed falls, are basically collections of mm- and sub-mm-sized solids, little modi- fied since agglomeration and lithification (compression) in their parent bodies. They exhibit nonvolatile element abun- dances comparable to the solar photosphere’s (Palme and Jones, 2005). Differentiated meteorites derive from parent bodies which underwent significant chemical fractionations on the scale of the parent body, resulting in the asteroid- wide segregation of an iron core and silicate mantle and crust. In the process, differentiated meteorites have lost not only their accretionary (presumably chondritic) texture, but also their primitive chemical composition, for, depending on which part of the parent body they sample, some may be essentially pure metal (the iron meteorites) while others are essentially metal-free (the achondrites). It is among the components of chondrites that the oldest solids of the solar system, the refractory inclusions (Krot et al., 2004; MacPherson, 2005), in particular Calcium- Aluminium-rich Inclusions (CAI), have been identified. Their age of 4567.3±0.16 Myr (Connelly et al., 2012) marks the commonly accepted “time zero” of the Solar Sys- tem. But the ubiquitous components of chondrites are the eponymous chondrules (Hewins et al., 2005; Connolly and Desch, 2004), which are millimeter-size silicate spherules presumably resulting from transient (and repeated) high- temperature episodes in the disk, but whose very nature remains a long-standing cosmochemical and astrophysical 2 Fig. 1.— The left panel indicates the initial state of a par- ent body of 100 km in radius, which was initially homoge- neous throughout, after heating by 26Al. The proportions of the least (3) to most (6) altered material are constrained by meteorite statistics. The right panel indicates the same body after catastrophic fragmentation and reassembly, as a rubble pile with some highly altered material now near the surface. This body is then cratered by subsequent impacts, releasing samples of all metamorphic grades (adapted from Scott and Rajan, 1981). enigma (Boss, 1996; Desch et al., 2012). All these compo- nents are set in a fine-grained matrix comprised of amor- phous and crystalline grains native to the disk as well as rare presolar grains (Brearley, 1996). While composition- ally primitive, chondrites may have undergone some degree of thermal metamorphism, aqueous alteration and shock processing on their parent body. Despite their general petrographic similarity and roughly solar composition, chondrites are actually quite variable, and 14 distinct chemical groups have been recognized so far, each of which believed to represent a single parent body – sometimes with supporting evidence from cosmic-ray ex- posure ages (Eugster et al., 2006a) – or at least a family of parent bodies formed in the same nebular reservoir (i.e. a compositionally distinctive space-time section of the disk). There are various levels of affinities between these groups (e.g. clans, classes) but we will be here content to distin- guish carbonaceous chondrites [henceforth “CCs”; com- prising the CI, CM, CO, CV, CK, CR, CH, CB groups] from non-carbonaceous chondrites [henceforth “EORs”; com- prising the enstatite (EH, EL), the ordinary (H, L, LL) and the Rumuruti (R) chondrite groups]. Carbonaceous chon- drites are more primitive in the sense that they have a so- lar Mg/Si ratio and a 16O-rich oxygen isotopic composition closer to that of the Sun (e.g. Scott and Krot, 2003). Non- carbonaceous chondrites, though poorer in refractory ele- ments, are more depleted in volatile elements, have subsolar Mg/Si ratios and a more terrestrial isotopic composition for many elements (Trinquier et al., 2009). EORs have gener- ally undergone thermal metamorphism (see sketch in Figure 1) while aqueous alteration has been prevalent in CCs (Huss et al., 2006; Brearley, 2003), but there are again exceptions. 2.1. Primary texture and aerodynamic sorting The texture of most chondrites has been reworked by impact fragmentation and erosion on their parent body. However, rare pieces of CM and CO chondrites have been found, referred to as “primary texture” (Metzler et al., 1992; Brearley, 1993), which seem to retain the nature of a pre- brecciated body. Primary texture appears to consist of noth- ing but dust-rimmed chondrules of very similar properties, loosely pressed together. The constituents of most chondrites appear well-sorted by size, with strong mean size differences from one group to another (Brearley and Jones, 1998). Whether these dif- ferences arise from some regionally or temporally variable bouncing-barrier (Jacquet, 2014a), some aerodynamic sort- ing process (Sections 5 and 7), or some aspect of the mys- terious chondrule formation process itself, they provide an important clue to primary accretion. Hezel et al. (2008) have emphasized the need for better particle counting statis- tics, and indeed one recent chondrule size distribution mea- surement taken from Allende, of a far larger sample than analyzed previously (Fisher et al., 2014), points to a distri- bution substantially broader for that chondrite than previ- ously reported. Aerodynamic sorting has been suggested often as an im- portant factor in selecting for the contents of primary tex- ture (see Cuzzi and Weidenschilling, 2006, for a review). Comparing the aerodynamical friction time of objects of greatly different density, such as silicate and metal grains, shows that their friction times are quite similar in the least altered meteorites (Dodd, 1976; Kuebler et al., 1999), sug- gesting that asteroids selectively incorporated components with specific aerodynamical properties (we discuss this fur- ther in Section 3.3). 2.2. The abundance and distribution of 26Al and 60Fe The melting of the parent bodies of differentiated mete- orites puts important constraints on the time-scale for plan- etesimal formation in the asteroid belt. While electromag- netic heating (Sonett and Colburn, 1968) or impact heat- ing (Keil et al., 1997) have been considered in the litera- ture, the most likely source of planetesimal heating is the decay of the short-lived radionuclides (SLRs) 26Al (with mean life-time τ = 1.0 Myr) and 60Fe (τ = 3.7 Myr) (Urey, 1955). Depending on their respective initial abundance, and on the time of planetesimal accretion, both could have sig- nificantly contributed to planetesimal heating. Additionally, short-lived nuclides provide crystallisation ages which can be calibrated using a long-lived radionuclide decay system such as Pb-Pb, under the assumption that the short-lived ra- dionuclide was homogeneously distributed in the solar pro- toplanetary disk. The content of SLRs in CAIs is usually identified to that of the nascent Solar System (Dauphas and Chaussi- don, 2011). Excesses of 26Mg linearly correlating with 27Al content were first observed in an Allende CAI in 1976 (Lee et al., 1976). This isochron diagram demonstrated the 3 presence of 26Al in the nascent Solar System. The CAIs from a diversity of chondrite groups formed with an ini- tial (26Al/27Al)0 of roughly 5 × 10−5 (MacPherson et al., 2014). A remarkably tight isochron for CAIs in the CV chondrites was obtained by Jacobsen et al. (2008). The deduced (26Al/27Al)0 ratio of (5.23±0.13)×10−5 is often considered as the initial value for the Solar System and the small dispersion as indicative of a narrow formation inter- val (≤ 40,000 yr). However, this interpretation should be limited to the region where the unusually large CV CAIs have formed (Krot et al., 2009). This is especially true since many CAIs are known to have formed without any 26Al (Liu et al., 2012; Makide et al., 2013). This indicates some level of heterogeneity in the 26Al distribution within the region where CAIs formed (assuming that region was unique, which is supported by the ubiquitous 16O enrich- ment of CAIs compared to e.g. chondrules). Larsen et al. (2011) used bulk magnesium isotopic measurements to sug- gest that the heterogeneity of 26Al distribution might have reached 80% of the canonical value in the solar protoplane- tary disk. However, Kita et al. (2013) and Wasserburg et al. (2012) argue that the observed variations can be better as- cribed to small heterogeneities in the stable isotope 26Mg. Although the presence of live 60Fe in the early Solar Sys- tem was demonstrated almost 20 years ago (Shukolyukov and Lugmair, 1993), the determination of the Solar Sys- tem initial abundance is complicated by the difficulty of obtaining good isochrons for CAIs (Quitt´e et al., 2007), given their low abundance in Ni. To bypass that difficulty most measurements were performed on chondrules which are believed to have formed from around the same time as CAIs up to three million years later (Connelly et al., 2012). High initial (60Fe/56Fe)0 ratios were originally reported in chondrules from unequilibrated ordinary chondrites (UOC) which experienced very little heating or aqueous alteration (e.g., Tachibana and Huss, 2003; Mostefaoui et al., 2005; Tachibana et al., 2006). Telus et al. (2012) showed that most of these previous data were statistically biased and that most chondrules do not show any 60Ni excesses in- dicative of the decay of 60Fe. The high values obtained in older publications could be due to statistical biases re- lated to low counts (Telus et al., 2012) or to thermal meta- morphism which would have lead to the redistribution of Ni isotopes (Chaussidon and Barrat, 2009). Recently im- proved techniques for measuring bulk chondrules in un- equilibrated ordinary chondrites have yielded initial Solar System values for (60Fe/56Fe)0 of ≈1×10−8 (Tang and Dauphas, 2012; Chen et al., 2013). This value is consis- tent with that inferred from Fe-Ni isotope measurements of a diversity of differentiated meteorites (Quitt´e et al., 2011; Tang and Dauphas, 2012). In conclusion, it seems that most CAIs formed with an initial ratio (26Al/27Al)0 of roughly 5 × 10−5, which can be considered in a first approximation as the Solar System average or typical initial value. Some heterogeneity was un- doubtedly present, but its exact level is still unknown. On the other hand, it is likely that the initial (60Fe/56Fe)0 of the 4 Solar System was lower than 1×10−8, though high levels of 60Fe (up to 10−6) have been detected in some compo- nents of chondrites. At the time of writing this review there was no evidence of 60Fe heterogeneity within the solar pro- toplanetary disk. 2.3. The origin of 26Al and 60Fe The initial Solar System ratio (26Al/27Al)0 of roughly 5 ×10−5 is well above the calculated average Galactic abun- dance (Huss et al., 2009). This elevated abundance indi- cates a last minute origin for 26Al. Production of 26Al by irradiation has been envisioned in different contexts (e.g., Lee, 1978; Gounelle et al., 2006), but fails to produce enough 26Al relative to 10Be (τ = 2.0 Myr), another SLR whose origin by irradiation is strongly supported by exper- imental data (Gounelle et al., 2013). This leaves stellar de- livery as the only possibility for 26Al introduction in the Solar System. Though Asymptotic Giant Branch Stars pro- duce elevated amounts of 26Al (Lugaro et al., 2012), it is extremely unlikely that stars at this evolutionary stage are present in a star forming region (Kastner and Myers, 1994). Thus massive stars are the best (unique) candidates for the origin of the Solar System’s 26Al. The recently obtained lower estimates for the abundance of 60Fe (see Section 2.2) are compatible with a Galactic background origin, independently of whether this is cal- culated crudely using a box model (Huss et al., 2009), or taking into account the stochastic nature of star formation in molecular clouds (Gounelle et al., 2009). However, if the initial 60Fe abundance corresponds instead to a much higher (60Fe/56Fe)0 ratio of ≈ 10−6, a last minute origin is needed. Irradiation processes cannot account for 60Fe production, because of its richness in neutrons (Lee et al., 1998). The winds of massive stars can also be excluded, given their low abundance in 60Fe. In contrast, supernovae are copious producers of 60Fe, essentially because this SLR is synthesized in abundance during the hydrostatic and ex- plosive phases of such massive stars (Woosley and Heger, 2007). Two astrophysical settings have been envisioned so In the first (classical) model (Cameron and Truran, far. 1977), the supernova ejecta hits a molecular cloud core and provokes its gravitational collapse (Boss and Keiser, 2013) as well as injecting 60Fe and other SLRs. A newer model injects 60Fe in an already formed protoplanetary disk (Hes- ter et al., 2004; Ouellette et al., 2007). In either cases, the supernova progenitor mass is in the range 20-60 M(cid:12), be- cause too massive supernovae are extremely rare and very disruptive to their environment (Chevalier, 2000). Alter- natively, generations of supernovae could have enriched the gas of the giant molecular cloud, so that the solar protoplan- etary disk simply inherited the elevated abundances of the birth cloud (Vasileiadis et al., 2013). All these models suffer from an important difficulty, namely the overabundance of 60Fe in supernova ejecta rela- tive to 26Al. The Solar System 60Fe/26Al mass ratio was ei- ther 4 × 10−3 or 0.4 depending on the adopted initial 60Fe abundance. In any case this is well below the 60Fe/26Al mass ratio predicted by supernovae nucleosynthetic mod- els whose variation domain extends from 1.5 to 5.5 for a progenitor mass varying between 20 and 60 M(cid:12). Hetero- geneity in the composition of supernova ejecta has been proposed as a possible solution to that discrepancy (Pan et al., 2012). However that variability is limited to a fac- tor of 4 which would help resolve the discrepancy only marginally in the case of the (unlikely) high 60Fe value. Fi- nally, the astrophysical context of any supernova model is difficult to reconcile with observations of star-forming re- gions (Williams and Gaidos, 2007; Gounelle and Meibom, 2008). The commonly proposed setting for supernova con- tamination of a protoplanetary disk or or a dense core is similar to that of the Orion Nebula where disks are seen within a few tenths of parsec of the massive star θ1 C Ori. The problem with that setting is that, when θ1 C Ori will ex- plode in 4 Myr from now, these disks will have long evap- orated or formed planets. New disks or cores will have obviously formed by then but they will be at the outskirts of the 10 pc wide HII region created by θ1 C Ori (Cheva- lier, 1999). At such a distance, the quantity of 26Al deliv- ered into these disks or cores is orders of magnitude lower than the quantity present in the Solar System (Looney et al., 2006). In other words, supernova remnants nearby dense phases are extremely rare (Tang and Chevalier, 2014). In conclusion, it seems very unlikely that a nearby supernova was close enough to the Solar System to provide the known inventory of 26Al. If the low value of 60Fe inferred from chondrites is cor- rect, then the presence of 26Al remains to be explained. Supernovae can be excluded as they would vastly overpro- duce 60Fe (see above). The winds of massive stars have long been known to be 26Al and 60Fe-poor (Arnould et al., In the model of Gaidos et al. (2009) and Young 1997). (2014), 26Al is injected at the molecular cloud phase by the winds from a large number of Wolf Rayet stars. The problem with these models is that the Wolf Rayet phase is followed by the supernova explosion and therefore they produce a large excess of 60Fe relative to 26Al and their re- spective Solar System values. To escape that caveat, Young (2014) has argued that Wolf Rayet stars do not explode as supernovae but directly collapse into black holes. Though this possibility has been theoretically envisioned, the recent observation of a Wolf Rayet star going supernova shows this is far from being the rule (Gal-Yam et al., 2014). In addi- tion, models considering injection at the global molecular cloud phase cannot account for the observed heterogeneity of 26Al (see Section 2.2). The last class of models considered injection at the scale of single massive stars. Tatischeff et al. (2010) have en- visioned a single star that escaped from its parent cluster and interacted with a neighboring molecular cloud, inject- ing 26Al through its dense wind. It is compromised again by the proximity of the Wolf Rayet phase with the supernova phase. In addition, Wolf Rayet stars are rare. In contrast Gounelle and Meynet (2012) proposed that 26Al has been injected in a dense shell of mass ∼1000 M(cid:12) collected by the wind of a massive star. Evolutionary models of rotating stars are used, so the injection in the shell starts as early as the entry of the star onto the main sequence, lasts for some Myr and ends well before the supernova explosion. When the collected shell has become dense enough and gravita- tionally unstable, it collapses and a second generation of stars form which contain 26Al. Detailed calculations have shown that as long as the parent star, baptized Coatlicue, is more massive than Mmin = 32 M(cid:12), the abundance of 26Al in the shell is equal to or larger than that of the Solar Sys- tem, depending on the mixing efficiency of the wind ma- terial with the shell. Because the mixing time-scale of the dense shell is comparable to its collapse time-scale, a cer- tain level of 26Al heterogeneity is expected (Gounelle and Meynet, 2012) in agreement with observations. This model is in line with observations of induced star-formation within dense shells around massive stars (Deharveng et al., 2010). Because it corresponds to a common – though not universal – mode of star formation, it implies that the Solar System is not the only of its kind to have formed with 26Al, and that early differentiation of planetesimals might have been common in exo-planetary systems. Gounelle (2014) has es- timated that the occurrence of planetary systems which are rich in 26Al and poor in 60Fe is on the order of 1%. 2.4. Timing of planetesimal accretion Planetesimal accretion itself cannot be dated directly with radioisotopic systems, since the mere agglomeration of different solids incurs no isotopic rehomogenization be- tween different mineral phases. Thus one can only obtain upper limits with the age-dating of pre-accretionary com- ponents (for chondrites) and lower limits from that of sec- ondary (parent body) processes. Important and ever-improving constraints have emerged since the publication of Asteroids III. In particular, Hf-W systematics of achondrites and irons have evidenced early differentiation, sometimes contemporaneous (within errors) with refractory inclusion formation (Kruijer et al., 2012). This indicates that planetesimal formation started very early in the evolution of the solar system. Intriguingly, Libourel and Krot (2007) ascribed some olivine aggregates in chon- drites to this first generation of planetesimals (but see What- tam et al., 2008; Jacquet et al., 2012a, for the alternative view that these formed in the protoplanetary disk). The above evidence for early differentiation is consistent with thermal modeling expectations, as the initial abundance of 26Al was sufficient to melt planetesimals, so that chondrites had to be accreted later to be preserved to the present day. Lower limits on chondrite accretion ages may be ob- tained from phases precipitated during aqueous alteration (see chapter by Krot et al.). Fujiya et al. (2013) obtained ages of 4562.5-4563.8 Myr in CI and CM chondrites (re- call the accepted age of CAI formation of 4567.3±0.16 Myr, Connelly et al., 2012). As to non-carbonaceous chon- drites, recent Mn-Cr dating of fayalite formed during in- 5 cipient fluid-assisted metamorphism in the least metamor- phosed LL ordinary chondrites yields an age of 4564.9+1.3−1.8 Myr (Doyle et al., 2014). Al-Mg systematics in the mildly metamorphosed H chondrite Sainte Marguerite indicate an age of 4563.1±0.2 Myr (Zinner and Gopel, 2002). Ther- mal modeling based on 26Al heating of the H chondrite parent body (one of the ordinary chondrite groups) con- strained by dating of chondrites of different metamorphic degrees indicate accretion ages ranging from 1.8 to 2.7 Myr after refractory inclusions (see Gail et al., 2014, and refer- ences therein), while Fujiya et al. (2013) advocate accre- tion of CI and CM chondrites 3-4 Myr after “time zero” based on similar calculations. If 26Al really is the heat source behind chondrite alteration, then it would indeed make sense that carbonaceous chondrites, little affected by thermal metamorphism, accreted later than their non- carbonaceous counterparts (Grimm and McSween, 1993). Upper limits are provided by the ages of chondrules. In both carbonaceous and non-carbonaceous chondrites, chon- drule Pb-Pb ages in individual meteorites span a range of ∼4564-4567 Myr (Connelly et al., 2012), i.e. 0-3 Myr af- ter refractory inclusions, with younger ages reported by Al- Mg dating for CR (Kita and Ushikubo, 2012) and enstatite chondrite (Guan et al., 2006) chondrules. Little correlation is seen between the age of chondrules and their composi- tion (Villeneuve et al., 2012). The significance of Al-Mg ages for chondrules, in particular in relationship with ap- parently somewhat older Pb-Pb ages (e.g. Connelly et al., 2012), nevertheless remains uncertain. To summarize, it seems clear that differentiated planetes- imals accreted from the outset of the disk evolution while the known chondrites accreted later in the evolution of the disk (∼2-4 Myr after refractory inclusions), as presumably required to escape differentiation, but the exact chronology of chondrite formation and alteration is still in the process of being firmly established. 2.5. Accretion of chondrules directly after formation? Alexander et al. (2008) proposed that the retention of sodium, a volatile element, in chondrules during their for- mation indicated high solid densities in the chondrule- forming regions, up to 7 orders of magnitude above ex- pectations for the solar protoplanetary disk and possibly gravitationally bound (see also Cuzzi and Alexander, 2006), although quite in excess of what the proportion of com- pound chondrules suggest (Ciesla et al., 2004). This raises the possibility that the formation of chondrules and chon- drites, respectively, may have been contemporaneous, as also advocated by Metzler (2012) based on the existence “cluster chondrites” comprised of mutually indented chon- drules (but see Rubin and Brearley, 1996, for a criticism of such hot accretion models). A further argument put forward by Alexander and Ebel (2012) is that chondrule populations in different chondrite groups are quite distinct (Jones, 2012). Indeed Cuzzi et al. (2010) noted that two populations of particles formed si- 6 multaneously at 2 and 4 AU would be well-mixed within 1 Myr for α = 10−4. Here α is the non-dimensional measure of the turbulent viscosity and diffusion. Relevant values for protoplanetary disks are discussed further in Section 6, but we note here that a value of 10−4 corresponds to the conditions which are expected if the mid-plane in the as- teroid formation region is stirred by turbulent surface layers (Oishi et al., 2007). This would suggest that chondrules had to accrete rapidly to avoid homogenization. However, we do not know exactly the turbulence level or original space- time separations between chondrule-/chondrite-forming lo- cations, as the asteroid belt may have undergone significant reshuffling (Walsh et al., 2011). Chondrite groups vary sig- nificantly in bulk composition. This indicates that there has been no thorough mixing of chondrite components, what- ever their individual transformations in the intervening time were, over the whole chondrite formation timescale. So whatever fraction of that time the period between chon- drule formation and chondrite accretion actually represents, it is not to be expected that chondrules should have been well-mixed over the different chondrite-forming reservoirs (Jacquet, 2014b). A link between chondrules and matrix is suggested, in the case of carbonaceous chondrites, by complementarity: the bulk meteorite is solar in some respect (e.g., the Mg/Si ratio) but its separate components (chondrules / matrix / refractory inclusions) are not (chondrules have a typically higher Mg/Si than solar while the converse is true for the matrix, Hezel and Palme, 2010). Complementarity – if verified, as for at least some elements it may reflect ana- lytical biases or parent body processes (Zanda et al., 2012) – would indicate a genetic relationship between chondrules and matrix, which would have exchanged chemical ele- ments upon formation, a relationship which would not be predicted e.g. in an X-wind scenario for chondrule forma- tion (Hezel and Palme, 2010) in which chondrules and ma- trix would have formed in different locations. But it does not require immediate accretion of chondrule and matrix. It only requires chondrules and dust grains to have been trans- ported in a statistically similar way, as was likely the case for a large portion of the disk evolution until accretion. Sev- eral batches of chondrule+dust may have contributed to a given chondrite-forming reservoir provided again they suf- fered no loss of chondrules relative to dust or vice-versa (Cuzzi et al., 2005; Jacquet et al., 2012b). This nonetheless does assume that at the stage of chondrule / matrix agglom- eration, there was no bias for or against the incorporation of any component (Jacquet, 2014a), which may be an impor- tant constraint on the accretion process. The problem is that small dust grains are much harder to incorporate into aster- oidal bodies than the macroscopic chondrules, due to their strong frictional coupling with the gas. One could never- theless envision that chondrules and matrix agglomerated together as compound objects (see Section 4.2.5) prior to incorporation in asteroidal bodies and/or that matrix-sized dust coagulated with ice into lumps with aerodynamical properties equivalent to chondrules. Fig. 2.— The cumulative size distribution of asteroids N (>D), as a function of asteroid diameter D, from Morbidelli et al. (2009). These coagulation models started with either km-sized planetesimals (left plot) or an initial size distribution following the current, observed size distribution of asteroids between 100 and 1000 km in diameter (right plot). The grey line shows the current size distribution of asteroids larger than 100 km in diameter. The model with small planetesimals overproduces asteroids smaller than 100 km in diameter (the upper dashed line represents the current size distribution of small asteroids while the lower dashed lines indicates a tighter constraint on the size distribution directly after accretion of the main belt). Starting with large asteroids gives a natural bump in the size distribution at 100 km in diameter, as the smaller asteroids are created in impacts between the larger primordial counterparts. So the jury is still out on whether chondrule formation immediately preceded incorporation in a chondrite or not. Given the chondrule age spread of 3 Myr within individual chondrites (Connelly et al., 2012), as well as the presence of refractory inclusions and presolar grains which would not have survived chondrule-forming events, it is possible that chondrite components did spend up to a few Myr as free-floating particles in the gaseous disk prior to accretion. 3. CONSTRAINTS FROM THE ASTEROID BELT The modern asteroid belt contains only a fraction of its original planetesimal population. However, the shape of the size distribution of the largest asteroids is primordial and gives important insights into the birth sizes of the planetesi- mals. Asteroid families provide a way to probe whether the asteroids are internally homogeneous or heterogeneous on large scales. 3.1. Asteroid size distribution The observed size distribution of asteroids in the main belt shows a quite steep slope for bodies with diameter D>100 km and a much shallower slope for smaller bod- ies (Bottke et al., 2005). A similar change of slope with an elbow at D∼130 km is observed in the Kuiper belt popula- tion (Fraser et al., 2014). It was expected that the transition from a steep to a shal- lower slope is the consequence of the collisional disruption of smaller bodies. However, Bottke et al. (2005) reached the opposite conclusion by examining the collisional evolution of the asteroid belt in detail. They used a number of con- straints (the total number of catastrophic asteroid families, the survival of the basaltic crust on Vesta, the existence of only 1-2 major basins on that body, etc.) to conclude that the integrated collisional activity of the asteroid belt had to be less than the one of the current main belt population in a putative time-span of 10 Gyr. If one supposes that initially the asteroid belt size distribution had a unique slope (the slope now observed for D>100 km), such a limited colli- sional evolution is not sufficient to reduce the slope of the size distribution of objects smaller than 100 km down to the observed value, i.e. to create the observed elbow. Therefore, Bottke et al. (2005) concluded that the elbow at D∼100 km is a fossil feature of the primordial size distribution. For the Kuiper belt, the constraints on the integrated collisional activity are not as tight as for the asteroid belt. Neverthe- less, models seem to suggest that collisional evolution alone could not create an elbow at diameters larger than 80 km (Pan and Sari, 2005), which is significantly smaller than the observed value. Morbidelli et al. (2009) failed to produce the elbow at D∼100 km in the asteroid belt in collisional coagulation simulations starting from a population of small planetesi- mals (see Figure 2). So, having in mind the new models of formation of large planetesimals from self-gravitating clumps of chondrules or larger pebbles and boulders (Jo- hansen et al., 2007; Cuzzi et al., 2008), they proposed that 100 km was the minimal diameter of the original plan- etesimals. Moreover, not being able to reproduce the cur- rent slope of D>100 km asteroids by mutual collisions be- tween bodies of 100 km in size, Morbidelli et al. (2009) argued that these large planetesimals were born with a sim- ilar slope. However, as we will see in Section 7, the cur- rent slope can be reproduced by considering the accretion 7 of chondrule-sized particles by 100-km-scale planetesimals during the gaseous disk phase, a process not considered by Morbidelli et al. (2009). Weidenschilling (2011) managed to reproduce the elbow at D∼100 km in the asteroid belt from collisional coagula- tion simulations starting from objects of 50 - 200 m radius. Because of the small size of these objects, collisional damp- ing and gas drag keep the disk very dynamically cold (i.e. with a small velocity dispersion among the objects). Hence, in the simulations of Weidenschilling (2011), the elbow at D∼100 km is produced by a transition from dispersion- dominated runaway growth to a regime dominated by Ke- plerian shear, before the formation of large planetary em- bryos. However, any external dynamical stirring of the pop- ulation, for instance due to gas turbulence in the disk, would break this process. Moreover, these simulations are based on the assumption that any collision which does not lead to fragmentation results in a merger, but 100-m-scale objects have very weak gravity and the actual capability of bodies so small to remain bound to each other is questionable. Fi- nally we stress that the formation of 100-m-scale bodies is an open issue, in view of the bouncing barrier and meter- size barrier discussed in Section 4. 3.2. Snowline problems Among the various meteorite types that we know, car- bonaceous meteorites (or at least some of them like CI and CM) contain today a considerable amount (5-10%) of wa- ter by mass. Evidence for water alteration is wide-spread, and it is possible that the original ice content of these bod- ies was higher, close to the 50% value expected from un- fractionated solar abundances. Instead, ordinary chondrites contain < 2% water by weight (e.g. Jarosewich, 1990; Krot et al., 2009); while ordinary chondrites do show signs of water alteration, it is unlikely that they ever contained 50% water ice by mass. This suggests that the parent bodies of these meteorites (CI and CM vs. ordinary) formed on ei- ther side of the condensation line for water, also called the snowline. In other words, the snowline was located in the middle of the asteroid belt at the time when the asteroids formed. The Earth is also very water poor (some 0.1% by mass, although uncertain by a factor of ∼5), suggesting that the planetesimals in its neighborhood were mostly dry and that water was delivered only by the small amount of plan- etesimals accreted from farther out (Morbidelli et al., 2000; Raymond et al., 2004; O’Brien et al., 2014). The very dry enstatite chondrites have been proposed to arise from an ex- tinct portion of the asteroid belt located between 1.6 and 2.1 AU from the Sun (Bottke et al., 2012). The problem with this picture is that chondritic bodies accreted late (2-4 Myr after CAIs) and that the snowline is expected to migrate towards the Sun with time. The temper- ature of the disk is set by the equilibrium between heating and cooling. In the inner part of the disk, the heating is predominantly due to the viscous friction of the gas in dif- ferential rotation around the Sun, so it is related to the gas accretion rate of the star. The cooling rate is governed by how much mass is in the disk in the form of micron-sized grains. In Bitsch et al. (2014a), the snowline is at 2-3 AU only in the earliest phases of disk evolution when the accre- tion rate is M = 10−7M(cid:12) yr−1; but the snowline migrates to 1-2 AU already after 500,000 yr when the accretion rate drops below M = 10−8M(cid:12) yr−1. These models are proba- bly optimistically warm, because it is assumed that the ratio between the mass in µm-sized dust and the mass in gas is 1% (this reflects the solar metallicity but planetesimal ac- cretion should eventually decrease the dust content) and ac- count for the heating produced by stellar irradiation by a star more luminous than our early Sun. According to the observation of the accretion rate of stars as a function of age (Hartmann et al., 1998) and photoevaporation models (Alexander and Armitage, 2006), the accretion rate on the star should drop to M = 10−9M(cid:12) yr−1 at 2-3 Myr, i.e. at the time of chondrite formation. The fact that there is no sign of primary accretion in the Solar System after ∼3 Myr also suggests that the disk disappeared at that time (or that all the solids could have been incorporated into parent bod- ies at that time – however, accretion of solids is unlikely to reach 100% efficiency). Thus, at the time of chondrite for- mation, the snowline should have been well inside the inner edge of the asteroid belt, possibly even inside 1 AU. This is at odds with the paucity of water in the ordinary chondrites and in the Earth. A way to keep the disk warm has been proposed recently by Martin and Livio (2012, 2013). In their model, a dead zone in the disk, with low turbulent viscosity, becomes very massive and develops gravitational instabilities which heat the gas. The unlimited pile-up of gas in the dead zone, how- ever, could be a consequence of a simplified disk transport model, and is not observed in more complex 3D hydrody- namical calculations (Bitsch et al., 2014b). Also, the model of Martin and Livio (2012) predicts a cold region where ice could condense inside the orbit of the Earth, which seems inconsistent with the compositions of Venus and Mercury. A possible way to keep the asteroid belt ice-poor, de- spite freezing conditions, is to stop the inwards flow of icy pebbles in a pressure bump. The inwards drift of peb- bles is caused by friction with the slower moving gas (this effect is discussed more in Section 6.1). Then when the nebula cooled locally below the ice condensation temper- ature, there was no water vapor present to condense, and the planetesimals formed there would have a volatile deple- tion similar to that achieved in a warm disk. Bitsch et al. (2014b) explored the effects of viscosity transitions at the snowline (as suggested by Kretke and Lin, 2007), but found that only very steep radial gradients in the α parameter al- low the formation of a pressure bump. Lambrechts et al. (2014) showed that the formation of a proto-Jupiter of about 20 ME can produce a strong pressure bump just exterior to its orbit. If proto-Jupiter formed at the snowline when the snowline was at the outer edge of the asteroid belt, then its presence would have shielded the asteroid belt and the ter- restrial planet region from the flow of icy pebbles, while 8 small silicate dust and chondrules would remain for longer times in the asteroid belt due to their slow radial drift. 3.3. Are asteroids internally homogeneous? The possibility that the asteroids grew incrementally by layered accretion of chondrules (Section 7) implies that as- teroids are internally heterogeneous, in the most extreme case with a chondritic surface layer residing on a differenti- ated interior. We discuss here briefly whether such internal heterogeneity is supported by observations of the asteroid belt. Burbine et al. (2002) note that 100-150 distinct me- teorite parent bodies, 3/4 of them differentiated, are repre- sented in the meteorite collection. However, this sample is biased towards the sturdy irons and against the weaker, never-melted primitive chondrites. Differentiation is a constraint on formation age. Most studies have suggested that sizeable asteroids forming < 106 years after CAIs are almost certain to have thoroughly melted, but those which formed more than 2 Myr after CAIs may have escaped melting except near their centers. More recent work paints a more complicated picture (see chap- ter by Elkins-Tanton et al.). Specifically, the Allende parent body, source of primitive CV3 chondrites, is thought to have melted near its center, as evidenced by the paleomagnetism detected in these meteorites (Elkins-Tanton et al., 2011); other CV chondrites may show a similar signal (Weiss et al., 2010). This suggests that the CV parent body was differen- tiated in its interior, but preserved an undifferentiated chon- dritic layer. Other authors have ascribed the magnetic field in the CV chondrites to impacts (e.g. Wasson et al., 2013). Modeling of the buoyancy of silicate melts of different composition suggests that, for C-type composition, melt might be dense and remain stable at depth, but for S-type (OC) compositions, and certainly for Enstatite composi- tions, melt is less dense than surrounding material and will rise to manifest on the surface (Fu and Elkins-Tanton, 2014). Thus there might be old, centrally-melted C-type as- teroids such as the Allende parent body, where the evidence for differentiation remains buried, but the lack of evidence for significant surface melting on most S-type asteroids may argue that most of them remain unmelted and undifferenti- ated throughout (see chapter by Elkins-Tanton et al., and Weiss and Elkins-Tanton, 2013). We can sample the internal properties of asteroids in two ways - from the observed color and albedo distributions of collisionally disrupted asteroids in families (see also chap- ters by Nesvorny et al. and Michel et al.) and from the mete- orites that derive from them (see next paragraph). The NE- OWISE mission has measured the albedos of a large num- ber of asteroids and families, finding a dichotomy in albedo, roughly corresponding to the classical S and C types, that is most evident in the outer main belt. Collisional families cover the entire belt, so they avoid the sampling bias that af- fects meteorites. Many families have internal albedo distri- butions that are narrower than the global spread of albedos across families. A similar story is told by observations at vi- 9 sual and near-IR wavelengths (Moth´e-Diniz and Nesvorn´y, 2008). This conclusion is nevertheless complicated by the identification of interlopers in (and exclusion from) the as- teroid families. Hence there is an inherent tendency to ob- serve low albedo variations within asteroid families. In con- trast, the Eos family and the Eunomia family have unusually large internal variance (Moth´e-Diniz et al., 2005, 2008) and the Eunomia family looks like what an internally differen- tiated S-type might produce (see discussion in Weiss and Elkins-Tanton, 2013). The three ordinary chondrite groups (H, L, and LL) are each thought to derive from a single parent body, based on a clustering of cosmic ray exposure ages of chondrites in each group, as if they were excavated by the same few large impacts (Eugster et al., 2006b). The thermal history of the H chondrite parent body under internal heating by 26Al was modeled by Trieloff et al. (2003), Monnereau et al. (2013), and Henke et al. (2013) who all concluded it was a roughly 100-km-radius body. Chondrites from all three OC groups show different amounts of thermal alteration from this pro- cess, designated as metamorphic grades 3-6. Models im- ply that the H chondrite parent, at least, first incurred ther- mal alteration in an “onion-shell” fashion, with the most strongly heated H5/H6 material heated deep in the interior, and then was catastrophically disrupted and reassembled as a rubble pile (Figure 1, Taylor et al., 1987; Scott et al., 2014). The small fractional abundance of minimally al- tered chondrites (H3/H4; Figure 1) constrains the accretion of the H parent body to have happened quickly - probably faster than 3 × 105 yr (Henke et al., 2013; Vernazza et al., 2014)). The constant mass growth rate adopted in Henke et al. (2013) is nevertheless not applicable to layered accre- tion, which results in run-away accretion and deposition of most of the mass towards the end of the growth phase (see further discussion in Section 7). While the major element chemical compositions, and the oxygen isotopic compositions, of the OC groups differ sig- nificantly, there is little or no discernible variation of ei- ther chemical or isotopic composition across metamorphic grades in any of the three groups (Jarosewich, 1990; Wood, 2005); see also Tables 1 and 2 of Clayton et al. (1991). The variation of chondrule size across metamorphic grade is less well studied (A. Rubin, personal communication 2014). Dodd (1976) demonstrated that the difference in the aerodynamical friction time between metal grains and sil- icate chondrules can explain silicate-metal fractionation in the ordinary chondrites. In this picture the LL chondrites are under-abundant in metal because the parent body was successfully able to accrete large silicate-rich chondrules (the chondrules in the LL chondrites are larger than those in both H and L chondrites, see e.g. Nelson and Rubin, 2002). The oxidation of metallic Fe could have happened as the LL chondrite parent body accreted water-bearing phyl- losilicates (Rubin, 2005). In the layered accretion model presented in Section 7, asteroids will accrete larger and larger particles as they grow. Metal grains had a more re- stricted size range and hence a relatively small parent body of the H chondrites would accrete mainly small chondrules together with metal grains. A small size of the H chon- drite parent body is further supported by the low fraction of strongly metamorphosed samples that we have from that group (Dodd, 1976). A similar story is told for the enstatite chondrites (Schneider et al., 1998): the EH group has more metal and smaller chondrules than the EL group. This aero- dynamical size sorting may be evidence of asteroid growth by layered accretion (Section 7) or asteroid formation by turbulent concentration (Section 5). 4. DUST GROWTH BY STICKING Now that we have given an overview of some of the con- straints from meteorites and asteroids, we can turn to the theoretical models of planetesimal formation. In this sec- tion we discuss the growth of dust by direct sticking; sub- sequent chapters discuss gravitational instability models. 4.1. Particle-gas interaction The dynamical behavior of a particle in gas depends on both its size and density, as determined by its friction time τf in the nebula gas (Whipple, 1972; Weidenschilling, 1977a). For particles smaller than the gas molecule mean free path (approximately 10-100 cm in the asteroid belt region, de- pending on the uncertain value of the gas density), the fric- tion time is τf = aρ• vthρg , (1) where a and ρ• are the particle radius and density, and vth and ρg are the gas thermal speed and mid-plane density. The thermal speed of the gas molecules is in turn connected to the isothermal sound speed through vth =(cid:112)8/πcs. The Stokes number St is defined as St = Ωτf, where Ω is the (local) orbital frequency of the protoplanetary disk. The translation from Stokes number to particle size follows (cid:16) (cid:17)−1.5 ρ• √ a = (2/π)ΣgSt ≈ 80 cm St fg(r) r 2.5 AU . (2) Here Σg = 2πρgHg is the gas column density and fg(r) a parameter which sets gas depletion relative to the MMSN as a function of semi-major axis r. We set in the second equality ρ• = 3.5 g/cm3, relevant for chondrules. The Stokes number controls many aspects of dust dynamics. Particles of larger Stokes numbers couple increasingly to larger, longer-lived, and higher-velocity eddies in nebula turbulence, thus acquiring larger relative velocities. Solu- tions for particle velocities have been developed by Voelk et al. (1980) and Mizuno et al. (1988), including closed- form analytical expressions by Cuzzi and Hogan (2003) and Ormel and Cuzzi (2007). Importantly, the Stokes number also controls the degree of sedimentation, with the scale- height of the sedimented mid-plane layer, Hp, given by (cid:114) α Hp = Hg St + α . (3) 10 Here Hg is the gas scale-height and α is a measure of the turbulent diffusion coefficient D normalised as D = αcsHg (see Johansen et al., 2014, for references). Significant sedi- mentation can occur when St (cid:38) α. Chondrules of mm sizes have a Stokes number of St ∼ 10−3; hence chondrules will only settle out of the gas if α (cid:28) 10−3. 4.2. Dust growth The study of dust growth has been an extremely active field, both experimentally and numerically, since Asteroids III. Subsequent reviews were presented by Dominik et al. (2007), Blum and Wurm (2008) and Chiang and Youdin (2010); the older review by Beckwith et al. (2000) is also valuable for basics. Two very recent overview chapters pre- senting both the basic physics and selected recent highlights are by Johansen et al. (2014) and Testi et al. (2014); for efficiency we build on those chapters and here emphasize specifics relevant to asteroid formation. 4.2.1. Sticking Particles can stick if their relative kinetic energy ex- ceeds certain functions of the surface energy of the mate- rial, which depends on composition. At the low relative velocities for small monomers (0.1-10 µm) under nebula conditions, both ice and silicate particles stick easily and form loose, porous aggregates. The process continues until the aggregates are at least 100 µm in radius. Aggregates can continue to grow and stick at larger velocities, if their open structure is able to deform and dissipate energy (Wada et al., 2009). The entire process of growth beyond roughly 100 µm fluffy aggregates depends on just how much these aggregates can be compacted by their mutual collisions. Recent studies that concentrate on icy particles outside the snowline have argued that the high surface energy of ice prevents significant compaction from occurring (and keeps relative velocities small) until particles have grown to ex- tremely large size - hundreds of meters - with extremely low density (Okuzumi et al., 2012). 4.2.2. Bouncing The surface energies of silicates are much smaller than those of ice, so it is easy for even mm-sized silicate parti- cles to compact each other in mutual collisions. Relative velocities large enough to cause compaction and bouncing are acquired by roughly mm-size silicate particles in nebula turbulence. Coagulation modeling by Zsom et al. (2010), consistent with experiments (Guttler et al., 2010; Weidling et al., 2012), revealed a bouncing barrier in this size range where growth of silicate aggregates by sticking ceased. This new barrier joins the long-known fragmentation barrier and radial drift barrier which, even if the bouncing barrier can be breached, tend to frustrate growth in the asteroid belt re- gion beyond dm-m size (Brauer et al., 2008; Birnstiel et al., 2010, 2011). 4.2.3. Fragmentation A simple critical velocity vfrag can be used to refer to fragmentation of two comparable masses. This approach has been modified in some treatments to include some mass transfer by a smaller projectile hitting a larger target at high velocity, even if the projectile is destroyed and some mass is ejected from the target (Wurm et al., 2005). An alternate approach is to treat the fragmentation threshold as a critical kinetic energy per unit mass Q∗, which can be thought of as a critical velocity squared for same-size particles (Stew- art and Leinhardt, 2009). The latter treatment automati- cally accounts for particle size differences and thus allows accretion of small particles to proceed at collision veloci- Q∗. Stewart and Leinhardt (2009) treated solids as weak rubble piles, all calibrated using experimental work by Setoh et al. (2007). These expressions allow for the higher efficiency of low- velocity collisions in fragmentation than for hypervelocity impacts. For particles made of small silicate grains, a value of Q∗ on the order of 104 cm2/s2 is suggested, with an asso- ciated vfrag ∼1 to several m/s, for weak cm-m-sized aggre- gates (Schrapler et al., 2012; Stewart and Leinhardt, 2009; Wada et al., 2009, 2013) or even less (Beitz et al., 2011). ties much higher than the nominal vfrag ∼ √ 4.2.4. Lucky particles The bouncing barrier, in preserving most of the avail- able solids at small sizes, may provide a target-rich envi- ronment for growth of much larger “lucky” particles, which experience few, or low-velocity, collisions and avoid de- struction while steadily growing from much smaller parti- cles to large sizes (Windmark et al., 2012a,b; Garaud et al., 2013; Dra¸zkowska et al., 2013). However, Windmark et al. (2012a,b) and Garaud et al. (2013) did not include radial drift, which is important because the growth time for these lucky particles greatly exceeds the drift time to the Sun. Dra¸zkowska et al. (2013) removed the radial drift prob- lem with a pressure bump at the edge of a dead zone, and still found that the total number of meter-size particles they could produce in 3 × 104 years was in the single digits. An isolated particle cannot trigger collective effects such as the streaming instability (Section 6), but can only keep growing by sweep up. Johansen et al. (2008) and Xie et al. (2010) have modeled this “snowball” stage and find that growth in this fashion is extremely slow unless the nebula turbulent α is very low, because the small feedstock particles are verti- cally diffused to a low spatial density otherwise. 4.2.5. Chondrule rims and chondrule aggregates The fine-grained component of chondrites is not only found in a featureless background matrix; it is also found rimming individual chondrules and other coarse particles, often filling cavities (Metzler et al., 1992; Brearley, 1993; Cuzzi et al., 2005). The origin of these fine grained rims has been debated (Lauretta et al., 2006). One school of thought regards them as accretionary rims, swept up as a cooled chondrule moves relative to the gas and entrained dust or small aggregates (MacPherson et al., 1985; Metzler et al., 1992; Bland et al., 2011). Matrix and rims were reviewed in depth by Scott et al. (1984); amongst many interesting re- sults, they found evidence that rims were accreted as numer- ous aggregates of variable mean composition, rather than as monomers. Rubin (2011) suggested that carbonaceous chondrites formed in dusty regions of the solar protoplan- etary disk and that matrix accumulated into mm-cm-sized highly porous dust balls. In this picture, chondrules ac- quired matrix rims by collisions with these dust balls rather than in collisions with smaller particles. The accretion of fine grained rims was modeled by Mor- fill et al. (1998), Cuzzi (2004), and Ormel et al. (2008). In these models, relative velocities predicted for the nebula en- vironment have been shown to be compatible with sticking and compaction under the theory of Dominik and Tielens (1997); Ormel et al. (2008) added the effects of interchon- drule collisions on compacting the rims further. Ormel et al. (2008) found that chondrules which had accreted porous rims of dust between collisions could stick more easily be- cause the porous rims acted as shock absorbers, resulting in composite, dm-size objects formed of rimmed chondrules. 4.2.6. Gravitational scattering barrier There is one further barrier for particles trying to grow incrementally, by sticking, to km and larger sizes. This gravitational scattering barrier arises because, in turbu- lence, nebula gas has small density fluctuations associated with pressure and vorticity. These density fluctuations, pri- marily on large scales, scatter growing planetesimals to achieve high relative speeds (Laughlin et al., 2004; Nel- son and Papaloizou, 2004; Nelson and Gressel, 2010; Yang et al., 2012). The random velocities acquired by 1-10-km- scale planetesimals in this way are sufficient to put them into an erosive, rather than accretionary, regime throughout most of the Solar System for a range of the most plausible α values (Ida et al., 2008; Stewart and Leinhardt, 2009). Ormel and Okuzumi (2013) found that planetesimals need to have radii of 100 km or more to be able to survive these destructive encounters. Overall, the barriers to formation of 100-km-scale aster- oids by incremental growth-by-sticking appear formidable. For these reasons, a number of models have arisen which avoid these problems with a leapfrog process in which 10- 1000-km-scale asteroids form directly from smaller parti- cles stuck at one of these barriers. 5. TURBULENT CONCENTRATION The leapfrog model sometimes referred to as turbulent concentration (TC) or turbulent clustering can in princi- ple make 10-100-km-radius planetesimals directly from chondrule-sized particles. The turbulent concentration model was motivated originally by laboratory and numeri- cal experiments that showed dense particle clusters forming spontaneously in isotropic turbulence, in which the most intense clustering was seen for particles with friction time 11 τf equal to the Kolmogorov (or smallest eddy) timescale (Squires and Eaton, 1990, 1991; Wang and Maxey, 1993; Eaton and Fessler, 1994). It was realized that this condition was very closely satisfied under canonical nebula condi- tions by the very chondrule-size particles that make up the bulk of primitive chondrites (Cuzzi et al., 1996, 2001). 5.1. A brief history of TC models What remains uncertain and a subject of current study, as with many properties of turbulence, is just how the labo- ratory experiments and numerical simulations translate into actual nebula conditions. The flow regime is described by its turbulence Reynolds number Re = U L/ν = αcsHg/ν where U and L are some typical velocity and length scale of the flow, and ν is the molecular viscosity. All current experiments and simulations are run at Re far smaller than likely nebula values. Cuzzi et al. (2001) noticed that the volume fraction of dense clumps increased with increasing Re, and suggested a scaling which would map the behavior to the much higher values of Re relevant for turbulence in protoplanetary disks. Hogan and Cuzzi (2007) showed that mass-loading feed- back of the particle burden on gas turbulence caused the concentration process to saturate at a mass loading Φ = ρp/ρg ∼ 100, precluding the small, dense clumps advo- cated by Cuzzi et al. (2001). Based on this, Cuzzi et al. (2008) advocated gravitational binding and ultimate sedi- mentation of much larger clumps, on the order of 103 − 104 km, with Φ ∼ 10− 100. They showed that a long-neglected discovery by Sekiya (1983) precludes genuine, dynamical- timescale collapse for dense clumps of chondrule-size parti- cles under plausible conditions. Particles of these sizes and friction times can only sediment slowly towards the center of their bound clump, on a much longer timescale (102−103 orbits). Cuzzi et al. (2008) suggested a criterion for stabil- ity of these bound, yet only slowly-shrinking, clumps which led directly to the conclusion that they would preferentially form large planetesimals comparable in size with the mass- dominant 100 km radius mode advocated for the early as- teroid belt by Bottke et al. (2005) – thus leaping over the long-troublesome mm-m-km-size barriers. Objects formed in this way are compatible with the primary texture seen in the most primitive unbrecciated CM and CO chondrites (Metzler et al., 1992; Brearley, 1993), and would display deep homogeneity of bulk chemical and isotopic properties. Cuzzi et al. (2010) and Chambers (2010) went on to de- scribe an end-to-end primary accretion scenario, combin- ing stability thresholds with calculated probability distri- butions of clump density, finding that a range of nebula conditions (all implying > 10× local enhancement of the usually assumed 1% cosmic solids-to-gas ratio within some few 104 km of the midplane) could match the required rate of planetesimal formation and the characteristic mass mode around 100-200 km diameter. Cuzzi et al. (2010) gave a number of caveats regarding the built-in assumptions of this model; one caveat regarding scale-dependence of the con- 12 centration process has been found to be important enough to change the predictions of the scenario quantitatively (see below). Subsequently, Cuzzi and Hogan (2012) resolved a discrepancy in a key timescale between Cuzzi et al. (2010) and Chambers (2010), which makes planetesimal formation 1000 times faster than in Cuzzi et al. (2010) (and corre- spondingly slower than in Chambers, 2010). 5.2. New insights into turbulent concentration The primary issues are whether it is always Kolmogorov friction time particles that are most effectively concen- trated, and whether the physics of their concentration is scale-invariant. Hogan and Cuzzi (2007) argued by anal- ogy with the observed scale-invariance of turbulent dis- sipation, which is dominated by Kolmogorov-time vortex tubes (little tornados in turbulence) that the concentration of Kolmogorov-friction-time particles would also be scale- invariant (see also Cuzzi and Hogan (2012). They devel- oped a so-called cascade model by which to extend the low- Re results to nebula conditions. The primary accretion scenarios of Cuzzi et al. (2010) and Chambers (2010) used this cascade model to gener- ate density-vorticity PDFs as a function of nebula scale. Pan et al. (2011) ran simulations at higher Re than Hogan and Cuzzi (2007) and found that the clump density PDFs dropped faster than would be predicted by the scale- invariant cascade. They suggested that the physics of parti- cle concentration might indeed be scale-dependent, and that planetesimal formation rates obtained using the Hogan and Cuzzi (2007) cascade might be significantly overestimated. Ongoing work supports this concern about scale depen- dence. Cuzzi et al. (2014) have analyzed much higher Re simulations (Bec et al., 2010) and found that the cascade measures called “multiplier distributions”, which determine how strongly particles get clustered at each spatial scale, do depend on scale at least over the largest decade or so of length scale; that is, the scale-invariant inertial range for particle concentration and dissipation does not become es- tablished at the largest scale, causing little concentration to occur until roughly an order of magnitude smaller scale. Because the cascade process is multiplicative, this slow start means that fewer dense zones are to be found at any given scale size than previously thought. New, scale-dependent cascades can now be implemented using the results of Cuzzi et al. (2014) at the largest scales. The quantitative implications are not clear as of this writing. One qualitative result is that particles with friction times 2-10× longer than originally suggested are more strongly clustered at the most relevant (larger) length scales (con- sistent with previous work by Bec et al., 2007; Pan et al., 2011). Because we know the size and density of chon- drules, the implications will be different preferred values of gas density and/or turbulent α (see Cuzzi et al., 2001). A related implication is that concentrated particle sizes may not be as narrowly peaked as shown in Cuzzi et al. (2001). However, at the same time, new observations are also showing broader chondrule size distributions in chon- drites (Fisher et al., 2014). 6. PRESSURE BUMPS AND STREAMING INSTA- BILITY The turbulent concentration mechanism described in the previous section operates on the smallest scales of the turbu- lent flow (although the vortical structures which expel par- ticles can be very elongated). The dynamical time-scales on such small length-scales are much shorter than the local orbital time-scale of the protoplanetary disk. In contrast, the largest scales of the turbulent flow are dominated by the Coriolis force, and this allows for the emergence of large- scale geostrophic structures (high-pressure regions in per- fect balance between the outwards-directed pressure gradi- ent force and the inwards-directed Coriolis force). Whipple (1972) found that particles are trapped by the zonal flow surrounding large-scale pressure bumps. Pres- sure bumps (in a way azimuthally extended analogs to the vortices envisioned in Barge and Sommeria, 1995) can arise through an inverse cascade of magnetic energy (Johansen et al., 2009a; Simon et al., 2012; Dittrich et al., 2013) in tur- bulence driven by the magnetorotational instability (Balbus and Hawley, 1991). Pressure bumps concentrate primarily large (0.1-10 m) particles which couple to the gas on an or- bital time-scale (Johansen et al., 2006), reaching densities at least 100 times the gas density which leads to the forma- tion of 1000-km-scale planetesimals (Johansen et al., 2007, 2011). The magnetorotational instability is nevertheless no longer favoured as the main driver of angular momentum transport in the asteroid formation region of the solar pro- toplanetary disk, since the ionisation degree is believed to be too low for coupling the gas to the magnetic field (see review by Turner et al., 2014). The magnetorotational instability can still drive turbu- lence (with α in the interval from 10−3 to 10−2) in the mid- plane close to the star (within approximately 1 AU where the ionisation is thermal) and far away from the star (beyond 20 AU where ionising cosmic rays and X-rays penetrate to the mid-plane). Accretion through the “dead zone”, situated between these regions of active turbulence, can occur in ionised surface layers far above the mid-plane (Oishi et al., 2007), from disk winds (Bai and Stone, 2013) and by purely hydrodynamical instabilities in the vertical shear of the gas (Nelson et al., 2013) or radial convection arising from the subcritical baroclinic instability (Klahr and Bodenheimer, 2003; Lesur and Papaloizou, 2010). The mid-plane is be- lieved to be stirred to a mild degree by these hydrodynam- ical instabilities or by perturbations from the active layers several scale-heights above the mid-plane, driving effective turbulent diffusivities in the interval from 10−5 to 10−3 in the mid-plane. The inner and outer edges of this “dead zone”, where the turbulent viscosity transitions abruptly, are also possible sites of pressure bumps and large-scale Rossby vortices which feed off the pressure bumps (Lyra et al., 2008, 2009). 13 6.1. Streaming instability The low degree of turbulent stirring in the asteroid for- mation region also facilitates the action of the streaming instability, a mechanism where particles take an active role in the concentration process (Youdin and Goodman, 2005; Youdin and Johansen, 2007; Johansen and Youdin, 2007). The instability arises from the speed difference between gas and solid particles. The gas is slightly pressure-supported in the direction pointing away from the star, due to the higher temperature and density close to the star, which mimics a reduced gravity on the gas. The result is that the gas orbital speed is approximately 50 m/s slower than the Keplerian speed at any given distance from the star. Solid particles are not affected by the global pressure gradient – they would move at the Keplerian speed in absence of drag forces, but drift radially due to the friction from the slower moving gas. The friction exerted from the particles back onto the gas leads to an instability whereby a small overdensity of parti- cles accelerates the gas and diminishes the difference from the Keplerian speed. The speed increase in turn reduces the local headwind on the dust. This slows down the ra- dial drift of particles locally, which leads to a run-away pro- cess where isolated particles drift into the convergence zone and the density increases exponentially with time. This pic- ture is a bit simplified as Youdin and Goodman (2005) and Jacquet et al. (2011) showed that the streaming instability operates only in presence of rotation, i.e. the instability re- lies on the presence of Coriolis forces. This explains why the instability occurs on relatively large scales of the pro- toplanetary disk where Coriolis forces are important, typi- cally a fraction of an astronomical unit, and operates most efficiently on large particles with frictional coupling times around 1/10 of the orbital time-scale (typically dm sizes at the location of the asteroid belt). Computer simulations which follow the evolution of the streaming instability into its non-linear regime show the emergence of axisymmetric filaments with typical separa- tions of 0.2 times the gas scale-height (Yang and Johansen, 2014) and local particle densities reaching several thousand times the gas density (Bai and Stone, 2010; Johansen et al., 2012). These high densities trigger the formation of large planetesimals (100–1000 km in diameter) by gravitational fragmentation of the filaments (Johansen et al., 2007), al- though planetesimal sizes decrease to approximately 100 km for a particle column density comparable to that of the solar protoplanetary disk (Johansen et al., 2012). An important question concerning planetesimal forma- tion through the streaming instability is whether the pro- cess can operate for particles as small as chondrules in the asteroid belt. In Figure 3 we show numerical experi- ments from Carrera et al. (2015) on the streaming instabil- ity in particles with sizes down to a fraction of a millimeter. The streaming instability requires a threshold particle mass- loading Z = Σp/Σg, where Σp and Σg are the particle and gas column densities, to trigger the formation of overdense filaments (Johansen et al., 2009b; Bai and Stone, 2010). Fig. 3.— Space-time plots of particle concentration by streaming instabilities, from Carrera et al. (2015), with the x-axis indicating the radial distance from the centre of the simulation box and the y-axis the time (on the left) and the dust-to-gas ratio (on the right). The four columns show particle sizes 0.5 mm (τf = 0.001Ω−1), 1.5 mm (τf = 0.003Ω−1), 5 mm (τf = 0.01Ω−1) and 1.5 cm (τf = 0.03Ω−1). Simulations start with a mean dust-to-gas ratio of Z = 0.01, but gas is removed on a time-scale of 30 orbits (1 orbit = 2πΩ−1), increasing the dust-to-gas ratio accordingly . While cm-sized particles concentrate in overdense filaments already at a modest increase in dust-to-gas ratio to Z = 0.015, smaller particles require consecutive more gas removal to trigger clumping. The simulations in Figure 3 start at Z = Z0 = 0.01, but the particle mass-loading is continuously increased by re- moving the gas on a time-scale of 30 orbital periods. This was done to identify how the critical value of Z depends on the particle size. The result is that overdense filaments form already at Z = 0.015 for cm-sized particles, while large chondrules of mm sizes require Z = 0.04 to trigger fila- ment formation. Chondrules smaller than mm do not form filaments even at Z = 0.08. A lowered gas column density may thus be required to trigger concentration of chondrule-sized particles by the streaming instability. It is possible that the solar proto- planetary disk had a lower gas density than what is in- ferred from the current mass of rock and ice in the planets (which multiplied by 100 gives the MMSN), if the planet- forming regions of the nebula were fed by pebbles drifting in from larger orbital distances (Birnstiel et al., 2012). In this picture the growing planetesimals and planets are fed by drifting pebbles, so that the current mass of the plan- ets was achieved by the integrated capture efficiency of the drifting solids; this allows for gas column densities lower than in the Minimum Mass Solar Nebula to be consistent with the current masses of planets in the Solar System. The gas will also be removed by accretion and photoevaporation (Alexander and Armitage, 2006). The high mass-loading in the gas could be obtained through pile-up by radial drift and release of refractory grains near the ice line (Sirono, 2011). Turbulence as weak as α ∼ 10−7 is necessary to allow the sedimentation of chondrules (with St ∼ 10−3) into a thin mid-plane layer with scale-height Hp = 0.01Hg and ρp ≈ ρg (see equation 2), the latter being a necessary den- sity criterion for activating particle pile up by streaming in- stabilities. Very low levels of α are consistent with proto- planetary disk models where angular momentum is trans- ported by disk winds and the mid-plane remains laminar (Bai and Stone, 2013), except for mild stirring by Kelvin- Helmholtz (Youdin and Shu, 2002) and streaming instabili- ties (Bai and Stone, 2010). Weak turbulence also facilitates the formation of dm- sized chondrule aggregates (Ormel et al., 2008), which would concentrate much more readily in the gas. Stirring by hydrodynamical instabilities in the mid-plane, such as the vertical shear instability (Nelson et al., 2013), would preclude significant sedimentation of chondrule-sized par- ticles and affect the streaming instability, as well as the for- mation of chondrule aggregates, negatively. An alternative possibility is that the first asteroid seeds in fact did not form from chondrules (or chondrule aggregates), but rather from larger icy particles which would have been present in the asteroid formation region in stages of the protoplanetary disk where the ice line was much closer to the star (Mar- tin and Livio, 2012; Ros and Johansen, 2013). Chondrules 14 could have been incorporated into by later chondrule accre- tion (see Section 7). for particle sizes (cid:18) R (cid:17)−3(cid:18) Σg 50 km (cid:19)−3 (cid:19)3(cid:18) ∆v (cid:19) 53 m/s . (5) 2.5 AU ΣMMSN 7. LAYERED ACCRETION The turbulent concentration model and the streaming in- stability, reviewed in the previous sections, are the leading contenders for primary accretion of chondrules into chon- drites. However, none of the two are completely successful in explaining the dominance of chondrules in chondrites: the turbulent concentration models may not be able to con- centrate sufficient amounts for gravitational collapse, while the streaming instability relies on the formation of chon- drule aggregates and/or gas depletion and pile up of solid material from the outer parts of the protoplanetary disk. In the layered accretion model the chondrules are instead ac- creted onto the growing asteroids over millions of years af- ter the formation of the first asteroid seeds – those first seeds forming by direct coagulation from a population of 100-m- sized planetesimals as envisioned in Weidenschilling (2011) or by one or more of the particle concentration mechanisms described in the previous sections. 7.1. Chondrule accretion RB R Chondrules are perfectly sized for drag-force-assisted accretion onto young asteroids. The ubiquity of chon- drules inside chondrites, and their large age spread (Con- nelly et al., 2012), indicates that planetesimals formed and orbited within a sea of chondrules. Chondrules would have been swept past these young asteroids with the sub- Keplerian gas. The gas is slightly pressure-supported in the radial direction and hence moves slower than the Ke- plerian speed by the positive amount ∆v (Weidenschilling, 1977b; Nakagawa et al., 1986). The Bondi radius RB = GM/(∆v)2 marks the impact parameter for gravitational scattering of a chondrule by an asteroid of mass M, with (cid:18) R (cid:19)2(cid:18) ∆v (cid:19)−2(cid:18) 50 km = 0.87 53 m/s 3.5 g/cm3 . (4) Here we have normalised by ∆v = 53 m/s, the nominal value in the Minimum Mass Solar Nebula model of Hayashi (1981), and used the chondrule density ρ• = 3.5 g/cm3 as a reference value. Chondrules with friction time comparable to the Bondi time-scale tB = RB/∆v are accreted by the asteroid (Johansen and Lacerda, 2010; Ormel and Klahr, 2010; Lambrechts and Johansen, 2012). The accretion ra- dius Racc can be calculated numerically as a function of asteroid size and chondrule size by integrating the trajec- tory of a chondrule moving with the sub-Keplerian gas flow past the asteroid. The accretion radius peaks at Racc ≈ RB for tf /tB in the range from 0.5 to 10 (Lambrechts and Jo- hansen, 2012). Accretion at the full Bondi radius happens (cid:19) ρ• a = [0.008, 0.16] mm ×(cid:16) r An asteroid of radius 50 km thus “prefers” to accrete chon- drules of sizes smaller than 0.1 mm, corresponding to the smallest chondrules found in chondrites. At 100 km in ra- dius the preferred chondrule size is closer to 0.2 mm, a 200 km radius body prefers mm-sized chondrules and larger bodies can only grow efficiently if they can accrete chon- drules of several mm or cm in diameter. Carbonaceous chondrites accreted significant amounts of CAIs and ma- trix together with their chondrules; Rubin (2011) suggested that matrix was accreted in the form of cm-sized porous aggregates with aerodynamical friction time comparable to chondrules and CAIs. Aerodynamical accretion of chondrules could explain the narrow range of chondrule sizes found in the various classes of meteorites. The model predicts that asteroids ac- crete increasingly larger chondrules as they grow. This pre- diction may be at odds with the little variation in chondrule sizes found within chondrite classes (70% of EH3 and CO3 chondrules have apparent diameters within a factor of 2 of the mean apparent diameters in the group, according to Ru- bin, 2000). The least metamorphosed LL chondrites nev- ertheless do seem to host on the average larger chondrules (Nelson and Rubin, 2002). More metamorphosed LL chon- drites actually show a lack of small chondrules; this could be due to the fact that the smallest chondrules disappeared from the strongly heated central regions of the parent body. The accretion rate of chondrules (and other macroscopic particles) is M = πf 2 BR2 Bρp∆v . (6) Here fB parameterises the actual accretion radius relative to the Bondi radius and ρp is the chondrule density. Accretion B ∝ M 2 of chondrules is a run-away process, since M ∝ R2 if the optimal chondrule size is present (so that fB = 1 in (cid:19)3 (cid:19)−3(cid:18) ∆v equation 6). The characteristic growth time-scale is (cid:19)−1 (cid:18) R (cid:19)−1(cid:18) (cid:17)2.75(cid:18) Σg = 1.66 Myr texp = 53 m/s 50 km M M ρ• r ×(cid:16) . (7) 2.5 AU ΣMMSN 3.5 g/cm3 We assumed here that the chondrules have sedimented to a thin mid-plane layer of thickness 1% relative to the gas scale-height. The strong dependence of the accretion rate on the planetesimal mass will drive a steep differential size distribution of a population of planetesimals accreting chondrules. This is illustrated in Figure 4 where asteroid seeds with initial sizes from 10 to 50 km in radius have been exposed to chondrule accretion over 5 Myr. The value of the 15 mid-plane layer of thickness 1% of the gas scale height; and indeed the time-scale to grow to the current asteroid popu- lation is 3 Myr in the simulation shown in Figure 4, about twice as long as in equation (7), but in good agreement with the ages of the youngest chondrules. 8. OPEN QUESTIONS The formation of asteroids is a complex problem which will only be solved through a collective effort from as- tronomers, planetary scientists and cosmochemists. Al- though many details of asteroid formation are still not un- derstood, we hope to have convinced the reader that new insights have been achieved in many areas in the past years. Here we highlight ten areas of open questions in which we believe that major progress will be made in the next decade: 1) Short-lived radionuclides. What is the origin of the short-lived radioactive elements which melted the differen- tiated parent bodies? Was 26Al heterogeneous in the solar protoplanetary disk? How did the young Solar System be- come polluted in 26Al without receiving large amounts of 60Fe, an element that is copiously produced in supernovae? 2) Maintaining free-floating chondrules and CAIs. How is it possible to preserve chondrules and CAIs for millions of years in the disk before storing them in a chondritic body, without mixing them too much to erase chondrule classes and chondrule-matrix complementarity? What are we miss- ing that makes this issue so paradoxical? 3) Chondrules versus matrix. Why do carbonaceous chondrites contain large amounts of matrix while ordinary chondrites contain very little matrix? Did the matrix en- ter the chondrites as (potentially icy) “matrix lumps” or on fine-grained rims attached to chondrules and other macro- scopic particles? 4) Initial asteroid sizes. What is the origin of the steep differential size distribution of asteroids beyond the knee at 100 km? Did asteroids form small as in the coagulation picture, medium-sized as in the layered accretion model or large as in some turbulent concentration models? Why do Kuiper belt objects, which formed under very different con- ditions in temperature and density, display a similar size distribution as asteroids? 5) The origin of asteroid classes. How is the radial gra- dient of asteroid composition produced and retained in the presence of considerable pre-accretionary turbulent mixing and post-accretionary dynamical mixing? Is asteroid for- mation a continuous process which happens throughout the life-time of the protoplanetary disk? What do the different chondrite groups mean in terms of formation location and time? 6) Dry and wet chondrites. Why do we have dry chon- drites (enstatite, ordinary)? If chondrites form at 2-4 Myr after CAIs, then the snowline should have been well in- side the inner edge of the asteroid belt. Are there over- looked heating sources which could keep the ice line at 3 AU throughout the life-time of the protoplanetary disk? Or did the asteroid classes form at totally different places only Fig. 4.— The size distribution of asteroids and planetary embryos after accreting chondrules with sizes from 0.1 to 1.6 mm in diameter for 5 Myr. The original asteroid sizes had sizes between 10 and 50 km in radius (yellow line), here envisioned to form by the streaming instability in a population of dm-sized icy particles. The resulting size dis- tribution of asteroids is in good agreement with the bump at 70 km in radius, the steep size distribution from 70 km to 200 km, and the shallower size distribution of larger aster- oids whose chondrule accretion is slowed down by friction within their very large Bondi radius. Figure based on the results of Johansen et al. (2015). turbulent viscosity is α = 2× 10−6. The resulting differen- tial size distribution (which manifests itself after around 3 Myr of chondrule accretion) shows a bump at 70 km in ra- dius, a steep decline towards 200 km in radius, and finally a more slow decline towards larger asteroids. The shallower decline is caused by the lack of cm-sized particles needed to drive the continued run-away accretion of large asteroids (this can be seen as a drop in fB in equation 6). All the fea- tures in the size distribution in Figure 4 are in good agree- ment with features of the observed size distribution of as- teroids which are not explained well in coagulation models (Morbidelli et al., 2009). Layered accretion of chondrules can readily explain the large age spread of individual chondrules inside chondrites (Connelly et al., 2012), as well as the remanent magnetisa- tion of the Allende meteorite (Elkins-Tanton et al., 2011), imposed on the accreted chondrules from the internal dy- namo in the parent body’s molten core. Efficient chon- drule accretion requires, as does the streaming instability discussed in the previous section, sedimentation of chon- drules to a thin mid-plane layer. The turbulent viscosity of α = 2×10−6 in Figure 4 is nevertheless significantly larger than the α ∼ 10−7 needed to sediment chondrules to a thin 16 101001000R [km]10−2100102104106dN/dR [km−1]CeresMoonMarsNominal model (t=5.0 Myr)Initial size distributionAsteroid belt (×3931)Embryos to be transported to their current orbits later? 7) Internal structure of asteroids. Does the chondrite and asteroid family evidence suggest that the primary asteroids – before internal heating – are homogeneous, roughly 100- km-diameter bodies composed of a physically, chemically, and isotopically homogeneous mix of chondrule-sized com- ponents? Or is internal heterogeneity, as may be the case for the Allende parent body, prevalent? 8) Turbulent concentration of chondrules. Under what nebula conditions can vortex tubes over a range of nebula scales concentrate enough chondrules into volumes that are gravitationally bound, at a high enough rate to produce the primordial asteroids and meteorite parent bodies directly? What other roles could turbulent concentration play in plan- etesimal formation given that the optimally concentrated particle is chondrule-sized under nominal values of the tur- bulent viscosity? 9) Streaming instability with chondrules. Will the con- ditions for streaming instabilities to concentrate chondrule- sized particles, i.e. gas depletion and/or particle pile up, be fulfilled in the protoplanetary disk? How does an over- dense filament of chondrule-sized particles collapse under self-gravity given the strong support by gas pressure? 10) Layered accretion. What is the origin of the apparent scarcity of heterogeneous asteroid families, given that aster- oids orbiting within an ocean of chondrules should accrete these prodigiously? What is the thermal evolution of early- formed asteroid seeds that continue to accrete chondrules over millions of years? Acknowledgments. AJ was supported by the Swedish Research Council (grant 2010-3710), the European Re- search Council under ERC Starting Grant agreement 278675-PEBBLE2PLANET and by the Knut and Alice Wallenberg Foundation. He would like to thank Benjamin Weiss for stimulating discussions on layered accretion. EJ wishes to remember his colleague and friend Guillaume Barlet (1985-2014), who as a short-lived radionuclide the- orist and chondrule/refractory inclusion specialist would certainly have contributed to the new paradigms discussed herein, true to his attachment to interdisciplinary interac- tions, but left us far too early. JC thanks Chris Ormel for a careful reading, and Ed Scott, Alan Rubin, Noriko Kita, and Gerry Wasserburg for helpful comments and references. We would like to thank Alan Rubin and an additional anony- mous referee for insightful referee reports. REFERENCES Alexander C. M. O. and Ebel D. S. (2012) Questions, questions: Can the contradictions between the petrologic, isotopic, ther- modynamic, and astrophysical constraints on chondrule for- mation be resolved? Meteoritics and Planetary Science, 47, 1157. Alexander C. M. O., Grossman J. N., Ebel D. S. et al. (2008) The formation conditions of chondrules and chondrites. Science, 320, 1617. Alexander R. D. and Armitage P. J. (2006) The Stellar Mass- Accretion Rate Relation in T Tauri Stars and Brown Dwarfs. 17 Astrophys. J. Lett., 639, L83. Arnould M., Paulus G., and Meynet G. (1997) Short-lived ra- dionuclide production by non-exploding Wolf-Rayet stars. As- tron. Astrophys., 321, 452. Bai X.-N. and Stone J. M. (2010) Dynamics of Solids in the Mid- plane of Protoplanetary Disks: Implications for Planetesimal Formation. Astrophys. J., 722, 1437. Bai X.-N. and Stone J. M. (2013) Wind-driven Accretion in Pro- toplanetary Disks. I. Suppression of the Magnetorotational In- stability and Launching of the Magnetocentrifugal Wind. As- trophys. J., 769, 76. Balbus S. A. and Hawley J. F. (1991) A powerful local shear in- stability in weakly magnetized disks. I - Linear analysis. II - Nonlinear evolution. The Astrophysical Journal, 376, 214. Barge P. and Sommeria J. (1995) Did planet formation begin in- side persistent gaseous vortices? Astron. Astrophys., 295, L1. Bec J., Biferale L., Cencini M. et al. (2007) Heavy Particle Con- centration in Turbulence at Dissipative and Inertial Scales. Physical Review Letters, 98, 8, 084502. Bec J., Biferale L., Cencini M. et al. (2010) Intermittency in the velocity distribution of heavy particles in turbulence. Journal of Fluid Mechanics, 646, 527. Beckwith S. V. W., Henning T., and Nakagawa Y. (2000) Dust Properties and Assembly of Large Particles in Protoplanetary Disks. Protostars and Planets IV, p. 533. Beitz E., Guttler C., Blum J. et al. (2011) Low-velocity Collisions of Centimeter-sized Dust Aggregates. Astrophys. J., 736, 34. Birnstiel T., Dullemond C. P., and Brauer F. (2010) Gas- and dust evolution in protoplanetary disks. Astron. Astrophys., 513, A79. Birnstiel T., Klahr H., and Ercolano B. (2012) A simple model for the evolution of the dust population in protoplanetary disks. Astron. Astrophys., 539, A148. Birnstiel T., Ormel C. W., and Dullemond C. P. (2011) Dust size distributions in coagulation/fragmentation equilibrium: nu- merical solutions and analytical fits. Astron. Astrophys., 525, A11. Bitsch B., Johansen A., Lambrechts L. et al. (2014a) The structure of protoplanetary discs around evolving young stars. Submitted to Astronomy & Astrophysics. Bitsch B., Morbidelli A., Lega E. et al. (2014b) Stellar irradiated discs and implications on migration of embedded planets. III. Viscosity transitions. Astron. Astrophys., 570, A75. Bland P. A., Howard L. E., Prior D. J. et al. (2011) Earliest rock fabric formed in the Solar System preserved in a chondrule rim. Nature Geoscience, 4, 244. Blum J. and Wurm G. (2008) The Growth Mechanisms of Macro- scopic Bodies in Protoplanetary Disks. Annu. Rev. Astron. As- trophys., 46, 21. Boss A. P. (1996) A concise guide to chondrule formation mod- in: Chondrules and the Protoplanetary Disk, (edited by els. R. H. Hewins, R. H. Jones, & E. R. D. Scott), pp. 257–263. Boss A. P. and Keiser S. A. (2013) Triggering Collapse of the Presolar Dense Cloud Core and Injecting Short-lived Radioiso- topes with a Shock Wave. II. Varied Shock Wave and Cloud Core Parameters. Astrophys. J., 770, 51. Bottke W. F., Durda D. D., Nesvorn´y D. et al. (2005) The fos- silized size distribution of the main asteroid belt. Icarus, 175, 111. Bottke W. F., Vokrouhlick´y D., Minton D. et al. (2012) An Ar- chaean heavy bombardment from a destabilized extension of the asteroid belt. Nature, 485, 78. Brauer F., Dullemond C. P., and Henning T. (2008) Coagulation, fragmentation and radial motion of solid particles in protoplan- etary disks. Astron. Astrophys., 480, 859. Brearley A. and Jones A. (1998) Chondritic meteorites. in: Plane- tary Materials, (edited by J. J. Papike), chap. 3, pp. 3–1–3–398, Mineralogical Society of America. Brearley A. J. (1993) Matrix and fine-grained rims in the unequili- brated CO3 chondrite, ALHA77307 - Origins and evidence for diverse, primitive nebular dust components. Geochim. Cos- mochim. Acta, 57, 1521. Brearley A. J. (1996) Nature of matrix in unequilibrated chon- in: Chon- drites and its possible relationship to chondrules. drules and the Protoplanetary Disk, (edited by R. H. Hewins, R. H. Jones, & E. R. D. Scott), pp. 137–151. Brearley A. J. (2003) Nebular versus Parent-body Processing. Treatise on Geochemistry, 1, 247. Burbine T. H., McCoy T. J., Meibom A. et al. (2002) Meteoritic Parent Bodies: Their Number and Identification. Asteroids III, pp. 653–667. Cameron A. G. W. and Truran J. W. (1977) The supernova trigger for formation of the solar system. Icarus, 30, 447. Carrera D., Johansen A., and Davies M. B. (2015) Formation of asteroids from mm-sized chondrules. Astron. Astrophys., in press. Chambers J. E. (2010) Planetesimal formation by turbulent con- centration. Icarus, 208, 505. Chaussidon M. and Barrat J.-A. (2009) 60Fe in Eucrite NWA 4523: Evidences for Secondary Redistribution of Ni and for Secondary Apparent High 60Fe/56Fe Ratios in Troilite. in: Lunar and Planetary Science Conference, vol. 40 of Lunar and Planetary Science Conference, p. 1752. Chen J. H., Papanastassiou D. A., Telus M. et al. (2013) Fe-Ni Iso- topic Systematics in UOC QUE 97008 and Semarkona Chon- drules. Lunar Planet. Sci., 44, 2649. Chevalier R. A. (1999) Supernova Remnants in Molecular Clouds. Astrophys. J., 511, 798. Chevalier R. A. (2000) Young Circumstellar Disks near Evolved Massive Stars and Supernovae. Astrophys. J. Lett., 538, L151. Chiang E. and Youdin A. N. (2010) Forming Planetesimals in So- lar and Extrasolar Nebulae. Annual Review of Earth and Plan- etary Sciences, 38, 493. Ciesla F. J., Lauretta D. S., and Hood L. L. (2004) The frequency of compound chondrules and implications for chondrule for- mation. Meteoritics and Planetary Science, 39, 531. Clayton R. N., Mayeda T. K., Olsen E. J. et al. (1991) Oxy- gen isotope studies of ordinary chondrites. Geochim. Cos- mochim. Acta, 55, 2317. Connelly J. N., Bizzarro M., Krot A. N. et al. (2012) The Abso- lute Chronology and Thermal Processing of Solids in the Solar Protoplanetary Disk. Science, 338, 651. Connolly Jr. H. C. and Desch S. J. (2004) On the origin of Chemie der the ”kleine Kugelchen” called Chondrules. Erde/Geochemistry, 64, 95. Cuzzi J. N. (2004) Blowing in the wind: III. Accretion of dust rims by chondrule-sized particles in a turbulent protoplanetary nebula. Icarus, 168, 484. Cuzzi J. N. and Alexander C. M. O. (2006) Chondrule formation in particle-rich nebular regions at least hundreds of kilometres across. Nature, 441, 483. Cuzzi J. N., Ciesla F. J., Petaev M. I. et al. (2005) Nebula Evolu- tion of Thermally Processed Solids: Reconciling Models and Meteorites. in: Chondrites and the Protoplanetary Disk, vol. 341 of Astronomical Society of the Pacific Conference Series, (edited by A. N. Krot, E. R. D. Scott, and B. Reipurth), p. 732. Cuzzi J. N., Dobrovolskis A. R., and Hogan R. C. (1996) Turbu- lence, chondrules, and planetesimals. in: Chondrules and the Protoplanetary Disk, (edited by R. H. Hewins, R. H. Jones, and E. R. D. Scott), pp. 35–43. Cuzzi J. N., Hartlep T., Weston B. et al. (2014) Turbulent Con- centration of mm-size Particles in the Protoplanetary Nebula: Scale-Dependent Multiplier Functions. in: Lunar and Plane- tary Science Conference, vol. 45 of Lunar and Planetary Sci- ence Conference, p. 2764. Cuzzi J. N. and Hogan R. C. (2003) Blowing in the wind. I. Veloc- ities of chondrule-sized particles in a turbulent protoplanetary nebula. Icarus, 164, 127. Cuzzi J. N. and Hogan R. C. (2012) Primary Accretion by Turbu- lent Concentration: The Rate of Planetesimal Formation and in: Lunar and Planetary Science the Role of Vortex Tubes. Conference, vol. 43 of Lunar and Planetary Science Confer- ence, p. 2536. Cuzzi J. N., Hogan R. C., and Bottke W. F. (2010) Towards initial mass functions for asteroids and Kuiper Belt Objects. Icarus, 208, 518. Cuzzi J. N., Hogan R. C., Paque J. M. et al. (2001) Size-selective Concentration of Chondrules and Other Small Particles in Pro- toplanetary Nebula Turbulence. Astrophys. J., 546, 496. Cuzzi J. N., Hogan R. C., and Shariff K. (2008) Toward Planetesi- mals: Dense Chondrule Clumps in the Protoplanetary Nebula. Astrophys. J., 687, 1432. Cuzzi J. N. and Weidenschilling S. J. (2006) Particle-Gas Dynam- ics and Primary Accretion. in: Meteorites and the Early Solar System II, (edited by Lauretta, D. S. & McSween, H. Y., Jr.), pp. 353–381, University of Arizona Press. Dauphas N. and Chaussidon M. (2011) A Perspective from Extinct Radionuclides on a Young Stellar Object: The Sun and Its Ac- cretion Disk. Annual Review of Earth and Planetary Sciences, 39, 351. Deharveng L., Schuller F., Anderson L. D. et al. (2010) A gallery of bubbles. The nature of the bubbles observed by Spitzer and what ATLASGAL tells us about the surrounding neutral mate- rial. Astron. Astrophys., 523, A6. Desch S. J., Morris M. A., Connolly H. C. et al. (2012) The im- portance of experiments: Constraints on chondrule formation models. Meteoritics and Planetary Science, 47, 1139. Dittrich K., Klahr H., and Johansen A. (2013) Gravoturbulent Planetesimal Formation: The Positive Effect of Long-lived Zonal Flows. Astrophys. J., 763, 117. Dodd R. T. (1976) Accretion of the ordinary chondrites. Earth and Planetary Science Letters, 30, 281. Dominik C., Blum J., Cuzzi J. N. et al. (2007) Growth of Dust as the Initial Step Toward Planet Formation. Protostars and Planets V, pp. 783–800. Dominik C. and Tielens A. G. G. M. (1997) The Physics of Dust Coagulation and the Structure of Dust Aggregates in Space. Astrophys. J., 480, 647. Doyle P. M., Krot A. N., Nagashima K. et al. (2014) Manganese- Chromium Ages of Aqueous Alteration of Unequilibrated Or- in: Lunar and Planetary Science Confer- dinary Chondrites. ence, vol. 45 of Lunar and Planetary Inst. Technical Report, p. 1726. Dra¸zkowska J., Windmark F., and Dullemond C. P. (2013) Plan- etesimal formation via sweep-up growth at the inner edge of dead zones. Astron. Astrophys., 556, A37. 18 Eaton J. K. and Fessler J. R. (1994) . Int. J. Multiph. Flow, 20, 20. Elkins-Tanton L. T., Weiss B. P., and Zuber M. T. (2011) Chon- drites as samples of differentiated planetesimals. Earth and Planetary Science Letters, 305, 1. Eugster O., Herzog G. F., Marti K. et al. (2006a) Irradiation Records, Cosmic-Ray Exposure Ages, and Transfer Times of in: Meteorites and the Early Solar System II, Meteorites. (edited by H. Y. Lauretta D. S. & McSween), pp. 829–851, University of Arizona Press. Eugster O., Herzog G. F., Marti K. et al. (2006b) Irradiation Records, Cosmic-Ray Exposure Ages, and Transfer Times of Meteorites, pp. 829–851. Fisher K. R., Tait A. W., Simon J. I. et al. (2014) Contrasting Size Distributions of Chondrules and Inclusions in Allende CV3. in: Lunar and Planetary Science Conference, vol. 45 of Lunar and Planetary Science Conference, p. 2711. Fraser W. C., Brown M. E., Morbidelli A. et al. (2014) The Abso- lute Magnitude Distribution of Kuiper Belt Objects. Astrophys. J., 782, 100. Fressin F., Torres G., Charbonneau D. et al. (2013) The False Pos- itive Rate of Kepler and the Occurrence of Planets. Astrophys. J., 766, 81. Fu R. R. and Elkins-Tanton L. T. (2014) The fate of magmas in planetesimals and the retention of primitive chondritic crusts. Earth and Planetary Science Letters, 390, 128. Fujiya W., Sugiura N., Sano Y. et al. (2013) Mn-Cr ages of dolomites in CI chondrites and the Tagish Lake ungrouped car- bonaceous chondrite. Earth and Planetary Science Letters, 362, 130. Gaidos E., Krot A. N., Williams J. P. et al. (2009) 26Al and the Formation of the Solar System from a Molecular Cloud Con- taminated by Wolf-Rayet Winds. Astrophys. J., 696, 1854. Gail H.-P., Trieloff M., Breuer D. et al. (2014) Early Thermal Evo- lution of Planetesimals and its Impact on Processing and Dat- ing of Meteoritic Material. ArXiv e-prints. Gal-Yam A., Arcavi I., Ofek E. O. et al. (2014) A Wolf-Rayet-like progenitor of SN 2013cu from spectral observations of a stellar wind. Nature, 509, 471. Garaud P., Meru F., Galvagni M. et al. (2013) From Dust to Plan- etesimals: An Improved Model for Collisional Growth in Pro- toplanetary Disks. Astrophys. J., 764, 146. Gounelle M. (2014) Aluminium-26 in the Early Solar System: A in: Lunar and Planetary Science Con- Probability Estimate. ference, vol. 45 of Lunar and Planetary Science Conference, p. 2113. Gounelle M., Chaussidon M., and Rollion-Bard C. (2013) Variable and Extreme Irradiation Conditions in the Early Solar System Inferred from the Initial Abundance of 10Be in Isheyevo CAIs. Astrophys. J. Lett., 763, L33. Gounelle M. and Meibom A. (2008) The Origin of Short-lived Ra- dionuclides and the Astrophysical Environment of Solar Sys- tem Formation. Astrophys. J., 680, 781. Gounelle M., Meibom A., Hennebelle P. et al. (2009) Supernova Propagation and Cloud Enrichment: A New Model for the Ori- gin of 60Fe in the Early Solar System. Astrophys. J. Lett., 694, L1. Gounelle M. and Meynet G. (2012) Solar system genealogy re- vealed by extinct short-lived radionuclides in meteorites. As- tron. Astrophys., 545, A4. Gounelle M., Shu F. H., Shang H. et al. (2006) The Irradiation Origin of Beryllium Radioisotopes and Other Short-lived Ra- dionuclides. Astrophys. J., 640, 1163. Grimm R. E. and McSween H. Y. (1993) Heliocentric zoning of the asteroid belt by aluminum-26 heating. Science, 259, 653. Guan Y., Huss G. R., Leshin L. A. et al. (2006) Oxygen isotope and 26Al-26Mg systematics of aluminum-rich chondrules from unequilibrated enstatite chondrites. Meteoritics and Planetary Science, 41, 33. Guttler C., Blum J., Zsom A. et al. (2010) The outcome of proto- planetary dust growth: pebbles, boulders, or planetesimals?. I. Mapping the zoo of laboratory collision experiments. Astron. Astrophys., 513, A56. Hartmann L., Calvet N., Gullbring E. et al. (1998) Accretion and the Evolution of T Tauri Disks. Astrophys. J., 495, 385. Hayashi C. (1981) Structure of the Solar Nebula, Growth and De- cay of Magnetic Fields and Effects of Magnetic and Turbu- lent Viscosities on the Nebula. Progress of Theoretical Physics Supplement, 70, 35. Henke S., Gail H.-P., Trieloff M. et al. (2013) Thermal evolu- tion model for the H chondrite asteroid-instantaneous forma- tion versus protracted accretion. Icarus, 226, 212. Hester J. J., Desch S. J., Healy K. R. et al. (2004) The Cradle of the Solar System. Science, 304, 1116. Hewins R. H., Connolly H. C., Lofgren G. E. et al. (2005) Exper- in: Chondrites imental Constraints on Chondrule Formation. and the Protoplanetary Disk, vol. 341 of Astronomical Soci- ety of the Pacific Conference Series, (edited by A. N. Krot, E. R. D. Scott, & B. Reipurth), pp. 286–316. Hezel D. C. and Palme H. (2010) The chemical relationship be- tween chondrules and matrix and the chondrule matrix com- plementarity. Earth and Planetary Science Letters, 294, 85. Hezel D. C., Russell S. S., Ross A. J. et al. (2008) Modal abun- dances of CAIs: Implications for bulk chondrite element abun- dances and fractionations. Meteoritics and Planetary Science, 43, 1879. Hogan R. C. and Cuzzi J. N. (2007) Cascade model for particle concentration and enstrophy in fully developed turbulence with mass-loading feedback. Phys. Rev. E, 75, 5, 056305. Huss G. R., Meyer B. S., Srinivasan G. et al. (2009) Stellar sources of the short-lived radionuclides in the early solar sys- tem. Geochim. Cosmochim. Acta, 73, 4922. Huss G. R., Rubin A. E., and Grossman J. N. (2006) Thermal in: Meteorites and the Early Metamorphism in Chondrites. Solar System II, (edited by Lauretta, D. S. & McSween, H. Y., Jr.), pp. 567–586, University of Arizona Press. Ida S., Guillot T., and Morbidelli A. (2008) Accretion and De- struction of Planetesimals in Turbulent Disks. Astrophys. J., 686, 1292. Jacobsen B., Yin Q.-z., Moynier F. et al. (2008) 26Al- 26Mg and 207Pb- 206Pb systematics of Allende CAIs: Canonical solar initial 26Al/ 27Al ratio reinstated. Earth and Planetary Science Letters, 272, 353. Jacquet E. (2014a) The quasi-universality of chondrule size as a constraint for chondrule formation models. Icarus, 232, 176. Jacquet E. (2014b) Transport of solids in protoplanetary disks: Comparing meteorites and astrophysical models. Comptes Rendus Geoscience, 346, 3. Jacquet E., Alard O., and Gounelle M. (2012a) Chondrule trace element geochemistry at the mineral scale. Meteoritics and Planetary Science, 47, 1695. Jacquet E., Balbus S., and Latter H. (2011) On linear dust-gas streaming instabilities in protoplanetary discs. Mon. Not. R. Astron. Soc., 415, 3591. Jacquet E., Gounelle M., and Fromang S. (2012b) On the aero- 19 dynamic redistribution of chondrite components in protoplan- etary disks. Icarus, 220, 162. Jarosewich E. (1990) Chemical analyses of meteorites - A com- pilation of stony and iron meteorite analyses. Meteoritics, 25, 323. Johansen A., Blum J., Tanaka H. et al. (2014) The multifaceted planetesimal formation process. Protostars and Planets V. Johansen A., Brauer F., Dullemond C. et al. (2008) A coagulation- fragmentation model for the turbulent growth and destruction of preplanetesimals. Astron. Astrophys., 486, 597. Johansen A., Klahr H., and Henning T. (2006) Gravoturbulent For- mation of Planetesimals. Astrophys. J., 636, 1121. Johansen A., Klahr H., and Henning T. (2011) High-resolution simulations of planetesimal formation in turbulent protoplane- tary discs. Astron. Astrophys., 529, A62. Johansen A. and Lacerda P. (2010) Prograde rotation of protoplan- ets by accretion of pebbles in a gaseous environment. Mon. Not. R. Astron. Soc., 404, 475. Johansen A., Mac Low M.-M., Lacerda P. et al. (2015) Growth of asteroids, planetary embryos and Kuiper belt objects by chon- drule accretion. Science Advances, 1, e1500109. Johansen A., Oishi J. S., Mac Low M.-M. et al. (2007) Rapid plan- etesimal formation in turbulent circumstellar disks. Nature, 448, 1022. Johansen A. and Youdin A. (2007) Protoplanetary Disk Turbu- lence Driven by the Streaming Instability: Nonlinear Satura- tion and Particle Concentration. Astrophys. J., 662, 627. Johansen A., Youdin A., and Klahr H. (2009a) Zonal Flows and Long-lived Axisymmetric Pressure Bumps in Magnetorota- tional Turbulence. Astrophys. J., 697, 1269. Johansen A., Youdin A., and Mac Low M.-M. (2009b) Particle Clumping and Planetesimal Formation Depend Strongly on Metallicity. Astrophys. J. Lett., 704, L75. Johansen A., Youdin A. N., and Lithwick Y. (2012) Adding parti- cle collisions to the formation of asteroids and Kuiper belt ob- jects via streaming instabilities. Astron. Astrophys., 537, A125. Jones R. H. (2012) Petrographic constraints on the diversity of chondrule reservoirs in the protoplanetary disk. Meteoritics and Planetary Science, 47, 1176. Kastner J. H. and Myers P. C. (1994) An observational estimate of the probability of encounters between mass-losing evolved stars and molecular clouds. Astrophys. J., 421, 605. Kato M. T., Fujimoto M., and Ida S. (2012) Planetesimal For- mation at the Boundary between Steady Super/Sub-Keplerian Flow Created by Inhomogeneous Growth of Magnetorotational Instability. Astrophys. J., 747, 11. Keil K., Stoeffler D., Love S. G. et al. (1997) Constraints on the role of impact heating and melting in asteroids. Meteoritics and Planetary Science, 32, 349. Kita N. T. and Ushikubo T. (2012) Evolution of protoplanetary disk inferred from 26Al chronology of individual chondrules. Meteoritics and Planetary Science, 47, 1108. Kita N. T., Yin Q.-Z., MacPherson G. J. et al. (2013) 26Al-26Mg isotope systematics of the first solids in the early solar system. Meteoritics and Planetary Science, 48, 1383. Klahr H. H. and Bodenheimer P. (2003) Turbulence in Accretion Disks: Vorticity Generation and Angular Momentum Transport via the Global Baroclinic Instability. Astrophys. J., 582, 869. Kretke K. A. and Lin D. N. C. (2007) Grain Retention and Forma- tion of Planetesimals near the Snow Line in MRI-driven Tur- bulent Protoplanetary Disks. Astrophys. J. Lett., 664, L55. Krot A., Petaev M., Russell S. S. et al. (2004) Amoeboid olivine aggregates and related objects in carbonaceous chondrites: records of nebular and asteroid processes. Chemie der Erde / Geochemistry, 64, 185. Krot A. N., Amelin Y., Bland P. et al. (2009) Origin and chronol- ogy of chondritic components: A review. Geochim. Cos- mochim. Acta, 73, 4963. Kruijer T. S., Sprung P., Kleine T. et al. (2012) Hf-W chronometry of core formation in planetesimals inferred from weakly irra- diated iron meteorites. Geochimica et Cosmochimica Acta, 99, 287. Kuebler K. E., McSween H. Y., Carlson W. D. et al. (1999) Sizes and Masses of Chondrules and Metal-Troilite Grains in Ordi- nary Chondrites: Possible Implications for Nebular Sorting. Icarus, 141, 96. Lambrechts M. and Johansen A. (2012) Rapid growth of gas-giant cores by pebble accretion. Astron. Astrophys., 544, A32. Lambrechts M., Johansen A., and Morbidelli A. (2014) Separat- ing gas-giant and ice-giant planets by halting pebble accretion. Astron. Astrophys., 572, A35. Larsen K. K., Trinquier A., Paton C. et al. (2011) Evidence for Magnesium Isotope Heterogeneity in the Solar Protoplanetary Disk. Astrophys. J. Lett., 735, L37. Laughlin G., Steinacker A., and Adams F. C. (2004) Type I Plane- tary Migration with MHD Turbulence. Astrophys. J., 608, 489. Lauretta D. S., Nagahara H., and Alexander C. M. O. (2006) Petrology and Origin of Ferromagnesian Silicate Chondrules, pp. 431–459. Lee T. (1978) A local proton irradiation model for isotopic anoma- lies in the solar system. Astrophys. J., 224, 217. Lee T., Papanastassiou D. A., and Wasserburg G. J. (1976) Demonstration of Mg-26 excess in Allende and evidence for Al-26. Geophys. Res. Lett., 3, 41. Lee T., Shu F. H., Shang H. et al. (1998) Protostellar Cosmic Rays and Extinct Radioactivities in Meteorites. Astrophys. J., 506, 898. Lesur G. and Papaloizou J. C. B. (2010) The subcritical baroclinic instability in local accretion disc models. Astronomy & Astro- physics, 513, A60. Libourel G. and Krot A. N. (2007) Evidence for the presence of planetesimal material among the precursors of magnesian chondrules of nebular origin. Earth and Planetary Science Let- ters, 254, 1. Liu M.-C., Chaussidon M., Gopel C. et al. (2012) A heterogeneous solar nebula as sampled by CM hibonite grains. Earth and Planetary Science Letters, 327, 75. Looney L. W., Tobin J. J., and Fields B. D. (2006) Radioactive Probes of the Supernova-contaminated Solar Nebula: Evidence that the Sun Was Born in a Cluster. Astrophys. J., 652, 1755. Lugaro M., Doherty C. L., Karakas A. I. et al. (2012) Short-lived radioactivity in the early solar system: The Super-AGB star hypothesis. Meteoritics and Planetary Science, 47, 1998. Lyra W., Johansen A., Klahr H. et al. (2008) Embryos grown in the dead zone. Assembling the first protoplanetary cores in low mass self-gravitating circumstellar disks of gas and solids. As- tron. Astrophys., 491, L41. Lyra W., Johansen A., Zsom A. et al. (2009) Planet formation bursts at the borders of the dead zone in 2D numerical sim- ulations of circumstellar disks. Astron. Astrophys., 497, 869. MacPherson G. J. (2005) Calcium-aluminum-rich inclusions in in: Meteorites, Comets and Planets: chondritic meteorites. Treatise on Geochemistry, Volume 1, (edited by Davis, A. M., Holland, H. D., & Turekian, K. K.), chap. 1.08, pp. 201–246, 20 Elsevier B. MacPherson G. J., Davis A. M., and Zinner E. K. (2014) Distri- bution of 26Al in the Early Solar System: A 2014 Reappraisal. in: Lunar and Planetary Science Conference, vol. 45 of Lunar and Planetary Science Conference, p. 2134. MacPherson G. J., Hashimoto A., and Grossman L. (1985) Accretionary rims on inclusions in the Allende meteorite. Geochim. Cosmochim. Acta, 49, 2267. Makide K., Nagashima K., Krot A. N. et al. (2013) Heterogeneous distribution of 26Al at the birth of the Solar System: Evidence from corundum-bearing refractory inclusions in carbonaceous chondrites. Geochim. Cosmochim. Acta, 110, 190. Martin R. G. and Livio M. (2012) On the evolution of the snow line in protoplanetary discs. Mon. Not. R. Astron. Soc., 425, L6. Martin R. G. and Livio M. (2013) On the evolution of the snow line in protoplanetary discs - II. Analytic approximations. Mon. Not. R. Astron. Soc., 434, 633. Metzler K. (2012) Ultrarapid chondrite formation by hot chon- drule accretion? Evidence from unequilibrated ordinary chon- drites. Meteoritics and Planetary Science, 47, 2193. Metzler K., Bischoff A., and Stoeffler D. (1992) Accretionary dust mantles in CM chondrites - Evidence for solar nebula pro- cesses. Geochim. Cosmochim. Acta, 56, 2873. Mizuno H., Markiewicz W. J., and Voelk H. J. (1988) Grain growth in turbulent protoplanetary accretion disks. Astron. As- trophys., 195, 183. Monnereau M., Toplis M. J., Baratoux D. et al. (2013) Thermal history of the H-chondrite parent body: Implications for meta- morphic grade and accretionary time-scales. Geochim. Cos- mochim. Acta, 119, 302. Morbidelli A., Bottke W. F., Nesvorn´y D. et al. (2009) Asteroids were born big. Icarus, 204, 558. Morbidelli A., Chambers J., Lunine J. I. et al. (2000) Source re- gions and time scales for the delivery of water to Earth. Mete- oritics and Planetary Science, 35, 1309. Morfill G. E., Durisen R. H., and Turner G. W. (1998) NOTE: an Accretion Rim Constraint on Chondrule Formation Theories. Icarus, 134, 180. Mostefaoui S., Lugmair G. W., and Hoppe P. (2005) 60Fe: A Heat Source for Planetary Differentiation from a Nearby Supernova Explosion. Astrophys. J., 625, 271. Moth´e-Diniz T., Carvano J. M., Bus S. J. et al. (2008) Miner- alogical analysis of the Eos family from near-infrared spectra. Icarus, 195, 277. Moth´e-Diniz T. and Nesvorn´y D. (2008) Visible spectroscopy of extremely young asteroid families. Astron. Astrophys., 486, L9. Moth´e-Diniz T., Roig F., and Carvano J. M. (2005) Reanaly- sis of asteroid families structure through visible spectroscopy. Icarus, 174, 54. Nakagawa Y., Sekiya M., and Hayashi C. (1986) Settling and growth of dust particles in a laminar phase of a low-mass solar nebula. Icarus, 67, 375. Nelson R. P. and Gressel O. (2010) On the dynamics of planetesi- mals embedded in turbulent protoplanetary discs. Mon. Not. R. Astron. Soc., 409, 639. Nelson R. P., Gressel O., and Umurhan O. M. (2013) Linear and non-linear evolution of the vertical shear instability in accretion discs. Monthly Notices of the Royal Astronomical Society, 435, 2610. Nelson R. P. and Papaloizou J. C. B. (2004) The interaction of giant planets with a disc with MHD turbulence - IV. Migration rates of embedded protoplanets. Mon. Not. R. Astron. Soc., 350, 849. Nelson V. E. and Rubin A. E. (2002) Size-frequency distributions of chondrules and chondrule fragments in LL3 chondrites: Im- plications for parent-body fragmentation of chondrules. Mete- oritics and Planetary Science, 37, 1361. O’Brien D. P., Walsh K. J., Morbidelli A. et al. (2014) Water de- livery and giant impacts in the “Grand Tack” scenario. Icarus, 239, 74. Oishi J. S., Mac Low M.-M., and Menou K. (2007) Turbulent Torques on Protoplanets in a Dead Zone. Astrophys. J., 670, 805. Okuzumi S., Tanaka H., Kobayashi H. et al. (2012) Rapid Coag- ulation of Porous Dust Aggregates outside the Snow Line: A Pathway to Successful Icy Planetesimal Formation. Astrophys. J., 752, 106. Ormel C. W. and Cuzzi J. N. (2007) Closed-form expressions for particle relative velocities induced by turbulence. Astron. As- trophys., 466, 413. Ormel C. W., Cuzzi J. N., and Tielens A. G. G. M. (2008) Co- Accretion of Chondrules and Dust in the Solar Nebula. Astro- phys. J., 679, 1588. Ormel C. W. and Klahr H. H. (2010) The effect of gas drag on the growth of protoplanets. Analytical expressions for the accre- tion of small bodies in laminar disks. Astron. Astrophys., 520, A43. Ormel C. W. and Okuzumi S. (2013) The Fate of Planetesimals in Turbulent Disks with Dead Zones. II. Limits on the Viability of Runaway Accretion. Astrophys. J., 771, 44. Ouellette N., Desch S. J., and Hester J. J. (2007) Interaction of Su- pernova Ejecta with Nearby Protoplanetary Disks. Astrophys. J., 662, 1268. Palme H. and Jones A. (2005) Solar system abundances of the el- ements. in: Meteorites, Comets and Planets: Treatise on Geo- chemistry, Volume 1, (edited by A. M. Davis, H. D. Holland, and K. K. Turekian), pp. 41–60, Elsevier. Pan L., Desch S. J., Scannapieco E. et al. (2012) Mixing of Clumpy Supernova Ejecta into Molecular Clouds. Astrophys. J., 756, 102. Pan L., Padoan P., Scalo J. et al. (2011) Turbulent Clustering of Protoplanetary Dust and Planetesimal Formation. Astrophys. J., 740, 6. Pan M. and Sari R. (2005) Shaping the Kuiper belt size distribution by shattering large but strengthless bodies. Icarus, 173, 342. Quitt´e G., Halliday A. N., Meyer B. S. et al. (2007) Correlated Iron 60, Nickel 62, and Zirconium 96 in Refractory Inclusions and the Origin of the Solar System. Astrophys. J., 655, 678. Quitt´e G., Latkoczy C., Schonbachler M. et al. (2011) 60Fe- 60Ni systematics in the eucrite parent body: A case study of Bou- vante and Juvinas. Geochim. Cosmochim. Acta, 75, 7698. Raymond S. N., Quinn T., and Lunine J. I. (2004) Making other earths: dynamical simulations of terrestrial planet formation and water delivery. Icarus, 168, 1. Ros K. and Johansen A. (2013) Ice condensation as a planet for- mation mechanism. Astron. Astrophys., 552, A137. Rubin A. E. (2000) Petrologic, geochemical and experimental constraints on models of chondrule formation. Earth Science Reviews, 50, 3. Rubin A. E. (2005) Relationships among intrinsic properties of ordinary chondrites: Oxidation state, bulk chemistry, oxygen- isotopic composition, petrologic type, and chondrule size. 21 Geochim. Cosmochim. Acta, 69, 4907. Rubin A. E. (2011) Origin of the differences in refractory- lithophile-element abundances among chondrite groups. Icarus, 213, 547. Rubin A. E. and Brearley A. J. (1996) A Critical Evaluation of the Evidence for Hot Accretion. Icarus, 124, 86. Schneider D. M., Akridge D. G., and Sears D. W. G. (1998) Size Distribution of Metal Grains and Chondrules in Enstatite Chondrites. Meteoritics and Planetary Science Supplement, 33, 136. Schrapler R., Blum J., Seizinger A. et al. (2012) The Physics of Protoplanetesimal Dust Agglomerates. VII. The Low-velocity Collision Behavior of Large Dust Agglomerates. Astrophys. J., 758, 35. Scott E. R. D. and Krot A. N. (2003) Chondrites and their Com- ponents. Treatise on Geochemistry, 1, 143. Scott E. R. D., Krot T. V., Goldstein J. I. et al. (2014) Thermal and impact history of the H chondrite parent asteroid during meta- morphism: Constraints from metallic Fe-Ni. Geochim. Cos- mochim. Acta, 136, 13. Scott E. R. D. and Rajan R. S. (1981) Metallic minerals, thermal histories and parent bodies of some xenolithic, ordinary chon- drite meteorites. Geochim. Cosmochim. Acta, 45, 53. Scott E. R. D., Rubin A. E., Taylor G. J. et al. (1984) Matrix ma- terial in type 3 chondrites - Occurrence, heterogeneity and re- lationship with chondrules. Geochim. Cosmochim. Acta, 48, 1741. Sekiya M. (1983) Gravitational instabilities in a dust-gas layer and formation of planetesimals in the solar nebula. Progress of Theoretical Physics, 69, 1116. Setoh M., Hiraoka K., Nakamura A. M. et al. (2007) Collisional disruption of porous sintered glass beads at low impact veloci- ties. Advances in Space Research, 40, 252. Shukolyukov A. and Lugmair G. W. (1993) Live Iron-60 in the early solar system. Science, 259, 1138. Simon J. B., Beckwith K., and Armitage P. J. (2012) Emer- gent mesoscale phenomena in magnetized accretion disc tur- bulence. Mon. Not. R. Astron. Soc., 422, 2685. Sirono S.-i. (2011) Planetesimal Formation Induced by Sintering. Astrophys. J. Lett., 733, L41. Sonett C. P. and Colburn D. S. (1968) Electrical Heating of Mete- orite Parent Bodies and Planets by Dynamo Induction from a Pre-main Sequence T Tauri “Solar Wind”. Nature, 219, 924. Squires K. D. and Eaton J. K. (1990) Particle response and turbu- lence modification in isotropic turbulence. Physics of Fluids, 2, 1191. Squires K. D. and Eaton J. K. (1991) Preferential concentration of particles by turbulence. Physics of Fluids, 3, 1169. Stewart S. T. and Leinhardt Z. M. (2009) Velocity-Dependent Catastrophic Disruption Criteria for Planetesimals. Astrophys. J. Lett., 691, L133. Tachibana S. and Huss G. R. (2003) The Initial Abundance of 60Fe in the Solar System. Astrophys. J. Lett., 588, L41. Tachibana S., Huss G. R., Kita N. T. et al. (2006) 60Fe in Chon- drites: Debris from a Nearby Supernova in the Early Solar Sys- tem? Astrophys. J. Lett., 639, L87. Tang H. and Dauphas N. (2012) Abundance, distribution, and ori- gin of 60Fe in the solar protoplanetary disk. Earth Planet. Sci. Lett., 359, 248. Tang X. and Chevalier R. A. (2014) Gamma-Ray Emission from Supernova Remnant Interactions with Molecular Clumps. As- trophys. J. Lett., 784, L35. Tatischeff V., Duprat J., and de S´er´eville N. (2010) A Runaway Wolf-Rayet Star as the Origin of 26Al in the Early Solar Sys- tem. Astrophys. J. Lett., 714, L26. Taylor G. J., Maggiore P., Scott E. R. D. et al. (1987) Original structures, and fragmentation and reassembly histories of as- teroids - Evidence from meteorites. Icarus, 69, 1. Telus M., Huss G. R., Ogliore R. C. et al. (2012) Recalculation of data for short-lived radionuclide systems using less-biased ratio estimation. Meteoritics and Planetary Science, 47, 2013. Testi L., Birnstiel T., Ricci L. et al. (2014) Dust Evolution in Pro- toplanetary Disks. ArXiv e-prints. Trieloff M., Jessberger E. K., Herrwerth I. et al. (2003) Structure and thermal history of the H-chondrite parent asteroid revealed by thermochronometry. Nature, 422, 502. Trinquier A., Elliott T., Ulfbeck D. et al. (2009) Origin of Nucle- osynthetic Isotope Heterogeneity in the Solar Protoplanetary Disk. Science, 324, 374. Turner N. J., Fromang S., Gammie C. et al. (2014) Transport and Accretion in Planet-Forming Disks. ArXiv e-prints. Urey H. C. (1955) The Cosmic Abundances of Potassium, Ura- nium, and Thorium and the Heat Balances of the Earth, the Moon, and Mars. Proceedings of the National Academy of Sci- ence, 41, 127. Vasileiadis A., Nordlund A., and Bizzarro M. (2013) Abundance of 26Al and 60Fe in Evolving Giant Molecular Clouds. Astro- phys. J. Lett., 769, L8. Vernazza P., Zanda B., Binzel R. P. et al. (2014) Multiple and Fast: The Accretion of Ordinary Chondrite Parent Bodies. ArXiv e- prints. Villeneuve J., Chaussidon M., and Libourel G. (2012) Lack of re- lationship between aluminum-26 ages of chondrules and their mineralogical and chemical compositions. Comptes Rendus Geoscience, 344, 423. Voelk H. J., Jones F. C., Morfill G. E. et al. (1980) Collisions be- tween grains in a turbulent gas. Astron. Astrophys., 85, 316. Wada K., Tanaka H., Okuzumi S. et al. (2013) Growth efficiency of dust aggregates through collisions with high mass ratios. As- tron. Astrophys., 559, A62. Wada K., Tanaka H., Suyama T. et al. (2009) Collisional Growth Conditions for Dust Aggregates. Astrophys. J., 702, 1490. Walsh K. J., Morbidelli A., Raymond S. N. et al. (2011) A low mass for Mars from Jupiter’s early gas-driven migration. Na- ture, 475, 206. Wang L.-P. and Maxey M. R. (1993) Settling velocity and concen- tration distribution of heavy particles in homogeneous isotropic turbulence. Journal of Fluid Mechanics, 256, 27. Wasserburg G. J., Wimpenny J., and Yin Q.-Z. (2012) Mg isotopic heterogeneity, Al-Mg isochrons, and canonical 26Al/27Al in the early solar system. Meteoritics and Planetary Science, 47, 1980. Wasson J. T., Isa J., and Rubin A. E. (2013) Compositional and petrographic similarities of CV and CK chondrites: A sin- gle group with variations in textures and volatile concentra- tions attributable to impact heating, crushing and oxidation. Geochim. Cosmochim. Acta, 108, 45. Weidenschilling S. J. (1977a) Aerodynamics of solid bodies in the solar nebula. Monthly Notices of the Royal Astronomical Soci- ety, 180, 57. Weidenschilling S. J. (1977b) Aerodynamics of solid bodies in the solar nebula. Mon. Not. R. Astron. Soc., 180, 57. Weidenschilling S. J. (2011) Initial sizes of planetesimals and ac- cretion of the asteroids. Icarus, 214, 671. 22 Weidling R., Guttler C., and Blum J. (2012) Free collisions in a microgravity many-particle experiment. I. Dust aggregate sticking at low velocities. Icarus, 218, 688. Weisberg M. K., McCoy T. J., and Krot A. N. (2006) Systemat- ics and Evaluation of Meteorite Classification. in: Meteorites and the Early Solar System II, (edited by Lauretta, D. S. & McSween, H. Y., Jr.), pp. 19–52, University of Arizona Press. Weiss B. P. and Elkins-Tanton L. T. (2013) Differentiated Plan- etesimals and the Parent Bodies of Chondrites. Annual Review of Earth and Planetary Sciences, 41, 529. Weiss B. P., Gattacceca J., Stanley S. et al. (2010) Paleomagnetic Records of Meteorites and Early Planetesimal Differentiation. Space Sci. Rev., 152, 341. Whattam S. A., Hewins R. H., Cohen B. A. et al. (2008) Gra- noblastic olivine aggregates in magnesian chondrules: Plan- etesimal fragments or thermally annealed solar nebula conden- sates? Earth and Planetary Science Letters, 269, 200. Whipple F. L. (1972) On certain aerodynamic processes for as- in: From Plasma to Planet, (edited by teroids and comets. A. Elvius), p. 211. Williams J. P. and Gaidos E. (2007) On the Likelihood of Super- nova Enrichment of Protoplanetary Disks. Astrophys. J. Lett., 663, L33. Windmark F., Birnstiel T., Guttler C. et al. (2012a) Planetesimal formation by sweep-up: how the bouncing barrier can be ben- eficial to growth. Astron. Astrophys., 540, A73. Windmark F., Birnstiel T., Ormel C. W. et al. (2012b) Breaking through: The effects of a velocity distribution on barriers to dust growth. Astron. Astrophys., 544, L16. Wood J. A. (2005) The Chondrite Types and Their Origins. in: Chondrites and the Protoplanetary Disk, vol. 341 of Astronom- ical Society of the Pacific Conference Series, (edited by A. N. Krot, E. R. D. Scott, and B. Reipurth), p. 953. Woosley S. E. and Heger A. (2007) Nucleosynthesis and remnants Wurm G., Paraskov G., and Krauss O. (2005) Growth of planetes- in massive stars of solar metallicity. Phys. Rep., 442, 269. imals by impacts at ∼25 m/s. Icarus, 178, 253. Xie J.-W., Payne M. J., Th´ebault P. et al. (2010) From Dust to Planetesimal: The Snowball Phase? Astrophys. J., 724, 1153. Yang C.-C. and Johansen A. (2014) On the Feeding Zone of Plan- etesimal Formation by the Streaming Instability. Astrophys. J., 792, 86. Yang C.-C., Mac Low M.-M., and Menou K. (2012) Planetesimal and Protoplanet Dynamics in a Turbulent Protoplanetary Disk: Ideal Stratified Disks. Astrophys. J., 748, 79. Youdin A. and Johansen A. (2007) Protoplanetary Disk Turbu- lence Driven by the Streaming Instability: Linear Evolution and Numerical Methods. Astrophys. J., 662, 613. Youdin A. N. and Goodman J. (2005) Streaming Instabilities in Protoplanetary Disks. Astrophys. J., 620, 459. Youdin A. N. and Shu F. H. (2002) Planetesimal Formation by Gravitational Instability. Astrophys. J., 580, 494. Young E. D. (2014) Inheritance of solar short- and long-lived ra- dionuclides from molecular clouds and the unexceptional na- ture of the solar system. Earth and Planetary Science Letters, 392, 16. Zanda B., Humayun M., and Hewins R. H. (2012) Chemical Com- position of Matrix and Chondrules in Carbonaceous Chon- drites: Implications for Disk Transport. in: Lunar and Plane- tary Institute Science Conference Abstracts, vol. 43, p. 2413. Zinner E. and Gopel C. (2002) Aluminum-26 in H4 chondrites: Implications for its production and its usefulness as a fine-scale chronometer for early solar system events. Meteoritics and Planetary Science, 37, 1001. Zsom A., Ormel C. W., Guttler C. et al. (2010) The outcome of protoplanetary dust growth: pebbles, boulders, or planetesi- mals? II. Introducing the bouncing barrier. Astron. Astrophys., 513, A57. This 2-column preprint was prepared with the AAS LATEX macros v5.2. 23
1108.5780
1
1108
2011-08-29T23:42:52
EURONEAR - Recovery, Follow-up and Discovery of NEAs and MBAs using Large Field 1-2m Telescopes
[ "astro-ph.EP" ]
We report on the follow-up and recovery of 100 program NEAs, PHAs and VIs using the ESO/MPG 2.2m, Swope 1m and INT 2.5m telescopes equipped with large field cameras. The 127 fields observed during 11 nights covered 29 square degrees. Using these data, we present the incidental survey work which includes 558 known MBAs and 628 unknown moving objects mostly consistent with MBAs from which 58 objects became official discoveries. We planned the runs using six criteria and four servers which focus mostly on faint and poorly observed objects in need of confirmation, follow-up and recovery. We followed 62 faint NEAs within one month after discovery and we recovered 10 faint NEAs having big uncertainties at their second or later opposition. Using the INT we eliminated 4 PHA candidates and VIs. We observed in total 1,286 moving objects and we reported more than 10,000 positions. All data were reduced by the members of our network in a team effort, and reported promptly to the MPC. The positions of the program NEAs were published in 27 MPC and MPEC references and used to improve their orbits. The O-C residuals for known MBAs and program NEAs are smallest for the ESO/MPG and Swope and about four times larger for the INT whose field is more distorted. The incidental survey allowed us to study statistics of the MBA and NEA populations observable today with 1--2m facilities. We calculate preliminary orbits for all unknown objects, classifying them as official discoveries, later identifications and unknown outstanding objects. The orbital elements a, e, i calculated by FIND_ORB software for the official discoveries and later identified objects are very similar with the published elements which take into account longer observational arcs; thus preliminary orbits were used in statistics for the whole unknown dataset. (CONTINUED)
astro-ph.EP
astro-ph
Appendix A. Unknown objects observed during out runs at ESO/MPG, Swope and INT. Table A.1: The asteroids officialy discovered at ESO/MPG (according to MPC DISCSTATUS Jan 2011). In the first line we give orbital elements calculated with FIND ORB and observational data from our run, while in the second line we include MPC orbital data derived from all available observations. Acronym Designation a e i MOID H pos arc rms R β ǫ µ VBSO004 2008 EB98 VBSO008 2008 EF155 VBSO036 2008 EG162 VBSO052 2008 EV168 VBSO059 2008 EF151 VBTO002 2008 ED145 VBTO005 2008 ED144 VBTO008 2008 EC144 VBTO011 2008 EY131 VBTO012 2008 EB145 VBTO013 2008 EA145 VBTO015 2008 EE145 VBTO021 2008 EN145 VBTO022 2008 EK145 VBTO023 2008 EH155 VBTO024 2008 EJ145 VBTO025 2006 WO37 VBTO029 2008 EH145 VBTO043 2008 EL145 2.94 2.90 2.73 2.86 2.79 2.49 2.41 2.36 2.33 2.25 2.60 3.43 3.22 2.23 2.23 2.54 2.42 3.08 2.79 3.23 3.12 2.49 2.24 2.38 4.11 2.99 2.25 2.59 2.31 2.56 2.99 3.02 2.47 0.12 0.16 0.05 0.21 0.18 0.06 0.16 0.01 0.10 0.14 0.25 0.21 0.07 0.17 0.15 0.21 0.07 0.22 0.20 0.07 0.07 0.21 0.17 0.38 0.48 0.18 0.09 0.16 0.09 0.08 0.13 0.14 0.07 7.7 7.1 11.4 1.9 2.3 3.3 2.7 7.4 7.5 4.7 5.1 9.7 10.5 7.8 8.1 3.8 4.4 8.7 5.6 14.6 9.2 21.6 6.6 8.3 10.7 11.3 3.8 17.0 4.4 6.9 11.1 11.1 6.1 1.58 1.61 1.28 1.33 1.33 0.93 1.76 0.86 1.03 1.39 1.22 1.92 1.02 0.49 1.17 1.05 1.11 1.60 1.29 16.8 17.2 17.3 18.5 18.2 17.5 18.2 17.7 18.1 18.7 18.5 16.3 16.3 17.4 17.7 17.7 17.4 15.8 17.9 15.9 16.7 16.2 17.5 16.9 15.3 15.6 17.4 16.8 18.5 17.2 17.4 17.6 17.9 VBTO055 2008 EO145 2.43 0.05 3.0 1.31 17.7 VBTO063 2008 EP145 VBTU011 2008 EA84 2.08 2.29 2.90 0.02 0.20 0.41 6.7 6.9 14.5 1.02 0.71 18.3 18.4 15.2 18 25 16 7 15 8 16 8 24 11 23 30 40 14 30 30 66 22 14 26 22 16 51 30 22 29 30 72 16 53 16 24 16 16 15 37 4 2d 10d 2d 1d 23d 1d 24d 1d 36d 1d 16d 3d 16d 1d 3d 3d 3y 2d 1d 16d 2d 2d 8y 3d 2d 16d 3d 5y 2d 5y 2d 16d 2d 0.07 20.0 0.14 20.5 0.24 20.7 0.10 20.1 0.06 20.3 0.18 20.3 0.09 20.5 0.37 20.7 0.12 20.2 0.26 21.0 0.09 20.4 0.37 20.6 0.10 19.9 0.13 20.4 0.13 20.3 0.16 20.4 0.20 20.6 0.16 20.6 0.26 20.9 2d 0.08 20.4 2d 16d 1d 0.17 20.4 0.05 20.7 10 11 2 2 2 -3 -3 -3 -4 -4 -4 -3 -3 -3 -3 -3 -3 -4 -4 -3 -3 -6 176 0.54 176 0.60 173 0.54 163 0.57 163 0.62 161 0.54 162 0.46 162 0.63 162 0.55 162 0.45 162 0.49 161 0.46 162 0.64 162 0.59 162 0.46 162 0.59 162 0.55 162 0.52 162 0.55 162 0.53 162 0.64 192 0.50 1 Table A.1 (continued) – Asteroids officially discovered at ESO/MPG. Acronym Designation a e i MOID H pos arc σ R β ǫ µ VBTU012 2008 EX154 VBTU013 2008 EU154 2.66 2.77 2.23 0.35 0.06 0.05 VBTU014 2008 EY83 2.06 0.16 VBTU015 2008 EZ83 VBTU016 2008 EW154 3.06 3.10 2.04 0.18 0.08 0.31 7.0 5.1 5.2 4.2 5.2 6.2 5.0 0.74 1.12 15.7 16.8 19.0 0.74 20.4 1.52 0.41 17.2 17.0 17.7 VBTU017 2008 EV154 2.55 0.02 15.4 1.49 17.9 VBTU020 2008 EC145 2.14 0.13 VBTU049 2008 EL154 2.61 0.22 VBTU052 2008 EE155 VBTU054 2008 ED155 VBTU056 2008 EK154 VBTU057 2008 EC155 VBTU058 2008 EA155 VBTU059 2008 EJ154 VBTU060 2008 EB155 VBTU061 2008 EZ154 2.56 2.67 3.05 2.99 2.81 3.06 3.04 2.60 2.60 2.86 2.89 3.11 3.20 2.25 0.23 0.29 0.15 0.02 0.31 0.05 0.09 0.02 0.02 0.31 0.08 0.14 0.05 0.09 VBTU062 2008 EY154 3.03 0.19 VBTU063 2008 EK155 2.27 0.23 VBTU064 2008 EJ155 2.20 0.14 VBTU074 2008 EG155 2.59 0.05 VBTU089 2008 EL155 VBTU090 2008 EA98 VBTU101 2008 EX152 VBTU104 2008 EH168 VBTU107 2008 EF145 2.85 2.86 2.48 2.53 2.62 2.62 2.02 2.67 0.04 0.09 0.10 0.15 0.16 0.16 0.24 0.03 1.8 9.6 6.6 5.8 3.3 3.7 5.4 9.3 10.3 7.8 7.9 16.9 11.5 9.4 7.7 3.0 8.9 2.8 4.2 2.4 9.3 8.0 4.7 15.4 14.0 6.6 2.6 4.3 0.87 19.4 1.04 19.5 0.97 1.61 1.93 1.93 1.56 1.01 1.66 1.06 17.0 17.3 17.9 17.5 16.4 17.6 17.5 17.9 17.9 16.3 17.0 16.5 17.0 19.3 1.46 17.2 0.75 17.7 0.89 19.6 1.48 18.5 1.75 1.24 1.16 1.21 1.60 17.9 18.3 19.1 16.0 16.4 18.1 20.4 17.5 2 10 32 9 5 9 13 10 10 6 12 12 17 18 14 22 15 13 22 22 24 24 23 23 10 15 12 12 11 11 15 3 14 44 9 17 17 2d 10y 2d 0.09 20.6 0.27 21.0 1d 0.31 21.1 2d 14d 2d 0.12 20.2 0.13 20.9 2d 1d 1d 2d 3d 2d 2d 3d 2d 3d 3d 3d 3d 3d 3d 3d 1d 2d 1d 1d 2d 2d 9d 1d 0.21 21.1 0.32 20.8 0.16 21.2 0.18 21.2 0.30 21.0 0.26 21.2 0.33 21.3 0.17 21.1 0.17 21.0 0.24 21.3 0.13 21.1 0.16 20.8 0.25 21.2 0.15 20.8 0.13 21.4 0.21 21.4 0.10 21.4 -6 -6 -6 -6 -6 -6 -3 -4 -4 -4 -4 -4 -4 -4 -4 -4 -4 -4 -4 -4 -4 -4 192 0.53 192 0.62 192 0.65 192 0.50 192 0.70 192 0.64 161 0.55 189 0.65 189 0.58 189 0.51 189 0.52 189 0.53 189 0.60 189 0.54 189 0.49 189 0.67 189 0.55 188 0.65 188 0.66 189 0.57 189 0.55 189 0.61 1d 3y 1d 34d 2d 0.14 20.4 9 135 0.25 0.29 20.7 0.30 20.8 -3 -4 161 0.50 161 0.50 Table A.1 (continued) – Asteroids officially discovered at ESO/MPG. Acronym Designation a e i MOID H pos arc σ R β ǫ µ VBTU113 2008 EG144 2.31 0.38 VBTU124 2008 EJ168 VBTU128 2008 EH144 2.80 2.39 3.03 0.04 0.25 0.61 5.1 2.8 1.9 4.4 0.43 16.1 1.68 0.19 17.0 19.2 17.3 VBTU199 2008 EL144 1.77 0.16 14.8 0.56 18.5 VBTU202 2008 EK144 2.52 0.03 9.2 1.45 17.1 VBTU203 2008 EN144 1.84 0.05 23.5 0.75 18.4 VBTU204 2008 EM144 2.79 0.18 21.6 1.28 16.0 VBTU207 2008 EW152 VBTU208 2008 EE157 VBTU213 2008 EQ144 VBTU224 2008 EW144 2.09 2.80 3.12 2.19 2.40 2.73 2.34 0.17 0.01 0.07 0.37 0.20 0.03 0.10 2.1 3.9 2.9 2.4 3.8 9.1 5.8 0.76 1.92 0.93 1.12 18.2 16.6 16.8 20.8 18.5 17.2 17.9 13 21 21 19 21 21 14 16 13 63 5 13 4 11 5 1d 0.12 20.4 0 163 0.59 3d 30d 3d 2d 2d 2d 2d 1d 8y 1d 24d 1d 16d 1d 0.27 20.9 -1 163 0.49 0.14 20.6 0.17 20.7 0.21 21.0 0.26 20.8 0.23 21.0 0.14 20.9 0.06 20.4 0.07 19.7 0 1 0 0 0 0 1 1 163 0.59 132 0.46 132 0.18 132 0.65 132 0.27 131 0.13 171 0.51 172 0.59 0.15 20.1 10 162 0.55 3 Table A.2: Later identification of unknown asteroids observed at ESO/MPG. First line represent our data and second line reffers to updated MPC data. Acronym Designation a e i MOID H pos arc rms R β ǫ µ VB001 2008 FG67 VB014 2004 RZ319 VB018 2008 FZ110 VB024 2009 TG3 VB027 2005 MW44 VB029 2007 AY28 VB036 2010 VT31 VB037 2001 TB234 VB039 2005 SA133 VB045 2008 JF20 VB047 2008 GD44 VB057 2005 TB197 VBOP008 2008 EF144 VBOP013 2008 EM90 VBOP017 2008 EB144 VBOP023 2008 EJ144 VBOP024 2008 EL163 VBSO007 2008 EG168 VBSO024 2008 DA83 VBSO063 2008 ER144 VBSO066 2008 ES144 VBSO068 2008 EZ95 VBTO001 2008 EJ20 VBTO028 2008 EE144 VBTO052 2008 EJ134 2.52 2.68 2.22 3.16 2.63 2.52 2.21 2.62 2.58 2.35 2.68 2.76 2.76 3.13 2.96 5.14 3.18 2.80 2.67 2.59 3.02 3.06 2.32 2.64 3.30 3.20 2.38 2.53 2.37 3.14 2.49 2.40 2.72 2.91 2.79 2.99 2.88 2.97 2.26 2.69 3.17 2.34 2.53 2.47 2.44 2.53 2.56 2.78 2.78 0.16 0.22 0.23 0.02 0.12 0.10 0.05 0.21 0.01 0.13 0.03 0.04 0.20 0.22 0.11 0.06 0.08 0.03 0.21 0.19 0.03 0.04 0.12 0.13 0.10 0.08 0.12 0.03 0.35 0.05 0.04 0.13 0.05 0.14 0.35 0.16 0.01 0.04 0.27 0.25 0.07 0.18 0.11 0.16 0.12 0.08 0.09 0.05 0.07 6.0 7.2 7.4 14.2 4.5 4.6 5.1 7.2 6.0 6.0 2.5 2.6 13.6 17.2 2.3 3.3 2.0 1.8 13.2 13.7 9.9 10.2 2.2 2.4 2.4 2.5 3.0 3.9 6.2 27.1 0.9 0.6 2.4 1.4 8.1 7.3 2.6 1.9 4.7 6.0 11.0 6.0 7.8 6.6 6.2 14.7 15.8 4.0 4.0 1.11 0.75 1.33 1.10 1.55 1.59 1.22 1.63 1.91 1.11 1.91 1.05 1.92 1.11 0.55 1.39 1.60 0.82 1.84 0.66 1.05 0.94 1.06 1.32 1.63 17.3 17.5 16.7 15.3 16.6 17.2 16.8 16.5 16.4 16.9 16.6 17.0 15.9 15.5 15.2 13.3 16.2 17.0 17.2 16.9 15.9 16.0 17.1 17.5 15.3 16.2 18.3 17.6 18.7 15.8 17.1 17.7 17.3 17.8 14.2 15.8 16.7 16.9 20.0 18.9 16.8 18.8 17.8 16.1 16.6 17.6 17.4 16.5 17.0 8 28 8 77 8 46 8 47 7 64 6 46 7 41 7 61 7 52 7 46 8 20 8 29 8 54 8 45 8 45 8 19 8 16 17 24 7 32 8 8 16 8 15 15 126 16 27 8 44 1d 48d 1d 11y 1d 9y 1d 7y 1d 19y 1d 14y 1d 7y 1d 8y 1d 14y 1d 78d 1d 43d 1d 9y 1d 8y 1d 20d 1d 12y 1d 15d 1d 42d 2d 45d 1d 11y 1d 1d 23d 1d 8d 1d 10y 2d 3d 1d 6y 0.27 20.4 0.07 20.0 0.08 20.1 0 0 0 124 0.23 124 0.14 124 0.13 0.07 20.4 -1 124 0.11 0.08 20.5 0.09 20.7 0.09 21.1 0.07 20.3 0.08 20.7 0.06 20.2 0.06 20.5 0.18 21.1 0.11 19.6 0.06 20.5 0.13 19.3 0.11 20.3 0.13 20.6 0 0 -1 -1 -1 -1 -1 -1 -3 -4 -4 0 0 123 0.07 124 0.04 126 0.16 126 0.09 126 0.04 126 0.32 126 0.14 126 0.07 162 0.45 162 0.54 161 0.61 163 0.57 163 0.53 0.36 19.2 11 176 0.53 0.30 20.3 3 173 0.55 0.11 20.4 0.18 20.7 0.08 20.2 0.23 19.8 0.10 20.3 0.11 20.4 11 11 11 -3 -4 -3 184 0.50 184 0.56 184 0.62 161 0.57 162 0.62 162 0.52 4 Table A.2 (continued) – Later identification of unknown asteroids observed at ESO/MPG. Acronym Designation a e i MOID H pos arc rms R β ǫ µ VBTU008 2008 CQ149 VBTU019 2008 ET69 VBTU023 2008 DU59 VBTU025 2008 EW83 VBTU028 2008 EX83 VBTU044 2008 EO84 VBTU047 2008 EP84 VBTU048 2008 EM84 VBTU050 2008 EL84 VBTU055 2005 SD127 VBTU065 2008 EQ84 VBTU098 2008 GS74 VBTU119 2008 EN90 VBTU123 2008 EM154 VBTU125 2008 EO90 VBTU136 2008 DE69 VBTU137 2008 EC155 VBTU145 2008 DM83 VBTU155 2001 SA350 VBTU163 2008 EG145 VBTU174 2008 FQ65 VBTU212 2008 FX41 VBVI001 2005 SO198 VBVI002 2008 EZ127 VBVI004 2008 EU144 VBVI006 2008 EV144 2.36 3.21 2.37 2.41 3.18 2.36 3.13 2.29 2.38 2.74 2.74 3.11 2.39 2.68 3.09 3.09 3.26 3.12 2.37 2.37 2.23 2.98 2.27 2.60 2.78 2.56 2.58 2.28 2.31 1.92 3.05 2.33 2.54 2.48 2.73 3.03 2.26 2.58 2.33 2.66 2.53 2.63 2.33 2.54 2.54 2.55 2.80 0.01 0.21 0.11 0.27 0.08 0.25 0.03 0.08 0.07 0.07 0.08 0.06 0.02 0.31 0.11 0.09 0.01 0.19 0.08 0.08 0.05 0.09 0.31 0.05 0.05 0.02 0.25 0.02 0.17 0.21 0.09 0.18 0.22 0.15 0.03 0.04 0.49 0.14 0.24 0.29 0.03 0.18 0.17 0.13 0.02 0.05 0.22 3.2 9.3 5.8 2.5 4.1 3.7 11.0 7.4 9.2 3.8 3.6 9.6 4.3 7.2 3.9 4.8 5.8 7.6 3.5 3.4 7.9 8.9 6.7 22.3 3.3 2.1 2.8 4.5 3.1 2.0 10.3 4.0 5.5 3.4 2.8 2.6 3.5 13.2 7.0 8.9 1.0 1.5 7.7 11.8 0.8 0.9 2.6 1.34 1.12 0.78 0.78 1.12 1.55 1.90 1.34 1.76 2.21 1.19 1.14 0.59 1.64 1.51 1.23 0.52 0.93 1.12 1.92 0.18 0.78 1.45 0.93 1.48 1.19 18.5 16.4 19.0 19.1 15.7 18.5 15.4 17.6 17.1 17.3 17.6 16.2 18.4 17.5 17.1 16.7 16.5 16.1 18.3 18.6 17.4 19.3 16.8 16.7 17.2 16.9 17.4 18.3 21.8 17.5 18.7 17.8 18.6 16.3 17.0 18.4 17.4 19.7 19.2 17.4 17.2 18.8 17.8 17.2 17.2 18.4 5 5 11 5 4 41 6 62 6 53 18 27 14 18 51 14 43 18 89 12 28 10 13 34 13 12 19 8 16 8 13 8 28 8 84 6 2 25 5 21 14 44 14 28 23 30 15 1d 30d 1d 1d 11y 1d 16y 1d 19y 2d 8d 1d 2d 2y 1d 6y 2d 17y 1d 9d 1d 1d 44d 1d 1d 17d 1d 15d 1d 3d 1d 3y 1d 10y 1d 1d 72d 1d 40d 1d 5y 1d 28d 2d 16d 1d 0.12 21.1 0.13 21.0 0.08 20.0 0.05 19.5 0.09 19.7 0.23 21.0 0.10 20.5 0.17 20.9 0.08 20.6 0.14 20.9 0.10 20.5 -5 -6 -6 -6 -6 -4 -4 -4 -4 -4 -4 192 0.63 192 0.63 192 0.50 193 0.52 192 0.66 189 0.55 188 0.52 189 0.62 189 0.53 189 0.48 188 0.64 0.24 20.8 9 135 0.18 0.09 20.1 0.13 20.6 0.11 20.5 0.07 20.0 0.18 21.5 0.06 20.1 0.06 20.6 0.12 20.9 0.00 18.9 0.14 20.4 0.11 20.3 0.11 20.1 0.18 20.3 0.17 20.5 -1 -1 -1 -4 -4 -4 -4 -3 0 1 0 0 0 0 163 0.69 162 0.51 162 0.54 189 0.66 189 0.55 189 0.61 189 0.60 162 0.47 132 0.28 172 0.65 190 0.60 190 0.67 190 0.58 190 0.54 Table A.2 (continued) – Later identification of unknown asteroids observed at ESO/MPG. Acronym Designation a e i MOID H pos arc rms R β ǫ µ VBVI009 2008 EZ145 VBVI010 2008 EY33 VBVI011 2008 ET144 VBVI016 2002 TT318 VBVI021 2008 EY144 VBVI024 2008 EX144 VBVI026 2008 CR152 VBVI035 2008 EA146 VBVI039 2008 CG89 VBVI041 2008 EB146 VBVI042 2008 EW33 VBTU147 2003 SF319 3.08 3.02 2.66 2.53 2.21 2.27 2.38 2.57 2.38 2.33 3.81 2.69 2.64 2.87 2.27 2.24 2.38 2.41 2.42 2.75 2.18 2.19 0.04 0.23 0.22 0.12 0.11 0.14 0.11 0.24 0.19 0.24 0.30 0.05 0.06 0.15 0.07 0.10 0.02 0.14 0.20 0.04 0.14 0.14 1.9 2.6 3.1 3.2 0.4 0.4 6.8 12.6 2.4 7.0 27.5 24.9 22.1 2.5 6.3 5.7 2.9 2.7 2.6 6.3 2.2 3.5 1.94 1.07 0.98 1.12 0.93 0.78 1.57 1.43 1.12 1.34 0.93 0.89 16.6 16.2 17.2 17.8 18.3 18.8 18.2 17.4 18.7 19.4 16.2 15.8 16.3 16.1 17.1 17.6 17.5 18.2 19.0 17.4 20.1 18.2 22 25 22 56 22 44 8 28 8 8 16 8 52 8 8 41 8 18 8 16 7 52 2d 13d 2d 15y 2d 8y 1d 7y 1d 1d 6d 1d 28d 1d 1d 35d 1d 25d 1d 12d 1d 12y 0.12 20.5 -1 190 0.50 0.21 19.4 0.09 19.9 0.11 20.3 0.08 20.2 0.17 20.4 0.09 19.3 0.07 20.5 0.09 19.1 0.07 20.1 0 0 4 4 4 0 0 -1 -1 191 0.57 190 0.61 193 0.63 193 0.55 192 0.59 190 0.72 191 0.53 191 0.67 191 0.63 0.04 20.4 0 191 0.57 0.22 21.4 -4 189 0.64 6 Table A.3: Outstanding unknown objects observed at ESO/MPG. Acronym a e i MOID H pos arc rms R β ǫ µ VB002 VB003 VB004 VB005 VB006 VB007 VB008 VB010 VB011 VB012 VB013 VB015 VB016 VB017 VB019 VB020 VB021 VB022 VB023 VB025 VB026 VB028 VB030 VB031 VB032 VB033 VB034 VB035 VB038 VB040 VB041 VB042 VB043 VB044 VB046 VB048 VB049 VB050 VB051 VB052 VB053 VB054 VB055 VB056 VB058 VB059 VB060 VB061 VB062 VB063 VB064 3.12 2.38 2.62 3.03 2.22 2.29 2.38 2.85 2.28 2.41 2.31 2.47 2.23 2.41 2.40 2.33 2.59 2.30 2.35 2.56 2.17 2.23 2.51 2.36 2.34 2.67 2.34 2.54 2.16 2.59 2.29 2.45 2.69 2.62 2.21 2.27 3.02 3.12 2.70 2.62 2.27 2.52 2.23 2.40 3.16 2.50 2.49 2.61 2.39 2.27 2.47 0.06 0.11 0.19 0.29 0.05 0.16 0.23 0.23 0.49 0.21 0.23 0.05 0.23 0.21 0.22 0.11 0.00 0.49 0.24 0.09 0.03 0.16 0.07 0.10 0.18 0.09 0.11 0.02 0.02 0.00 0.02 0.05 0.08 0.55 0.13 0.15 0.09 0.06 0.04 0.11 0.14 0.16 0.22 0.03 0.19 0.04 0.04 0.00 0.20 0.07 0.18 8.4 3.0 14.8 2.1 6.5 1.1 2.3 2.3 6.7 4.0 5.9 6.0 5.6 3.4 5.9 5.2 12.2 2.6 4.9 0.9 0.6 6.9 1.5 2.7 2.1 1.6 6.9 9.5 0.7 8.2 8.9 5.5 1.3 2.2 6.7 4.1 14.2 1.6 3.8 3.2 6.6 0.6 5.5 7.3 2.2 4.0 0.9 3.4 3.8 5.3 2.1 8 8 8 7 8 8 7 8 8 7 8 8 8 8 8 8 7 6 8 8 8 7 7 7 7 6 4 6 7 7 7 7 6 7 7 8 8 8 7 8 8 8 8 8 8 8 7 8 8 8 8 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 0.16 0.08 0.16 0.09 0.10 0.14 0.16 0.11 0.13 0.15 0.09 0.09 0.09 0.12 0.18 0.22 0.28 0.09 0.17 0.07 0.08 0.15 0.12 0.25 0.22 0.09 0.07 0.11 0.12 0.14 0.12 0.22 0.10 0.12 0.98 0.11 0.18 0.09 0.21 0.20 0.13 0.17 0.06 0.08 0.19 0.09 0.11 0.09 0.10 0.09 0.13 1.94 1.11 1.11 1.13 1.12 0.92 0.85 1.17 0.19 0.91 0.82 1.35 0.75 0.91 0.89 1.08 1.58 0.18 0.79 1.33 1.10 0.87 1.33 1.13 0.92 1.44 1.08 1.49 1.10 1.59 1.25 1.33 1.47 0.18 0.93 0.92 1.74 1.94 1.59 1.33 0.94 1.11 0.77 1.33 1.56 1.40 1.39 1.59 0.92 1.11 1.03 15.9 17.7 17.6 17.1 17.1 18.2 16.3 17.0 20.5 16.6 17.6 17.0 17.7 16.5 16.7 17.2 17.2 20.6 16.4 17.2 17.9 17.0 17.5 16.7 18.8 16.2 - 17.4 18.1 17.2 17.4 17.8 16.7 20.5 19.3 18.7 16.2 15.9 17.2 16.7 17.1 18.3 17.9 17.7 15.5 17.2 17.0 16.8 18.1 18.0 16.8 7 21.1 20.8 20.6 20.6 20.6 20.8 20.9 21.0 20.8 21.2 20.8 21.0 21.0 21.1 21.2 21.3 21.3 20.8 21.0 20.8 21.0 21.0 21.0 20.7 21.3 20.8 0 0 0 0 0 0 0 0 0 0 0 0 0 0 -1 -1 -1 -1 -1 -1 -1 0 -1 -1 -1 -1 - -1 -1 -1 -1 -1 -1 -1 -1 -1 -1 -1 -1 -1 -1 -1 -1 -1 -1 -1 -1 -1 -2 -1 -1 -1 21.4 21.1 21.2 20.9 21.3 21.3 21.3 20.7 21.2 21.3 21.1 21.2 21.2 21.1 21.3 21.1 21.2 21.2 21.2 21.0 20.8 20.6 21.0 21.3 123 123 124 124 124 124 124 124 124 124 124 124 124 124 124 124 124 124 124 124 124 124 123 125 126 125 125 126 126 126 126 126 126 126 191 126 126 127 127 126 126 126 126 126 126 126 126 126 126 126 126 0.09 0.19 0.38 0.24 0.11 0.26 0.10 0.11 0.55 0.10 0.10 0.09 0.12 0.09 0.12 0.09 0.15 0.28 0.12 0.11 0.13 0.09 0.11 0.07 0.25 0.07 0.10 0.15 0.08 0.14 0.18 0.13 0.08 0.20 0.68 0.22 0.16 0.09 0.07 0.10 0.14 0.14 0.15 0.15 0.13 0.09 0.05 0.08 0.27 0.18 0.11 Table A.3 (continued) – Outstanding unknown objects observed at ESO/MPG. Acronym a e i MOID H pos arc rms R β ǫ µ VB065 VBOP001 VBOP002 VBOP003 VBOP004 VBOP008 VBOP009 VBOP010 VBOP011 VBOP012 VBOP014 VBOP015 VBOP016 VBOP018 VBOP019 VBOP020 VBOP021 VBOP022 VBOP025 VBOP026 VBOP027 VBOP028 VBOP029 VBOP030 VBOP031 VBOP032 VBSO002 VBSO003 VBSO005 VBSO006 VBSO009 VBSO013 VBSO015 VBSO016 VBSO017 VBSO018 VBSO019 VBSO020 VBSO021 VBSO022 VBSO023 VBSO025 VBSO026 VBSO027 VBSO028 VBSO029 VBSO030 VBSO031 VBSO032 VBSO033 VBSO034 2.60 2.43 2.79 2.20 2.04 3.30 2.35 1.85 2.21 2.46 2.34 3.19 2.34 2.33 2.30 2.35 2.73 2.57 2.31 2.61 2.40 2.50 2.34 2.12 2.84 2.69 2.09 2.62 2.01 2.60 2.35 2.13 2.70 2.90 3.14 2.20 1.94 2.24 2.40 2.30 2.79 2.71 3.15 2.41 2.14 2.86 2.54 2.34 2.95 3.11 2.29 0.00 0.08 0.20 0.02 0.28 0.11 0.18 0.25 0.05 0.05 0.00 0.03 0.10 0.00 0.13 0.16 0.05 0.01 0.09 0.23 0.20 0.07 0.22 0.01 0.03 0.25 0.10 0.05 0.20 0.22 0.10 0.12 0.08 0.26 0.05 0.23 0.22 0.27 0.20 0.27 0.09 0.07 0.19 0.20 0.12 0.17 0.19 0.25 0.01 0.06 0.22 3.1 5.6 5.0 4.3 6.3 2.4 4.0 5.0 2.7 2.3 4.2 2.7 4.1 0.6 3.1 2.1 4.5 2.3 4.3 3.2 2.1 0.8 3.3 1.7 3.8 4.2 5.2 6.9 4.2 5.6 7.1 5.5 14.0 5.6 11.5 14.8 3.6 4.7 2.8 2.3 1.8 4.1 2.8 2.5 3.1 2.1 2.0 3.5 2.5 3.1 8.4 6 7 6 8 7 8 8 8 8 8 8 8 8 8 8 8 8 8 6 6 8 8 8 8 8 7 14 11 11 9 8 8 9 8 8 7 8 7 7 6 7 7 7 7 6 7 7 7 6 7 7 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 0.10 0.14 0.11 0.24 0.47 0.12 0.15 0.48 0.18 0.15 0.08 0.18 0.10 0.09 0.09 0.09 0.20 0.10 0.04 0.24 0.06 0.09 0.07 0.11 0.22 0.25 0.21 0.09 0.09 0.08 0.11 0.15 0.06 0.05 0.18 0.36 0.16 0.08 0.10 0.17 0.13 0.12 0.10 0.05 0.24 0.11 0.07 0.09 0.28 0.10 0.49 21.4 20.3 20.2 21.1 21.0 19.6 21.0 20.8 21.0 21.1 20.2 21.0 21.0 20.7 19.9 20.8 20.7 20.8 19.5 20.7 20.6 20.4 20.7 20.5 21.0 20.9 20.7 19.6 20.4 20.5 20.6 20.7 19.8 20.2 20.8 20.7 20.7 20.2 20.7 20.6 20.6 20.7 20.5 20.1 20.6 20.6 20.4 19.9 20.5 20.0 20.3 -1 -3 -3 -3 -3 -3 -4 -4 -4 -4 -4 -4 -4 0 0 0 0 0 0 0 -1 -1 -1 -1 -1 -1 11 11 10 11 11 10 10 10 10 10 10 10 3 3 3 3 3 3 2 2 2 2 2 3 3 126 162 162 162 162 162 162 162 162 162 162 162 162 163 163 163 163 163 163 163 163 163 163 163 163 163 176 176 176 176 176 176 176 176 176 176 176 176 173 173 173 173 173 173 173 173 173 173 173 173 173 0.07 0.56 0.47 0.61 0.76 0.45 0.53 0.74 0.61 0.55 0.58 0.45 0.57 0.60 0.62 0.56 0.54 0.56 0.60 0.50 0.53 0.56 0.52 0.64 0.52 0.48 0.70 0.60 0.61 0.56 0.66 0.69 0.62 0.53 0.52 0.73 0.54 0.53 0.60 0.52 0.57 0.58 0.49 0.59 0.67 0.56 0.59 0.60 0.54 0.52 0.65 1.58 1.24 1.22 1.17 0.51 1.92 0.93 0.42 1.09 1.33 1.32 2.08 1.11 1.32 1.00 0.99 1.60 1.54 1.11 1.02 0.93 1.33 0.82 1.11 1.77 1.03 0.90 1.47 0.63 1.04 1.12 0.90 1.47 1.15 1.98 0.72 0.52 0.65 0.93 0.68 1.54 1.51 1.54 0.93 0.89 1.39 1.07 0.77 1.92 1.92 0.79 17.4 17.3 17.8 18.8 20.5 15.3 19.5 20.8 18.3 18.4 17.5 16.4 18.9 18.0 16.7 19.1 17.5 17.6 17.5 18.9 19.1 17.8 19.6 18.5 17.1 19.0 19.4 16.3 20.2 18.8 18.6 19.5 16.0 18.2 16.4 17.4 21.1 19.9 19.4 20.3 17.5 17.5 15.6 18.9 19.5 16.8 18.7 19.2 16.9 16.4 16.9 8 Table A.3 (continued) – Outstanding unknown objects observed at ESO/MPG. Acronym a e i MOID H pos arc rms R β ǫ µ VBSO035 VBSO037 VBSO038 VBSO039 VBSO040 VBSO041 VBSO042 VBSO043 VBSO044 VBSO045 VBSO046 VBSO047 VBSO048 VBSO049 VBSO050 VBSO051 VBSO053 VBSO054 VBSO055 VBSO056 VBSO057 VBSO058 VBSO060 VBSO061 VBSO062 VBSO064 VBSO065 VBSO067 VBSO069 VBSO070 VBSO071 VBSO072 VBSO073 VBSO074 VBSO075 VBSO076 VBSO077 VBSO078 VBSO079 VBTO003 VBTO004 VBTO007 VBTO009 VBTO014 VBTO016 VBTO017 VBTO018 VBTO020 VBTO026 VBTO030 VBTO031 2.48 2.38 2.72 2.26 2.69 2.25 2.61 2.21 2.76 3.19 2.74 2.47 1.90 2.30 2.10 2.80 2.39 2.27 2.41 2.54 2.87 2.48 2.45 2.50 2.32 2.30 2.98 2.22 2.27 2.35 3.04 1.95 2.21 2.24 2.20 2.31 2.11 2.43 2.64 2.22 2.57 2.33 2.27 2.40 1.40 3.07 2.40 2.32 2.26 2.34 2.47 0.10 0.17 0.17 0.07 0.03 0.06 0.00 0.05 0.19 0.05 0.19 0.06 0.21 0.01 0.11 0.20 0.12 0.22 0.15 0.26 0.02 0.06 0.26 0.04 0.09 0.27 0.24 0.26 0.22 0.01 0.22 0.13 0.01 0.27 0.14 0.17 0.12 0.28 0.10 0.10 0.25 0.35 0.15 0.16 0.15 0.15 0.12 0.09 0.11 0.18 0.22 10.8 5.4 22.1 6.0 7.7 2.0 6.8 2.6 4.4 1.7 4.3 11.5 1.5 6.0 1.0 1.7 5.3 4.6 7.1 2.1 2.1 2.9 3.7 3.2 3.7 4.5 8.3 4.5 4.8 6.2 9.9 17.5 6.9 5.2 5.4 5.5 7.0 5.0 7.2 1.8 2.8 9.1 7.0 8.1 1.7 3.4 3.1 3.3 1.8 4.0 3.5 4 7 7 8 8 6 7 7 8 6 8 8 8 8 8 8 7 8 6 8 8 8 6 6 6 8 8 8 7 5 7 8 8 8 8 5 4 5 5 14 14 11 12 13 14 14 14 8 8 8 8 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 0.39 0.09 0.11 0.14 0.09 0.05 0.23 0.13 0.08 0.22 0.08 0.17 0.14 0.07 0.12 0.16 0.04 0.07 0.36 0.08 0.11 0.07 0.05 0.07 0.13 0.12 0.18 0.14 0.30 0.09 0.25 0.12 0.26 0.27 0.19 0.38 0.10 0.31 0.35 0.18 0.13 0.65 0.44 0.26 0.87 0.17 0.27 0.14 0.11 0.12 0.11 20.6 20.0 20.4 20.3 20.2 19.7 20.3 20.1 20.4 20.3 20.5 18.7 20.6 20.2 20.1 20.6 19.7 20.4 20.0 20.5 20.3 20.4 19.7 20.4 20.2 20.7 20.5 20.6 20.6 19.6 20.6 20.0 20.5 20.9 20.7 20.7 20.6 20.6 21.1 20.6 20.6 20.8 20.8 20.9 20.6 20.5 20.0 19.6 20.9 21.0 20.9 3 3 3 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 11 11 11 11 11 11 11 11 11 11 11 11 11 11 -3 -3 -3 -4 -4 -4 -3 -3 -3 -3 -4 -4 173 173 173 117 117 117 117 117 163 163 163 163 163 163 164 163 164 164 164 163 163 163 163 163 163 184 184 184 184 184 184 184 184 184 184 184 184 184 184 161 161 162 162 162 161 161 161 162 162 162 162 0.69 0.64 0.70 0.30 0.19 0.28 0.14 0.16 0.51 0.49 0.51 0.63 0.49 0.63 0.62 0.51 0.59 0.55 0.61 0.49 0.51 0.57 0.53 0.57 0.59 0.51 0.55 0.53 0.57 0.65 0.55 1.17 0.69 0.53 0.64 0.62 0.73 0.53 0.56 0.58 0.52 0.76 0.59 0.57 0.68 0.46 0.53 0.56 0.58 0.53 0.47 1.23 0.96 1.38 1.10 1.59 1.10 1.59 1.10 1.23 2.03 1.22 1.34 0.51 1.27 0.88 1.26 1.11 0.77 1.05 0.88 1.81 1.33 0.83 1.39 1.11 0.68 1.29 0.66 0.78 1.33 1.39 0.70 1.18 0.66 0.90 0.94 0.85 0.78 1.38 1.01 0.91 0.56 0.93 1.04 0.20 1.60 1.11 1.11 1.02 0.93 0.93 18.4 16.6 16.9 17.1 16.0 16.5 16.1 16.4 18.0 16.3 18.1 15.7 20.9 17.5 18.7 18.1 17.7 19.5 18.1 19.2 16.5 17.8 18.5 17.2 18.1 20.2 18.1 20.3 19.7 17.0 17.9 19.4 18.3 20.5 19.4 19.2 18.3 19.8 17.5 18.6 16.3 18.5 19.2 18.9 22.7 17.1 17.8 17.5 18.9 19.4 19.3 9 Table A.3 (continued) – Outstanding unknown objects observed at ESO/MPG. Acronym a e i MOID H pos arc rms R β ǫ µ VBTO032 VBTO033 VBTO034 VBTO037 VBTO041 VBTO042 VBTO048 VBTO050 VBTO053 VBTOb10 VBTOb13 VBTU003 VBTU004 VBTU007 VBTU009 VBTU010 VBTU021 VBTU022 VBTU026 VBTU031 VBTU032 VBTU036 VBTU040 VBTU042 VBTU043 VBTU046 VBTU051 VBTU053 VBTU067 VBTU081 VBTU082 VBTU084 VBTU086 VBTU087 VBTU088 VBTU091 VBTU093 VBTU094 VBTU095 VBTU096 VBTU097 VBTU099 VBTU100 VBTU102 VBTU105 VBTU106 VBTU109 VBTU111 VBTU112 VBTU114 VBTU115 2.23 2.50 1.93 2.30 2.33 2.34 2.22 3.08 2.31 2.22 2.49 2.40 2.30 2.39 2.88 2.29 2.60 2.22 1.97 1.88 1.99 2.33 2.47 2.39 1.92 2.72 2.38 2.77 2.77 2.31 2.74 1.98 2.77 2.22 2.45 2.12 2.36 2.06 3.19 2.42 2.48 1.95 2.28 2.27 2.78 2.36 2.35 2.33 2.39 2.44 2.45 0.16 0.19 0.17 0.19 0.22 0.10 0.18 0.25 0.21 0.05 0.06 0.13 0.16 0.12 0.25 0.19 0.35 0.32 0.36 0.38 0.30 0.00 0.16 0.19 0.25 0.05 0.11 0.05 0.27 0.01 0.05 0.06 0.06 0.13 0.22 0.10 0.01 0.16 0.08 0.50 0.05 0.13 0.02 0.03 0.17 0.19 0.14 0.31 0.12 0.21 0.28 1.9 6.1 3.2 7.0 3.2 5.6 6.6 3.2 2.3 2.5 3.6 4.2 5.5 5.4 15.8 4.6 6.7 5.2 18.7 19.1 16.8 3.8 6.6 4.5 1.4 3.3 5.4 2.8 3.3 2.9 9.7 9.2 4.0 9.0 2.2 2.5 6.4 5.5 17.1 7.7 7.1 4.8 6.1 6.6 19.6 5.8 1.7 3.2 0.5 2.9 1.3 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 2d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 0.21 0.16 0.49 0.14 0.13 0.17 0.47 0.16 0.12 0.15 0.11 0.32 0.16 0.13 0.15 0.09 0.14 0.20 0.57 0.53 0.67 0.24 0.23 0.33 0.08 0.12 0.18 0.39 0.24 0.10 0.04 0.24 0.10 0.18 0.16 0.07 0.05 0.20 0.21 0.20 0.21 0.03 0.09 0.14 0.24 0.19 0.07 0.20 0.10 0.07 0.08 20.9 21.0 20.7 21.1 21.1 21.1 21.0 20.9 20.9 20.9 20.5 21.1 21.3 20.7 20.9 20.8 20.5 18.9 20.8 21.1 21.1 21.4 20.7 20.7 20.6 20.9 21.2 21.2 21.2 20.8 21.4 21.3 21.5 21.4 21.4 20.9 20.8 21.1 20.7 21.1 18.6 19.6 20.6 20.8 20.6 20.7 19.6 20.4 19.7 19.7 20.0 -4 -4 -4 -4 -4 -4 -3 -3 -3 -4 -4 -5 -5 -6 -5 -6 -3 -3 -6 -6 -6 -6 -3 -3 -4 -4 -4 -4 -4 -5 -4 -5 -4 -4 -4 -4 -4 9 9 9 9 9 9 9 -3 -3 0 0 0 -1 -1 162 162 162 162 162 162 162 162 162 162 162 192 192 192 192 192 161 161 193 192 192 192 161 161 188 188 188 189 188 189 189 189 189 189 189 189 189 135 135 135 135 135 135 135 161 161 162 162 163 163 163 0.55 0.53 0.54 0.61 0.48 0.56 0.61 0.45 0.51 0.60 0.52 0.59 0.62 0.62 0.53 0.69 0.66 0.64 0.78 0.84 0.87 0.62 0.55 0.51 0.49 0.57 0.65 0.54 0.55 0.65 0.60 0.79 0.56 0.74 0.56 0.71 0.66 0.22 0.19 0.51 0.24 0.23 0.19 0.17 0.57 0.51 0.55 0.64 0.53 0.50 0.41 8 8 8 8 8 8 7 8 8 8 16 14 3 4 5 4 15 14 6 4 5 5 14 8 14 14 10 10 13 3 3 3 3 3 3 4 4 9 13 12 9 14 14 14 13 10 13 13 13 11 11 0.87 1.02 0.61 0.87 0.82 1.11 0.82 1.30 0.82 1.10 1.33 1.11 0.93 1.12 1.17 0.84 0.72 0.54 0.24 0.17 0.39 1.31 1.06 0.93 0.44 1.59 1.12 1.62 1.04 1.27 1.61 0.85 1.61 0.93 0.93 0.89 1.33 0.75 1.92 0.26 1.37 0.70 1.23 1.21 1.31 0.93 1.04 0.63 1.11 0.93 0.77 19.5 19.1 20.3 19.6 19.8 19.0 19.8 18.1 19.7 18.2 17.9 19.1 19.8 18.7 16.1 18.6 18.4 18.3 17.5 18.0 17.9 18.8 17.0 19.1 21.4 17.7 19.3 17.5 19.6 18.2 18.2 19.4 18.3 19.5 20.1 18.4 18.3 18.5 16.2 20.4 14.7 17.5 17.3 17.5 16.3 19.1 17.6 19.0 17.6 18.2 18.9 10 Table A.3 (continued) – Outstanding unknown objects observed at ESO/MPG. Acronym a e i MOID H pos arc rms R β ǫ µ VBTU116 VBTU117 VBTU118 VBTU120 VBTU121 VBTU122 VBTU126 VBTU127 VBTU129 VBTU130 VBTU131 VBTU132 VBTU133 VBTU140 VBTU142 VBTU143 VBTU144 VBTU146 VBTU148 VBTU151 VBTU152 VBTU156 VBTU162 VBTU164 VBTU166 VBTU167 VBTU168 VBTU169 VBTU170 VBTU171 VBTU172 VBTU173 VBTU175 VBTU176 VBTU177 VBTU178 VBTU179 VBTU180 VBTU181 VBTU183 VBTU184 VBTU188 VBTU189 VBTU190 VBTU191 VBTU192 VBTU193 VBTU194 VBTU195 VBTU196 VBTU197 2.43 2.22 2.21 2.54 2.39 2.54 2.32 3.12 2.32 2.36 3.07 2.31 2.32 2.22 2.73 2.10 2.37 2.50 2.46 2.35 2.61 2.36 2.36 2.42 2.44 2.26 2.24 2.74 2.78 2.42 2.29 2.38 2.46 2.61 2.27 3.11 2.43 2.37 3.04 2.37 2.20 2.68 2.93 2.42 3.15 2.01 3.40 2.21 2.30 2.20 3.02 0.13 0.13 0.17 0.02 0.08 0.07 0.01 0.07 0.17 0.11 0.05 0.12 0.12 0.17 0.05 0.30 0.25 0.07 0.28 0.25 0.22 0.25 0.08 0.04 0.07 0.15 0.16 0.15 0.24 0.07 0.08 0.02 0.18 0.12 0.03 0.06 0.07 0.11 0.09 0.10 0.18 0.22 0.26 0.04 0.24 0.14 0.24 0.18 0.46 0.06 0.68 3.6 7.4 4.2 2.5 11.6 3.9 2.3 2.4 5.1 3.2 3.3 22.7 22.8 4.7 2.5 3.6 4.5 3.6 2.7 2.8 9.3 2.1 3.9 3.1 6.5 6.8 7.0 13.8 16.0 1.4 3.6 1.2 7.5 10.6 7.0 2.0 0.6 7.1 2.2 5.2 7.3 9.2 11.3 7.6 11.8 5.7 14.2 7.1 5.3 8.5 8.2 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 0.15 0.19 0.19 0.07 0.72 0.22 0.14 0.18 0.20 0.25 0.42 0.95 0.49 0.05 0.24 0.75 0.20 0.10 0.21 0.13 0.39 0.20 0.08 0.22 0.24 0.28 0.18 0.13 0.21 0.20 0.13 0.14 0.23 0.17 0.19 0.16 0.20 0.09 0.09 0.11 0.21 0.09 0.08 0.09 0.08 0.09 0.12 0.18 0.17 0.13 0.38 20.8 20.7 20.7 19.6 20.9 20.8 20.8 20.8 20.6 20.6 21.0 21.1 21.0 20.4 21.5 21.7 21.4 20.9 21.4 21.3 21.3 21.5 20.9 20.9 21.2 21.2 21.3 21.5 21.0 21.3 20.7 20.8 21.1 20.9 21.0 21.1 21.0 19.4 20.9 21.3 21.3 20.9 20.7 19.7 21.1 21.0 20.9 21.5 21.3 21.2 21.4 -1 -1 -1 -1 -1 -1 -1 -1 0 0 0 -28 -28 -4 -4 -4 -4 -4 -4 -4 -4 -4 -3 -3 1 1 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 11 10 10 10 10 11 10 10 10 10 163 163 163 162 162 162 162 162 162 162 162 183 183 189 189 189 189 189 189 189 189 189 162 162 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 135 136 136 136 136 136 136 135 135 135 0.53 0.62 0.62 0.54 0.59 0.53 0.58 0.46 0.55 0.55 0.47 0.70 0.70 0.67 0.56 0.75 0.56 0.60 0.50 0.67 0.66 0.53 0.57 0.55 0.17 0.21 0.21 0.27 0.40 0.15 0.10 0.09 0.22 0.21 0.17 0.16 0.15 0.16 0.17 0.18 0.22 0.27 0.22 0.19 0.20 0.22 0.19 0.30 0.47 0.32 0.68 6 11 13 10 10 12 12 12 12 9 8 5 8 8 6 3 6 8 8 8 8 7 8 5 7 7 8 6 8 8 8 8 6 8 6 6 5 8 3 4 7 8 7 7 7 7 7 8 7 8 7 1.11 0.93 0.83 1.48 1.21 1.35 1.30 1.92 0.93 1.11 1.92 1.09 1.10 0.84 1.61 0.47 0.78 1.34 0.78 0.74 1.04 0.78 1.18 1.33 1.28 0.93 0.89 1.33 1.11 1.25 1.11 1.33 1.01 1.31 1.21 1.93 1.26 1.11 1.78 1.14 0.81 1.09 1.18 1.34 1.40 0.74 1.59 0.84 0.26 1.08 0.09 18.7 19.1 17.4 16.3 17.9 17.7 18.1 16.9 19.0 18.5 17.1 17.6 17.5 17.2 18.3 18.8 20.6 18.3 20.6 18.0 19.6 20.7 18.5 18.2 17.3 17.3 17.4 18.1 18.2 17.4 17.8 17.5 16.6 16.4 17.6 16.0 17.1 16.5 15.8 17.4 17.4 16.7 15.9 16.4 16.3 18.9 16.1 17.7 21.5 17.9 18.9 11 Table A.3 (continued) – Outstanding unknown objects observed at ESO/MPG. Acronym a e i MOID H pos arc rms R β ǫ µ VBTU198 VBTU200 VBTU201 VBTU205 VBTU206 VBTU209 VBTU210 VBTU211 VBTU214 VBTU215 VBTU216 VBTU217 VBTU218 VBTU219 VBTU220 VBTU221 VBTU222 VBTU223 VBTU225 VBTU226 VBTU227 VBTU228 VBTU229 VBTU230 VBTU231 VBTU232 VBTU233 VBTU234 VBTU235 VBTU236 VBTU237 VBTU238 VBTU239 VBTU240 VBTU241 VBTU242 VBTU243 VBTU244 VBTU245 VBTU246 VBTU247 VBTU248 VBTUb61 VBVA002 VBVA003 VBVI005 VBVI007 VBVI008 VBVI012 VBVI013 VBVI014 2.42 2.42 2.66 2.25 2.38 2.39 2.89 2.44 2.22 2.64 3.16 2.75 3.05 2.61 2.38 2.54 1.01 2.23 2.23 1.94 2.47 2.28 2.60 2.54 2.80 1.94 2.34 2.56 2.70 2.54 2.52 2.36 2.11 2.06 2.25 2.97 2.68 2.26 2.53 2.98 2.61 2.41 2.29 2.50 2.84 2.92 2.64 2.70 2.38 2.35 2.20 0.04 0.07 0.10 0.06 0.51 0.12 0.01 0.04 0.17 0.02 0.04 0.06 0.04 0.01 0.08 0.02 0.08 0.26 0.27 0.20 0.10 0.16 0.07 0.03 0.14 0.20 0.11 0.02 0.04 0.02 0.16 0.10 0.23 0.18 0.15 0.13 0.09 0.03 0.08 0.11 0.11 0.12 0.22 0.39 0.25 0.25 0.20 0.25 0.18 0.25 0.06 6.9 7.5 10.2 5.0 1.3 4.6 2.1 3.7 2.0 3.3 11.3 5.2 8.8 2.6 2.1 2.7 2.2 5.3 4.8 25.7 6.1 5.1 11.2 14.8 13.1 0.8 5.9 3.5 4.1 8.6 3.8 5.9 0.4 4.7 6.4 11.5 2.4 6.0 13.1 0.6 3.2 3.0 12.0 13.8 1.4 3.0 14.6 5.7 11.5 6.1 8.0 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 0.17 0.10 0.09 0.41 0.87 0.26 0.10 0.14 0.07 0.12 0.06 0.09 0.03 0.15 0.22 0.09 0.16 0.36 0.25 0.52 0.19 0.07 0.17 0.46 0.16 0.08 0.20 0.09 0.15 0.17 0.35 0.06 0.09 0.16 0.18 0.14 0.09 0.10 0.25 0.07 0.10 0.10 0.14 0.08 0.12 0.12 0.17 0.50 0.09 0.26 0.28 21.5 21.1 20.6 21.3 21.3 21.2 20.6 20.9 20.7 20.5 20.7 20.6 20.9 20.6 21.1 21.1 20.4 20.9 21.1 21.0 20.9 20.7 20.2 20.6 21.4 21.0 21.2 21.1 21.5 20.7 21.3 20.7 21.2 21.2 21.1 21.3 20.7 20.8 21.2 20.8 20.7 20.9 21.2 19.8 21.1 20.4 20.2 19.9 20.4 18.8 20.4 10 0 1 0 0 1 1 1 1 1 1 1 1 1 1 1 10 10 10 10 10 10 -17 -17 -1 -1 0 0 0 0 -1 0 -1 -1 -1 -1 -1 -1 -1 -1 -1 -1 -4 24 -1 0 0 0 4 4 4 135 132 132 132 131 171 172 172 172 171 171 171 171 171 171 171 162 162 162 162 162 162 188 188 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 189 154 126 190 190 190 192 192 192 0.20 0.18 0.20 0.11 0.28 0.63 0.54 0.63 0.68 0.59 0.52 0.57 0.54 0.59 0.65 0.60 2.12 0.63 0.63 0.89 0.54 0.53 0.58 0.58 0.18 0.07 0.09 0.04 0.07 0.12 0.11 0.11 0.11 0.09 0.11 0.19 0.08 0.09 0.22 0.10 0.08 0.16 0.69 0.60 0.17 0.52 0.73 0.56 0.64 0.67 0.69 6 8 14 7 4 4 5 5 5 5 5 4 4 4 5 4 5 5 4 5 5 5 7 6 5 8 8 8 5 3 8 8 8 6 8 8 8 8 8 8 8 8 4 3 8 14 11 8 7 7 7 1.34 1.25 1.41 1.11 0.19 1.11 1.85 1.34 0.84 1.60 2.02 1.58 1.92 1.59 1.17 1.47 0.03 0.66 0.65 0.56 1.22 0.94 1.43 1.52 1.42 0.58 1.08 1.52 1.59 1.49 1.13 1.12 0.63 0.71 0.92 1.59 1.45 1.19 1.33 1.66 1.33 1.11 0.79 0.57 1.12 1.19 1.12 1.04 0.96 0.76 1.08 18.2 17.2 16.1 18.4 20.6 19.3 17.0 18.5 17.7 17.5 16.4 16.9 17.2 17.6 18.1 18.1 24.7 19.6 19.8 18.2 18.4 19.1 17.2 17.3 17.4 18.3 17.2 17.1 17.4 16.7 16.8 16.7 17.9 18.4 17.1 17.2 16.1 17.3 17.7 15.6 17.2 17.9 17.7 17.8 17.7 18.2 18.3 18.2 16.8 17.4 18.2 12 Table A.3 (continued) – Outstanding unknown objects observed at ESO/MPG. Acronym a e i MOID H pos arc rms R β ǫ µ VBVI015 VBVI017 VBVI018 VBVI019 VBVI020 VBVI022 VBVI023 VBVI027 VBVI029 VBVI030 VBVI031 VBVI032 VBVI033 VBVI034 VBVI036 VBVI038 VBVI040 VBVI044 VBVI213 2.39 2.41 2.40 2.64 2.40 2.26 2.55 2.24 2.24 2.27 2.45 2.50 2.00 2.28 2.86 2.01 2.41 2.37 2.45 0.20 0.18 0.20 0.01 0.20 0.27 0.17 0.07 0.08 0.20 0.19 0.12 0.21 0.16 0.29 0.21 0.23 0.21 0.57 7.1 3.4 2.9 12.9 2.1 7.9 14.5 8.0 8.0 0.2 2.3 5.5 4.2 7.5 3.4 4.2 2.2 7.3 2.0 0.93 0.99 0.93 1.60 0.93 0.65 1.12 1.09 1.07 0.82 0.97 1.20 0.59 0.93 1.04 0.59 0.84 0.89 0.07 19.0 19.1 18.8 16.1 19.3 19.1 16.8 18.6 18.6 19.4 16.8 16.7 20.4 19.1 18.6 20.2 17.8 18.9 25.3 7 8 8 8 7 8 8 8 8 7 8 8 8 7 8 8 8 8 5 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 0.12 0.14 0.12 0.07 0.12 0.28 0.07 0.57 0.58 0.14 0.08 0.06 0.60 0.08 0.10 0.56 0.10 0.55 0.31 20.5 20.8 20.3 19.3 20.8 19.9 18.9 20.5 20.4 20.4 20.6 20.2 20.7 20.5 20.3 20.5 20.8 20.5 20.0 4 4 4 3 4 4 4 0 0 0 0 0 0 0 0 -1 -1 0 -1 192 193 193 193 193 192 192 190 191 191 191 191 191 191 191 191 191 190 191 0.62 0.58 0.54 0.63 0.54 0.74 0.71 0.69 0.69 0.59 0.61 0.61 0.69 0.67 0.50 0.69 0.67 0.69 7.17 13 Table A.4: Later identification of unknown asteroids observed with Swope telescope. First line represent our data and second line reffers to updated MPC data. Acronym Designation a e i MOID H pos arc rms R β ǫ µ PBS007 2008 UC335 PBS009 2008 UE35 PBS015 2008 SM192 PBT010 2008 TZ101 PBT012 2008 UZ25 PBT021 2008 UL55 PBT023 2008 TW81 PBT025 1998 UB46 2.44 2.14 2.27 2.21 2.38 2.8 2.48 2.42 2.33 3.1 2.13 2.37 2.44 2.99 2.65 2.9 0.27 0.25 0.09 0.1 0.23 0.14 0.17 0.19 0.24 0.19 0.13 0.12 0.19 0.11 0.24 0.17 3.2 1.9 5.2 3.1 3.7 5.4 1.7 1.5 4.6 12.2 0.7 1 0.4 1.6 3.7 5.4 0.78 1.07 0.83 1.06 0.77 0.86 0.99 1.03 19.3 20.2 17.1 18.3 18.6 16.9 18.0 18.5 19.0 17 18.5 17.2 18.4 17.2 18.1 17.4 5 11 8 30 10 20 7 38 3 18 6 79 5 24 10 31 1d 18d 1d 36d 1d 30d 1d 48d 1d 35d 1d 12y 1d 35d 1d 11y 0.27 19.9 0.22 19.8 1 0 182 0.53 182 0.67 0.22 19.5 -7 175 0.59 0.12 19.8 0.19 19.8 0.33 20.0 0.20 20.1 0 0 0 0 188 0.59 188 0.57 197 0.56 197 0.49 0.13 19.8 -3 187 0.55 14 Table A.5: Outstanding unknown objects observed with Swope telescope. Acronym a e i MOID H pos arc rms R β ǫ µ PBS001 PBS002 PBS003 PBS004 PBS005 PBS006 PBS008 PBS010 PBS011 PBS014 PBS016 PBS017 PBT001 PBT002 PBT006 PBT007 PBT008 PBT009 PBT011 PBT013 PBT014 PBT015 PBT016 PBT017 PBT018 PBT022 PBT024 PBT026 PBT027 PBT028 PBV002 PBV003 PBV004 2.20 2.44 2.40 2.50 2.20 2.49 2.72 2.24 2.25 2.97 2.37 2.77 2.72 2.91 2.20 2.30 2.33 2.72 2.06 2.82 2.21 2.74 2.14 2.32 3.08 2.28 2.40 2.49 2.67 2.40 2.15 2.35 2.30 0.21 0.24 0.13 0.16 0.21 0.03 0.18 0.19 0.22 0.16 0.07 0.04 0.03 0.18 0.05 0.01 0.22 0.32 0.16 0.27 0.15 0.30 0.12 0.21 0.02 0.24 0.20 0.23 0.09 0.19 0.15 0.22 0.40 4.1 3.6 4.2 9.4 4.3 5.4 6.0 6.1 7.7 13.5 10.4 4.8 24.0 10.1 4.9 5.3 3.1 0.5 4.3 2.8 3.2 3.8 2.3 7.9 2.1 4.3 0.7 2.6 6.2 3.7 2.3 3.2 2.4 0.74 0.85 1.10 1.15 0.74 1.42 1.23 0.83 0.78 1.49 1.21 1.66 1.65 1.40 1.09 1.27 0.82 0.85 0.74 1.06 0.89 0.91 0.89 0.85 2.04 0.72 0.93 0.93 1.42 0.95 0.84 0.83 0.37 19.5 19.3 18.3 18.0 18.8 17.2 17.2 18.8 17.8 14.9 16.8 16.5 15.6 16.5 16.4 17.0 19.1 15.3 18.8 18.4 18.9 15.1 18.9 19.0 16.0 16.4 18.7 18.7 16.7 16.4 18.4 15.8 19.7 6 6 6 6 6 8 8 8 6 10 8 10 12 4 5 4 3 7 8 3 6 6 6 4 6 6 3 9 9 6 4 4 4 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 0.23 0.14 0.18 0.28 0.07 0.25 0.15 0.17 0.11 0.19 0.32 0.48 0.16 0.40 0.33 0.27 0.43 0.36 0.16 0.31 0.28 0.14 0.34 0.21 0.17 0.45 0.42 0.22 0.30 0.46 0.22 0.34 0.40 20.5 20.6 20.4 20.4 19.8 19.9 19.3 20.3 19.1 19.5 19.7 19.6 19.1 19.2 19.9 20.0 20.1 20.1 19.5 20.2 20.0 19.8 20.1 20.2 19.9 20.1 20.3 20.0 19.9 20.0 19.2 19.3 19.5 -7 -7 -7 -8 -8 1 0 -12 -17 -7 -7 -8 -12 -12 4 -7 0 0 0 0 -3 -3 -3 -3 -3 0 0 -3 -3 -1 -5 -5 -5 194 194 194 194 194 182 182 153 197 175 175 175 174 174 -108 -149 188 188 188 188 184 184 184 184 184 197 197 187 187 196 176 176 176 0.52 0.46 0.53 0.59 0.52 0.62 0.60 0.46 0.49 0.54 0.68 0.59 0.71 0.57 0.57 0.37 0.57 0.54 0.66 0.52 0.65 0.55 0.66 0.72 0.51 0.58 0.46 0.55 0.59 0.56 0.68 0.64 0.60 15 Table A.6: Later identification of unknown asteroids observed at the INT. First line represent our data and second line reffers to updated MPC data. Acronym Designation a e i MOID H pos arc rms R β ǫ µ VDT070 247902 VDT072 2005 YW147 VDT073 2008 SZ203 VIT003 2005 TO93 VIT005 2000 SN27 VTD001 2010 CD146 VTD003 2010 CJ33 VTD004 2001 TY142 VTD012 2010 CU38 VTD025 2004 RJ308 VTD027 241531 VTD067 2003 UJ281 VTD073 2007 UN95 VTU006 2009 BC142 VTU008 2009 BF55 VTU010 2009 BR86 VTU013 2010 JR112 VTU017 2009 BS86 VTU018 2009 BY86 VTU030 2009 FS67 VTU033 2009 FK20 VTU035 2002 GU116 VTU038 215040 2.30 2.86 2.38 2.42 2.43 2.52 2.77 2.60 1.58 2.66 2.40 2.76 1.40 2.43 2.34 3.20 2.40 2.29 2.15 2.37 2.44 2.68 2.35 2.85 2.14 2.78 2.50 2.71 2.33 3.16 2.41 2.81 2.57 2.55 2.49 2.58 2.11 2.32 2.39 2.76 2.36 2.65 2.24 2.33 2.45 2.55 0.08 0.20 0.11 0.15 0.13 0.06 0.06 0.10 0.13 0.20 0.20 0.09 0.03 0.13 0.30 0.13 0.20 0.14 0.08 0.18 0.02 0.04 0.01 0.20 0.09 0.08 0.19 0.09 0.22 0.14 0.20 0.14 0.03 0.28 0.15 0.05 0.24 0.15 0.12 0.09 0.29 0.28 0.04 0.17 0.05 0.13 6.4 9.0 6.3 5.8 2.7 3.4 6.5 8.3 6.0 14.0 3.9 4.0 2.1 3.3 5.8 10.3 6.5 7.4 9.0 9.6 0.9 7.5 7.4 9.0 1.5 2.6 2.2 4.5 5.4 16.2 3.5 5.7 4.2 4.9 3.6 4.5 2.0 1.0 5.1 7.7 3.8 4.9 7.1 7.3 6.9 7.2 1.12 1.12 1.12 1.60 0.38 0.93 0.35 0.67 0.93 0.98 1.38 1.33 0.95 1.05 0.83 0.94 1.50 1.12 0.60 1.11 0.67 1.14 1.35 17.8 16.3 17.3 17.6 17.8 17.2 16.2 16.9 19.9 16.2 18.2 16.7 20.0 17.9 18.5 14.9 18.1 18.3 17.5 17.0 16.4 16.2 17.8 16.3 16.8 16.1 18.6 16.8 18.9 16.8 18.1 16.7 16.8 15.7 18.0 17.4 17.7 18.8 18.0 16.5 18.8 17.2 17.1 17.4 15.6 16.0 4 51 5 33 2 30 4 53 4 78 4 11 4 11 4 50 5 18 5 29 5 63 4 51 3 33 6 25 4 19 5 31 8 59 5 17 4 16 5 26 5 22 5 49 5 52 1d 10y 1d 6y 1d 3y 1d 14y 1d 10y 1d 11d 1d 11d 1d 10y 1d 11d 1d 18y 1d 11y 1d 10y 1d 4y 1d 10y 1d 40d 1d 4y 1d 6y 1d 38d 1d 18d 1d 10y 1d 35d 1d 9y 1d 7y 0.19 20.3 0.41 19.7 0.00 19.7 0.42 20.4 10 10 5 4 151 0.37 151 0.39 173 0.61 240 0.17 0.38 19.0 -9 167 0.95 0.16 19.6 0.56 19.3 6 6 193 0.57 193 1.15 0.19 18.9 14 183 0.49 0.28 19.2 0.41 19.6 0.38 19.1 0 0 0 178 0.67 178 0.78 178 0.63 0.16 20.8 10 151 0.43 0.89 20.8 -2 106 0.39 0.48 20.1 0.19 19.9 0.19 19.5 0.18 19.6 0.27 20.0 0.33 20.3 0.23 20.2 0.41 20.2 0.19 19.6 4 1 1 7 1 1 3 8 8 183 0.58 189 0.65 189 0.59 182 0.62 189 0.60 189 0.75 202 0.52 208 0.43 208 0.50 0.20 19.1 10 232 0.16 16 Table A.7: Outstanding unknown objects observed at the INT. Acronym a e i MOID H pos arc rms R β ǫ µ VBA001 VBA002 VBA003 VBA004 VBA005 VBA006 VDT055 VDT056 VDT057 VDT058 VDT071 VIBC01 VIBC02 VIBC03 VIBC04 VIT001 VIT002 VIT004 VIT006 VITB01 VTD005 VTD006 VTD007 VTD008 VTD009 VTD010 VTD011 VTD013 VTD014 VTD015 VTD016 VTD017 VTD018 VTD019 VTD021 VTD022 VTD026 VTD028 VTD029 VTD032 VTD033 VTD034 VTD035 VTD036 VTD037 VTD038 VTD039 VTD043 VTD044 VTD045 VTD046 2.26 2.72 1.80 2.66 2.65 2.27 2.61 2.38 2.24 2.44 2.28 2.32 2.54 2.41 2.48 2.52 1.97 2.33 2.24 3.28 2.20 2.25 2.99 2.34 2.22 2.63 3.08 2.37 2.46 2.35 2.21 2.37 2.30 2.06 2.01 2.28 2.28 2.14 2.40 2.34 2.78 3.06 2.38 2.29 2.37 2.24 2.21 2.56 3.17 2.37 2.42 0.13 0.23 0.02 0.30 0.31 0.22 0.00 0.11 0.24 0.21 0.40 0.07 0.50 0.47 0.06 0.24 0.02 0.32 0.19 0.01 0.30 0.32 0.02 0.10 0.31 0.26 0.05 0.08 0.22 0.30 0.10 0.25 0.02 0.34 0.02 0.31 0.08 0.12 0.20 0.17 0.15 0.17 0.22 0.06 0.06 0.14 0.38 0.02 0.18 0.02 0.27 5.6 11.0 23.5 11.7 4.7 4.6 2.3 6.0 5.4 0.7 5.0 5.1 2.5 4.9 3.7 8.3 23.9 1.0 5.9 21.1 5.3 5.4 10.7 5.9 1.0 21.3 2.9 0.5 3.1 1.2 4.5 2.4 6.2 4.7 5.6 3.6 6.9 2.7 3.3 3.3 9.8 2.7 7.7 6.3 6.4 5.5 4.7 19.9 17.4 3.4 3.5 5 5 5 3 5 5 5 3 4 4 4 5 5 5 5 4 4 5 4 5 3 3 3 4 5 5 5 5 5 5 5 5 5 5 5 4 5 5 5 5 4 4 5 5 5 4 4 4 4 5 4 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 0.23 0.22 0.08 0.53 0.85 0.26 0.24 0.17 0.44 0.31 0.43 0.08 0.07 0.09 0.13 0.28 0.54 0.13 0.25 0.26 0.44 0.16 0.43 0.41 0.27 0.30 0.24 0.19 0.26 0.22 0.16 0.80 0.43 0.35 0.64 0.81 0.41 0.55 0.45 0.39 0.26 0.60 0.29 0.22 0.17 0.25 0.29 0.65 0.78 0.35 0.31 20.2 20.8 21.1 20.2 20.2 19.7 20.6 20.8 20.6 20.6 19.8 20.1 19.9 20.0 19.8 20.2 20.2 20.4 20.8 20.1 19.4 19.6 19.4 20.8 20.1 20.2 20.4 20.0 20.1 20.1 19.4 20.1 20.2 20.0 19.9 19.6 19.4 20.3 20.3 20.4 20.0 20.5 20.5 20.1 20.4 20.8 20.2 20.2 20.0 20.5 20.5 8 8 43 -9 -8 -8 -1 -1 -1 -1 10 -1 -1 -1 -1 4 4 -1 -9 -9 14 14 15 11 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 10 10 10 11 11 17 17 -1 -1 208 208 219 165 165 165 167 167 167 167 151 137 137 137 137 240 240 137 167 92 179 179 179 152 178 178 178 177 177 178 178 177 177 177 177 178 178 178 178 178 178 177 152 152 152 153 152 137 137 167 166 0.45 0.51 0.45 0.67 0.68 0.47 0.57 0.62 0.71 0.53 0.51 0.26 0.38 0.39 0.19 0.50 0.54 0.35 0.69 0.43 0.46 0.44 0.55 0.40 0.44 1.02 0.54 0.67 0.60 0.49 0.69 0.57 0.70 0.70 0.74 0.54 0.71 0.69 0.63 0.65 0.62 0.50 0.35 0.45 0.43 0.39 0.49 0.49 0.16 0.61 0.49 0.98 1.11 0.79 0.90 0.85 0.77 1.60 1.12 0.71 0.93 0.39 1.15 0.28 0.29 1.35 0.93 0.92 0.57 0.80 2.25 0.56 0.56 1.94 1.13 0.55 0.97 1.93 1.19 0.93 0.65 1.00 0.78 1.25 0.35 0.99 0.58 1.08 0.89 0.93 0.96 1.38 1.53 0.86 1.18 1.23 0.94 0.40 1.51 1.63 1.34 0.78 18.2 18.4 19.0 17.4 17.8 18.7 17.4 18.7 19.2 19.1 19.0 16.6 18.9 19.6 16.5 17.5 17.5 16.2 18.6 14.5 19.4 19.7 15.5 18.3 20.6 19.0 16.8 18.1 19.0 20.1 18.1 19.5 17.9 16.7 18.5 20.0 16.8 19.3 19.2 19.2 17.6 17.7 18.3 17.5 17.4 18.9 19.8 16.3 16.1 17.9 19.6 17 Table A.7 (continued) – Outstanding unknown objects observed at the INT. Acronym a e i MOID H pos arc rms R β ǫ µ 4 5 5 5 5 4 5 3 4 4 4 5 5 5 5 5 4 4 7 6 7 5 7 8 4 6 3 5 5 3 6 3 4 4 4 3 4 4 3 5 5 4 3 5 5 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 1d 0.61 0.12 0.52 0.31 0.30 0.36 0.24 0.17 0.44 0.31 0.36 1.04 0.90 0.46 0.30 0.56 0.28 0.51 0.50 1.27 0.80 0.47 0.17 0.11 0.62 0.32 0.17 0.38 0.21 0.20 0.31 0.43 0.43 0.19 0.23 0.19 0.24 0.44 0.26 0.46 0.14 0.38 0.12 0.54 0.22 19.9 20.1 20.5 20.5 20.2 19.9 20.6 20.8 20.6 20.6 20.4 20.3 20.1 20.7 20.7 21.0 21.1 20.5 21.3 21.2 21.4 20.3 20.0 18.9 20.2 19.8 20.2 20.2 19.9 19.9 20.3 18.9 20.5 20.6 20.6 20.6 20.7 20.6 20.6 20.5 20.5 20.7 20.7 20.9 20.1 -1 -1 -1 -1 -1 -1 -1 -1 -1 -1 10 10 10 10 10 10 10 10 0 0 0 -13 4 4 1 6 7 7 1 7 41 -9 3 3 3 3 3 3 3 3 3 8 8 8 16 166 166 166 166 166 166 167 167 167 167 151 151 151 151 151 151 151 151 162 162 162 102 183 183 189 182 182 183 189 183 229 173 202 202 202 202 202 202 202 202 202 208 208 208 187 0.55 0.57 0.52 0.57 0.48 0.50 0.57 0.62 0.71 0.53 0.40 0.45 0.45 0.49 0.38 0.44 0.47 0.60 0.50 0.58 0.64 0.52 0.63 0.63 0.71 0.70 0.60 0.51 0.60 0.63 0.67 4.61 0.52 0.39 0.50 0.51 0.52 0.56 0.72 0.59 0.90 0.40 0.43 0.33 0.53 VTD047 VTD049 VTD051 VTD052 VTD053 VTD054 VTD055 VTD056 VTD057 VTD058 VTD059 VTD060 VTD061 VTD062 VTD063 VTD064 VTD066 VTD069 VTD070 VTD071 VTD072 VTU004 VTU005 VTU007 VTU009 VTU011 VTU012 VTU014 VTU016 VTU019 VTU020 VTU021 VTU022 VTU023 VTU024 VTU025 VTU026 VTU027 VTU028 VTU029 VTU031 VTU032 VTU034 VTU036 VTU037 2.45 2.31 3.03 2.39 3.06 3.06 2.61 2.38 2.24 2.44 2.57 2.66 2.51 2.23 2.03 3.11 2.31 3.01 2.35 2.44 2.51 2.49 2.40 2.31 2.34 2.27 2.69 2.38 2.57 2.54 1.84 1.03 2.43 2.31 2.47 2.41 2.36 2.43 2.39 2.20 1.99 2.30 2.26 2.62 2.99 0.09 0.21 0.03 0.17 0.23 0.05 0.00 0.11 0.24 0.21 0.12 0.46 0.31 0.17 0.16 0.51 0.20 0.52 0.18 0.17 0.35 0.06 0.14 0.23 0.20 0.03 0.12 0.29 0.01 0.05 0.14 0.02 0.07 0.23 0.16 0.03 0.01 0.16 0.46 0.15 0.03 0.16 0.07 0.16 0.03 3.0 3.5 8.1 0.8 2.3 3.6 2.3 6.0 5.4 0.7 6.9 4.7 9.2 6.3 4.8 4.8 7.2 10.2 2.2 7.4 3.9 11.3 2.6 4.1 3.4 6.6 6.6 2.8 2.4 7.6 22.0 2.1 2.1 1.2 1.9 2.5 2.8 9.9 7.2 3.5 19.6 4.6 4.9 6.1 11.2 1.22 0.83 1.97 0.99 1.36 1.92 1.60 1.12 0.71 0.93 1.25 0.46 0.77 0.87 0.72 0.53 0.85 0.51 0.93 1.01 0.66 1.35 1.07 0.79 0.89 1.20 1.39 0.69 1.53 1.41 0.58 0.02 1.27 0.77 1.09 1.33 1.33 1.03 0.30 0.86 0.92 0.93 1.12 1.22 1.93 17.6 18.9 16.5 18.8 15.1 16.1 17.4 18.7 19.2 19.1 17.0 18.9 16.7 17.1 19.4 19.3 19.3 19.4 19.7 17.5 19.7 16.2 18.4 18.1 19.0 17.4 17.6 19.8 16.8 16.7 18.3 26.4 17.1 19.5 16.9 17.8 17.9 18.5 17.4 18.9 18.8 18.8 18.3 18.2 16.2 18 1 1 0 2 g u A 9 2 . ] P E h p - o r t s a [ 1 v 0 8 7 5 . 8 0 1 1 : v i X r a EURONEAR - Recovery, Follow-up and Discovery of NEAs and MBAs using Large Field 1-2m Telescopes✩ O. Vaduvescua,b,c,d,∗, M. Birlanb,e, A. Tudoricaf,g,h,i, A. Sonkaj,k, F. Pozo N.c, A. Barr D.c, D. J. Asherl, J. Licandrod,m, J. L. Ortizn, E. Unda-Sanzanac, M. Popescub,k,o, A. Nedelcub,e, D. Dumitruh,k, R. Tomag,h,p, I. Comsaq, C. Vanceah, D. Vidicank, C. Opriseanuk, T. Badescuh, M. Badeah, M. Constantinescuk aIsaac Newton Group of Telescopes (ING), Apartado de Correos 321, E-38700 Santa Cruz de la Palma, Canary Islands, Spain bIMCCE, Observatoire de Paris, 77 Avenue Denfert-Rochereau, 75014 Paris Cedex, France cInstituto de Astronom´ıa, Universidad Cat´olica del Norte (IA/UCN), Avenida Angamos 0610, Antofagasta, Chile dInstituto de Astrof´ısica de Canarias (IAC), C/V´ıa L´actea s/n, 38205 La Laguna, Spain eAstronomical Institute of the Romanian Academy, Cutitul de Argint 5, Bucharest 040557, Romania fBonn Cologne Graduate School of Physics and Astronomy, Germany gArgelander-Institut fur Astronomie, Universitat Bonn, Auf dem Hugel 71 D-53121 Bonn, Germany hUniversity of Bucharest, Department of Physics, CP Mg-11, Bucharest Magurele 76900, Romania iInstitute for Space Sciences, Bucharest - Magurele, Ro-077125 Romania jAstronomical Observatory "Admiral Vasile Urseanu", B-dul Lascar Catargiu 21, Bucharest, Romania kBucharest Astroclub, B-dul Lascar Catargiu 21, sect 1, Bucharest, Romania lArmagh Observatory, College Hill, Armagh BT61 9DG, UK mDepartamento de Astrof´ısica, Universidad de La Laguna, E-38205 La Laguna, Tenerife, Spain nInstituto de Astrof´ısica de Andaluc´ıa (IAA), CSIC, Apt 3004, 18080 Granada, Spain oPolytechnic University of Bucharest, Faculty of Applied Sciences, Department of Physics, Bucharest, Romania pRomanian Society for Meteors and Astronomy (SARM), CP 14 OP 1, 130170, Targoviste, Romania qBabes-Bolyai University, Faculty of Mathematics and Informatics, 400084 Cluj-Napoca, Romania Abstract We report on the follow-up and recovery of 100 program NEAs, PHAs and VIs using the ESO/MPG 2.2m, Swope 1m and INT 2.5m telescopes equipped with large field cameras. The 127 fields observed during 11 nights covered 29 square degrees. Using these data, we present the incidental survey work which includes 558 known MBAs and 628 unknown moving objects mostly consistent with MBAs from which 58 objects became official discoveries. We planned the runs using six criteria and four servers which focus mostly on faint and poorly observed objects in need of confirmation, follow-up and recovery. We followed 62 faint NEAs within one month after discovery and we recovered 10 faint NEAs having big uncertainties at their second or later opposition. Using the INT we eliminated 4 PHA candidates and VIs. We observed in total 1,286 moving objects and we reported more than 10,000 positions. All data were reduced by the members of our network in a team effort, and reported promptly to the MPC. The positions of the program NEAs were published in 27 MPC and MPEC references and used to improve their orbits. The O–C residuals for known MBAs and program NEAs are smallest for the ESO/MPG and Swope and about four times larger for the INT whose field is more distorted. For the astrometric reduction, the UCAC-2 catalog is recommended instead of USNO-B1. The incidental survey allowed us to study statistics of the MBA and NEA populations observable today with 1–2m facilities. We calculate preliminary orbits for all unknown objects, classifying them as official discoveries, later identifications and unknown outstanding objects. The orbital elements a, e, i calculated by FIND ORB software for the official discoveries and later identified objects are very similar with the published elements which take into account longer observational arcs; thus preliminary orbits were used in statistics for the whole unknown dataset. We present a basic model which can be used to distinguish between MBAs and potential NEAs in any sky survey. Based on three evaluation methods, most of our unknown objects are consistent with MBAs, while up to 16 unknown objects could represent NEO candidates and four represent our best NEO candidates. We assessed the observability of the unknown MBA and NEA populations using 1m and 2m surveys. Employing a 1m facility, one can observe 1 today fewer unknown objects than known MBAs and very few new NEOs. Using a 2m facility, a slightly larger number of unknown than known asteroids could be detected in the main belt. Between 0.1 and 0.8 new NEO candidates per square degree could be discovered using a 2m telescope. Keywords: minor planets, near Earth asteroids, main belt asteroids, orbits, astrometry, follow-up, survey, discovery 1. Introduction Near Earth Asteroids (NEAs) are defined as minor planets with a perihelion distance q ≤ 1.3 AU and an aphelion distance Q ≥ 0.983 AU (Morbidelli et al., 2002). Potentially Hazardous Asteroids (PHAs) are NEAs having a minimum orbital intersection distance MOID ≤ 0.05 AU and an absolute magnitude H ≤ 22 mag (Bowell and Muinonen, 1994). Virtual Impactors (VIs) represent objects for which the future Earth impact probability is non-zero according to the actual orbital uncertainty (Milani and Gronchi, 2009). According to present data (e.g. Bowell, 2011), there are more than half a million orbits of known Main Belt Asteroids (MBAs) and about 7,600 catalogued NEAs of which about 1,200 are PHAs (NASA, 2011) and ca 100 VIs (NEODyS, 2011). During the last two decades the total numbers of discovered NEAs and PHAs have continued to grow, mainly thanks to five dedicated surveys led by the United States (CSS, LINEAR, Spacewatch, LONEOS and NEAT) which have been using large field mostly 1m class telescopes. Together, they have discovered 86.8% of the entire NEA population known today (MPC Jan 2011 database), while Europe accounts for less than 1% (led by La Sagra and Crni Vrh mostly run by amateurs). Some European initiatives were taken by national institutions or local collaborations (ASIAGO/ADAS in Italy and Germany, CINEOS in Italy, KLENOT in the Czech Republic, NEON in Finland) and a few programs to study physical properties of NEAs have been carried out by groups in Europe (led by P. Pravec in the Czech Republic, SINEO led by M. Lazzarin in Italy, another program led by J. Licandro in Spain, etc). Thus although there is still no common European program and no dedicated telescope to observe NEAs, the existence of this expertise, in addition to extensive observational facilities, evidently provides an opportunity for Europe to improve its number of discoveries. In spite of the larger facilities apparently available today, extremely few researchers have used 2m class or larger facilities to observe fainter NEAs. During two short runs at ESO La Silla, Boattini et al. (2004) employed the ESO/MPG 2.2m as a search facility and the NTT 3.5m as a follow-up telescope to survey faint NEAs and MBAs beyond 22nd magnitude. During three nights using the ESO/MPG facility, the authors incidentally observed about 700 MBAs as faint as R = 21.5 mag (V ∼ 22 mag) exposing between 60 and 150 sec in the R band. Using only four hours in override service mode at the Yepun VLT 8.2m telescope, Boattini et al. (2003) eliminated 4 very faint VIs (22 < V < 25 mag), shifting them to simply the NEA or PHA class. Although the known NEA and PHA populations have increased during the last few decades, the annual growth of the number of known NEAs and PHAs appears to be becoming constant during recent years (EARN, 2011), probing some size threshold due to the limiting magnitude ✩ Based on observations taken with the telescopes ESO/MPG 2.2m in La Silla (ESO Run number 080.C-2003), Swope 1m in Las Campanas (CNTAC 2008) and the INT 2.5m in La Palma (CAT DDT 2010). ∗Corresponding author Email address: [email protected] (O. Vaduvescu) Preprint submitted to Planetary and Space Science May 21, 2018 V ∼ 21 mag reached by the present 1m class surveys. It is unclear whether new 2m surveys such as the Spacewatch 1.8m and especially the new Pan-STARRS PS1 1.8m (although not entirely dedicated to NEAs) will make a significant contribution to the completeness of NEAs through the small size objects. During the last four years, the European Near Earth Asteroids Research (EURONEAR) pro- gram has observed to date 234 program NEAs, defined as NEAs specifically planned to be ob- served according to a few selection criteria to be discussed below. Throughout our program, we used 10 mostly 1m class telescopes, in visiting mode, contributing with follow-up astrom- etry and recovery of some important NEAs, PHAs and VIs, allowing their orbits to be secured or improved (Birlan et al., 2010). Part of this work, three telescopes in Chile and the Canaries, represent our largest facilities employed, namely the ESO/MPG 2.2m in La Silla, Swope 1m in Las Campanas, and the INT 2.5m in La Palma, all equipped with large field cameras. To take advantage of these facilities, incidentally to our main program NEA work, we have identified many known MBAs in the observed fields and have also discovered many new objects. In this paper we review our EURONEAR observations at ESO/MPG and Swope and we present our new observations at the INT. Besides presenting our main NEA follow-up program, we introduce and discuss our incidental survey work using the observed fields, classifying all the observed sources as known MBAs, unknown MBAs and NEA candidates. By comparing the statistics obtained from our survey on the three 1–2m facilities, and also with those obtained from other authors using 4m and 8m telescopes, we assess the observability limits of the MBA and NEA populations observable with 1m and 2m telescopes. In Section 2 of the paper we ex- plain the basic principles driving the planning, observations and data reduction of our runs. In Section 3 we present the results, including program NEAs, known MBAs and other unknown objects. In Section 4 we discuss the results, focusing on the known and unknown MBA and NEA populations, comparing all three facilities and presenting some statistics. The conclusions are presented in Section 5. 2. Observations and Data Reduction As part of the EURONEAR project, between 2008 and 2010 we obtained three runs and a total of 11 nights for proposals devoted to the recovery and follow-up of some important NEAs, PHAs and VIs using the ESO/MPG and Swope telescopes in Chile and the INT in La Palma. For each run, the targets were selected based on the daily updated known NEA population. 2.1. Planning the Runs To search and prioritize the objects, we have used four planning servers with which one can check on a daily or hourly basis the updated NEA database. To complement the existing sur- veys focused on discovery, we focused our EURONEAR runs on some important objects in need of orbital improvement, especially on newly discovered faint asteroids which need to be secured against loss and also on other older objects having a short observed arc which needs to be improved at the second or a later opposition. We selected our targets taking into account the following six criteria: 1. Object class: observe objects classified as VIs, PHAs or NEAs (in this order), to improve their orbits and confirm or change their classification; 2. Time interval from discovery: secure and follow-up poorly observed objects, a few days or weeks from discovery; 3 3. Number of oppositions: recover objects previously observed at very few oppositions (espe- cially one-opposition objects); 4. Object brightness: recover and follow-up faint and very faint objects, accessible only to larger facilities (larger than 2m) in danger of being lost by current surveys; 5. Positional uncertainty: recover and follow-up poorly observed objects having large posi- tional uncertainty (up to one degree), less accessible to other smaller field facilities; 6. New object confirmation: recover newly discovered objects, preferably a few to several hours after their discovery; To implement these criteria, we have used the following four planning servers: 1. EURONEAR Planner 1: queries the Spaceguard database for mostly newly discovered ob- jects; 2. EURONEAR Planner 2: queries the MPC Bright and Faint Recovery Opportunity databases for mostly older objects in need of being recovered at a new opposition; 3. MPC NEO Confirmation Page: includes 1-night objects in need of being confirmed by independent observers; 4. NASA/JPL Close Approaches List: includes closest approaches of old and new objects visible from Earth; Both EURONEAR planner web-services are accessible online (EURONEAR, 2011), being written in PHP by our team and offered to the community for planning other NEA follow-up campaigns. The servers query the current Spaceguard or MPC databases (both updated on a daily basis) and return the prioritized observing lists given the observing place, facility and observing date. The planning is based on eight calculated observability factors, namely: the asteroid class (according to MPC), the apparent magnitude, proper motion, ephemeris uncertainty, altitude, star density in the field, and Moon illumination and distance, calculated with a time step (e.g., one hour) in a given time interval (e.g., one night). The result consists of a few tables listing at each step the recommended observable objects prioritized according to the object apparent magni- tude, its altitude (or airmass), proper motion, sky plane error, or some proposed "Observability" factor calculated as the product of all the above individual observability factors. Two accurate ephemeris servers, namely NEODyS (NEODyS, 2011) and IMCCE (IMCCE, 2011), are queried by the planning server automatically, returning sky coordinates, magnitudes and uncertainties according to the observed orbital arc. The response time for both servers is short, usually less than one minute for one night's information. 2.2. Observations We present here our observing runs at ESO/MPG in La Silla, Swope in Las Campanas and INT in La Palma. 2.2.1. ESO/MPG Observing Run During 3 nights from 10–13 Mar 2008 we used the ESO/MPG 2.2m telescope at ESO La Silla, Chile to observe 15 program NEAs, PHAs and VIs. At the Cassegrain F/5.9 focus of the telescope we used the Wide Field Imager (WFI) which consists of a 2 × 4 mosaic of CCDs 2K × 4K pixels each, covering a total field of view of 34′ × 33′ with a pixel scale of 0.24 ′′/pix. Due to the relatively high proper motion of NEAs at opposition (around 2 − 3 ′′/min), the average seeing of about 1 − 1.5′′, and taking into account the large raw image size on disk (140 4 MB) combined with the relatively slow readout time compared with our fast planned cadence, we observed the entire run in binning mode 3 × 3 (0.71′′ pixel size). We used an R band filter for the entire run. This binning does not affect the quality of our astrometry (set by the goal of having astrometric errors less than 0.3′′, i.e., comparable with the star catalog reference), as can be observed from the statistics in Section 3. To take advantage of the large MPG and WFI facility, we focused our run on two aspects, namely to follow-up some important NEAs, and to discover and recover many MBAs appearing in the observed fields. The weather was clear all three nights, with only two hours lost due to high humidity at the beginning of the first night. The sky was dark, with the Moon 3–6 days past new. In Table A.1. of our past paper (Birlan et al., 2010) we listed the 15 observed NEAs during our ESO/MPG run, 6 VIs, 4 PHAs and 5 other NEAs. Besides the 15 NEA program fields, during the following available nights we observed nine neighbouring fields, in order to secure some MBAs discovered in the previous nights. The neighbouring fields were chosen assuming a proper motion of 0.7 ′′/min for MBAs observed near opposition. During the second night only, we also surveyed 8 WFI fields (2.5 square degrees) in the ecliptic, about 50◦ from opposition to avoid crowding from the Milky Way. Besides the program NEAs, we identified and measured all moving sources in all the observed fields, reporting all known and new objects visible up to V ∼ 22 mag. During all three nights at ESO/MPG, we observed in total 42 WFI fields covering about 13 square degrees. 2.2.2. Swope Observing Run During 5 nights on 18–19 Oct and 22–24 Oct 2008 we used the Swope 1m telescope in Las Campanas Observatory (LCO), Chile, to observe 50 program NEAs. At the Cassegrain F/7 focus of the telescope we used the S IT e#3 2K × 3.6K pixel camera giving a field of 15.1′ × 26.5′ with a pixel scale 0.43 ′′/pix. We used an R band filter and no binning for the entire run. The sky was gray (Moon up to 4 days from last quarter) and the weather was very good, with seeing around 1′′. In Table A.1. of our past paper (Birlan et al., 2010) we listed our 50 observed NEAs during the Swope run, 12 PHAs and 38 other NEAs. Besides the program NEAs, we identified and measured all moving sources in the observed fields, reporting all known and new objects visible up to R ≤ 20.4 mag. During five nights we observed in total 50 Swope fields covering some 6 square degrees. 2.2.3. INT Observing Runs During two discretionary nights (D-nights) on 12 Feb and 15 Apr 2009 (3+3 hours), one hour in 13/14 Nov 2009 and 4 nights awarded by the Spanish Director's Discretionary Time (DDT) in 20–25 Feb and 3 Mar 2010, we used the INT 2.5m telescope in Roque de Los Muchachos Observatory (ORM) in La Palma to observe 35 program NEAs, namely 1 VI, 13 PHAs and 21 other NEAs. At the prime focus of the INT we used the Wide Field Camera (WFC) which consists of 4 CCDs 2K × 4K pixels each, covering an L-shape 34′ × 34′ with a pixel scale of 0.33 ′′/pix. Both D-nights and 13/14 Nov 2009 were observed without binning, while the Feb- Mar 2010 run was observed with 2 × 2 binning (0.66 ′′/pix) to minimize the readout time and match the poor weather conditions. We used an R filter for all runs. Most of the INT time was bright and gray, with only three hours dark time. The weather was good during the first 6 hours on the first two D-nights and one hour in 13/14 Nov 2009, but very bad (wet and windy) during the whole DDT run (Feb-Mar 2010) when the Moon was 5 gray and bright. In Table 1 we list the INT program NEAs, including the object classification, proper motion, 3σ positional uncertainty (according to the MPC for the observing date) and the orbital arc length at the observing date. Besides the program NEAs, at the INT we identified and measured all moving sources in all observed fields, reporting all known and new objects visible up to R ∼ 21.2 mag. In total at the INT, we observed for about three clear nights, a total of 35 WFC fields covering some 10 square degrees. 2.3. Data Reduction For all runs, we processed the data within 2–3 days from observations using an IRAF pipeline (for the INT and Swope data) and IDL (for ESO/MPG data), taking into account the usual sub- traction of the appropriate bias and sky flat field. For the WFI and WFC mosaic cameras we sliced CCD images and treated them independently, calculating CCD centres based on the point- ing of the telescope (included in the image headers) and the geometry of the two cameras. All images were processed on-site, then archived and eventually transferred via FTP to the remote available data reducers, allocated on demand. In order to secure some unknown objects, we observed at ESO/MPG nine fields in multiple nights, seven imaged during two nights and two fields during three nights. To identify objects observed in multiple nights, we extrapolated in time the arcs of all moving sources observed during the first night using a least square fitting code written by our team. By comparing the extrapolated positions of objects observed in one field during the first night for the observing date corresponding to the second night with the positions of the objects observed in the follow- up field during the second night, one could easily pair objects in the α − δ position space. In case of crowding and close matches, the magnitude represents a second indicator. Using this pairing technique we matched all the unknown objects observed during multiple nights, namely 43 objects observed during two nights and 8 objects observed during 3 nights. Thanks to this confirmation, most of these objects became credited discoveries. Figure 1 presents an example of such identification of objects observed in a crowded field (9◦ from opposition and 4◦ south of the ecliptic). With cyan circles we plot predicted positions and with magenta circles observed positions. The reduced images were analysed and measured using Astrometrica (Raab, 2011), carefully visually blinking all images of the same field in order to detect all moving objects. Around 100 reference stars (UCAC-2 for Swope data or USNO-B1 for ESO/MPG and INT data) were used for each CCD to perform the astrometry, using a linear model in the cases of WFI and Swope known to have small field distortion and fitting a 3-degree polynomial in the case of WFC to accommodate its larger optical field distortion. According to the literature and also to our derived astrometry (Section 3.1), the astrometric errors due to the field distortion appear very small for the WFI field, being mostly within 0.1′′ across most of the WFI field (about 90%), according to the field distortion pattern derived by Assafin et al (2010) which shows the largest distortion in the corners of the mosaic (up to 0.53′′) and some distortion around 0.2′′ in the upper and lower edge and the centre of the mosaic. Relative photometry was derived by Astrometrica using from a dozen to a few hundred catalog stars visible in the field, the reduced asteroid magnitudes having an uncertainty of about 0.1 mag. For all runs, we decided to use Astrometrica in preference to other software because of its simplicity, common platform and simple installation and usage by all members of the team, as many of the data reducers were students and amateurs. We inspected the data visually instead of using automated software, because of the relatively low volume of data per run, and mostly because the human eye and brain are known to detect faint moving 6 sources better than the computer. We detected asteroids as low as ca 1.5σ level from noise, which allowed us to recover many faint targets inaccessible to other automated surveys. All the reduced data (output of Astrometrica in MPC format) were collected by the PI of the run, checked for errors using the EURONEAR O–C calculator or the FITSBLINK residual calculator (Skvarc, 2010a), then submitted to the MPC in three groups: the observed NEAs, the known asteroids, and the unknown asteroids (possibly discovered by us). 3. Results During all three runs, we observed effectively in total about 11 nights, reporting positions and magnitudes for 100 program NEAs (7 VIs, 29 PHAs and 64 NEAs), 558 known MBAs and 628 unknown moving objects, in total 1286 objects observed in 29 square degrees total surveyed field. Table 2 presents the overview of our observations at ESO/MPG, Swope and INT, listing the number of observed objects, number of reported positions and the standard deviation of the or- bital fit using our data (according to NEODyS). We classify the objects in three groups: observed NEAs, known MBAs, and the unknown (unidentified) objects. We further classify the unknown objects in three groups: official discoveries, later identifications and outstanding objects. We in- clude the number of nights observed at each facility, the number of observed fields, the total sky coverage (in square degrees) and the limiting magnitude for each facility. Next, we give the total number of objects. Finally, we conclude with some density statistics which will be discussed in Section 4.4. 3.1. Program NEAs In the left panels of Figures 2, 3 and 4 we plot the O–C residuals (observed minus calculated) in α and δ for the program NEAs observed at ESO/MPG, Swope and INT, based on the orbital fits available on 22 Nov 2010 (NEODyS, 2011). The O–C standard deviations are 0.15′′ for the ESO/MPG dataset, 0.39′′ for Swope and 0.42′′ for the INT. In total, 27 MPC and MPEC publications include data from our three runs. Table 3 includes these references: 12 publications containing our Swope observations, 10 our ESO/MPG obser- vations and 6 publications including our INT data. Birlan et al. (2010) presented the most important NEAs recovered at ESO/MPG and Swope. At ESO/MPG we observed 12 faint objects (6 VIs, 3 PHAs and 3 NEAs) less than one month after discovery, and also recovered 2 faint objects (1 PHA and 1 NEA) at their second or later opposition. With Swope we observed 25 objects (all NEAs) soon after discovery, recovering also 2 objects (one NEA and one PHA) at their second or later opposition. From Table 1 we can count the number of recoveries using the INT: 25 objects soon after discovery (1 VI, 8 PHAs and 16 NEAs) and 6 faint objects having large uncertainty recovered at their second or later opposition (2 PHAs and 4 NEAs). Nine especially important NEAs were observed with the INT. We mark them with * in the first column of Table 1 and we discuss them next. Thanks to the the large aperture of the INT and the large field of the WFC, we eliminated 3 NEA candidates and 1 VI. First classified as a NEA, 2009 CB2 had a poor orbit (2 day arc) and a very high sky uncertainty (3σ = 1000′′). Thanks to the large field of WFC, we recovered this object one week later, allowing its orbit to be improved and eliminating it from the NEA list. A similar case was 2010 CF12, originally classified as a NEA based on a small 3 day arc. Although the MPC did not list its sky uncertainty by the time 7 of our INT run, we recovered this object one week later and eliminated it from the NEA list. Another object degraded from the NEA class to the MBA class was 2010 DC. It had a small arc based on 4 nights data and a sky uncertainty of about 1′, allowing fast recovery two weeks later at the INT. One of the most important detections of the INT was 2009 VR25, a new object classified as a VI based on its original poor two night orbit. Although quite faint (V = 20.7 mag) and having very large sky uncertainty (3σ = 1400′′), we could find it, enabling its reclassification from the VI to NEA class. We recovered 3 NEAs at their second opposition with the INT. 2007 RM133 was discovered in 2007, having a short arc (one month) at the time of the INT run. Despite its very faint mag- nitude (V = 21.4 mag) and very large sky uncertainty (3σ = 490′′), we recovered it 3 years later, allowing a substantial improvement of its orbit (Holvorcem et al, 2010). Another second opposition recovery was 2003 SJ84, another NEA observed for only one month, following its discovery in 2003. Although very faint (V = 21.4 mag) and having quite large positional un- certainty (3σ = 81′′), we recovered it six years later about 21′ away from its predicted position (15 times more than its nominal MPC 3σ value) and we improved its orbit (Fitzsimmons et al., 2009). Our best ever second opposition recovery was 2000 SV20. This NEA was observed for 3 months following its discovery in 2000 and recovered 10 years later by the INT/WFC about 7′ away from its predicted MPC position (7 times more than the nominal MPC 3σ ∼ 1′ value). Using the INT, in 2010 we followed two Arecibo radar targets. 2010 DJ1 was requested by NASA in Feb 2010 (Benner, 2010), having an uncertainty (3σ = 190′′) which allowed its INT recovery only three days after discovery. 2007 EF was another Arecibo target, moving very fast (µ = 20 ′′/min) but allowing successful WFC imaging at V = 18.2 mag in only 10 sec exposures. 3.2. Known MBAs During all three runs in the observed program NEA fields, we observed incidentally a total of 558 known MBAs. In the right panels of Figures 2, 3 and 4 we plot the O–C residuals for the known MBAs observed in all three runs based on their current orbits available at 22 Nov 2010 (AstDyS, 2011). The standard deviations are 0.15′′ for the ESO/MPG dataset, 0.18′′ for Swope and 0.66′′ for the INT. The standard deviations of the ESO/MPG dataset for both NEAs and MBAs are very small, proving the excellent quality of this telescope equipped with the WFI which allowed very ac- curate astrometry across its whole large field. Nevertheless, a systematic offset to the north of 0.111′′ ± 0.001′′ in δ and 0.037′′ ± 0.001′′ in α shows up in the right panel of Figure 2 for the known MBA dataset observed at ESO/MPG, with the observed median position located north- west of the calculated ephemerides. This effect is consistent with results of Tholen et al. (2008a) who found a surprising systematic offset of the astrometry of the asteroid Apophis of about 0.2′′ based on 200 observations reduced with the USNO-B1 catalog (which was also used by us to reduce our ESO/MPG data). By comparing USNO-B1 with ICRF, the authors determined for the USNO-B1 an average declination offset of +0.116′′, in perfect agreement with our findings. The standard deviation of the Swope datasets is 0.39′′ for the NEA data and 0.11′′ for the MBA data, and no sample shows any systematic offset with respect to the origin. We reduced Swope data with the UCAC-2 catalog, known to have better astrometry than USNO-B1. Both for Swope and ESO/MPG runs, the standard deviation for the MBA sample is smaller than that of the NEA sample probably because of the higher S/N due to brighter and slower moving MBA objects, compared with the fainter and faster moving program NEAs. 8 The INT/WFC shows relatively large residuals for both NEA and MBA samples (up to about 2′′ and standard deviation 0.66′′ as quoted above) and only for this facility the deviation of the MBA sample is larger than that of the program NEAs. Bad weather did not impede the astrometry for our INT run, catalog stars being quite bright in a 2m telescope resulting in good S/N and sub- arcsec stellar positions. The USNO-B1 catalog has an average astrometric accuracy of 0.2′′, too small to explain the larger residuals of the INT astrometry. The larger residuals for known MBAs than for program NEAs observed with INT/WFC can however be explained taking into account that MBAs were imaged across the entire WFC field which is heavily affected by the field distortion at the INT prime focus, compared with the central CCD#4 where most of the NEAs were observed and the WFC field has the smallest distortion. Given this, for any future INT/WFC work we plan to correct the image field before data reduc- tion. 3.3. Unknown Objects A total of 467 unknown moving objects were identified at ESO/MPG, 41 at Swope and 120 at the INT (Table 2). The observed proper motion for most objects was compatible with MBAs with a few exceptions discussed next. Throughout our paper and in Table 2 we classify the unknown objects in three categories: of- ficial discoveries (confirmed by the MPC according to their DISCSTATUS monthly list), later identifications (unknown objects which could be linked with existing arcs) and outstanding ob- jects (waiting for orbital links from independent observations, possibly to be credited to us later). 3.3.1. Unknown MBAs We include in the Appendix A (available only in the online version of our paper) seven tables listing all the unknown objects observed at ESO/MPG (Tables A.1, A.2 and A.3), Swope (Tables A.4 and A.5) and the INT (Tables A.6 and A.7). We give first the official discoveries, then later identifications and finally outstanding objects. Table A.1 includes our official discovered asteroids at ESO/MPG. We give first the EU- RONEAR object acronym (based on the initials of the surname of the observers and reducers1), then the official designation (from the MPC), three main orbital parameters (semimajor axis a, eccentricity e and inclination i), Earth minimum orbital intersection distance MOID, absolute magnitude H, number of observed positions, arc length (in days or years), rms of the O–C resid- uals for the orbital fit σ, observed apparent magnitude R, ecliptic latitude β, Solar elongation ǫ (both in degrees) and apparent proper motion µ (in ′′/min). We distinguish directions east and west of opposition by letting ǫ increase above 180◦ for fields to the east. For each object we give in the first line the orbital elements fitted with the FIND ORB software (Gray, 2011a) based on our observations only and for the standard epoch MJD = 54520.0. On the second line we list the orbital elements taken from the MPC database (MPC, 2011) calculated by fitting all MPC available observations for an epoch close to the mid-point interval of those observations. Table A.2 includes the unknown objects (at the date of the run) which were identified later (in Nov 2010) with known objects, based on the checks of the MPC database and the MPC automatically assigning designations. Our calculated orbits, presented again in the first line, are 1VB = Vaduvescu and Birlan; TU = Tudorica; SO = Sonka; OP = Opriseanu; VI = Vidican; TO = Toma; VA = Vancea. 9 based on the very short available arc (observations acquired in less than one hour), and so should be regarded with caution. For both official discoveries and later identified objects, the orbital elements from the first line are very close to the official elements in the second line. Although most of our fits are based on very short arcs observed during 1–3 nights at ESO/MPG and only one night at Swope and INT, one can observe the success of the FIND ORB fit for most objects, especially for the a, e and i parameters which will be used in statistics later. In particular, based on 116 paired orbits available from the three runs, we can compare FIND ORB versus MPC by calculating the median values of the differences in a, e, i and H, which are 0.20 AU, 0.06, 1.05◦ and 0.70 mag, respectively. Table A.3 includes the remaining unknown objects observed at ESO/MPG which could not be identified according to the present MPC database (Nov 2010). For them we give only our calculated orbits, based on very short arcs observed only in one night. The exact elements should be regarded with caution but are usable for statistics in Section 4. Table A.4 lists the later identified objects observed with the Swope telescope, while Table A.5 gives the outstanding unknown objects observed with Swope. Similarly, Tables A.6 and A.7 give the later identified objects and outstanding unknown objects observed at the INT. According to the MPC (Jan 2011 DISCSTATUS), our ESO/MPG run produced 58 official dis- coveries. Most of the people involved in EURONEAR work on a voluntary basis and include students and amateur astronomers based in Romania. Indeed, the entire ESO/MPG team in- cluded people of Romanian origin who became the first Romanian discoverers of minor planets (Vaduvescu, 2009). Given these, in 22 December 2008 we proposed to the Working Group for Small Body Nomenclature (CSBN) of the International Astronomical Union (IAU) a list includ- ing 12 Romanian names. In Jan 2011 the first two of our discovered asteroids received numbers, namely (257005) = 2008 EW152 = VBTU207 (our acronym) and (263516) = 2008 EW144 = VBTU224, being eligible for naming. 3.3.2. NEO Candidates In this section we will use three tools to check all our unknown objects for potential Near Earth Objects (NEOs). In Table 4 we include all NEO candidates derived from all three methods, marking in bold the best candidates. Besides the observed properties (quantities µ, ǫ, R, number of positions and O–C standard deviation), we include in this table the orbital parameters (a, e, i) and data derived from the three methods (MOID in column 4, MPC score in column 10 and the Model in the last column). For the first method we plot in Figure 5 the apparent proper motion µ versus the Solar elon- gation ǫ for all the unknown objects observed at ESO/MPG (red), Swope (green) and the INT (blue). Observed in a given field near opposition (close to ǫ ∼ 180◦), MBAs are expected to show proper motions distributed in a small vertical "finger"-shaped region with proper motions between µ ∼ 0.3 − 0.7 ′′/min, depending on their location in the main belt. Observed further away from opposition, MBAs should show smaller proper motions owing to the larger distance and velocity projection effect. Both µ and ǫ represent quantities measured directly from obser- vations, not being affected by the uncertain orbits derived from the short arc. Therefore the µ − ǫ plot represents an important method to search a survey for fast moving objects including NEAs, PHAs and other NEOs. Let us consider the very basic orbital model which assumes the (prograde) asteroid orbit cir- cular and coplanar with the circular orbit of Earth, and the asteroid at least 90◦ away from the Sun. Following Kolena (1999) we express the asteroid proper motion as a function ǫ. Let ∆ be the angle between the directions of Sun and asteroid as seen from Earth (∆ is ǫ or 360◦ − ǫ, with 10 0 < ∆ < 180◦). Let vE and va be the orbital velocities of Earth and asteroid, φ the angle as seen from the asteroid between the direction of Earth and that of the asteroid's orbital motion, and E the difference 180◦ minus the angle (as seen from Earth) between the direction to the asteroid and the direction of Earth's orbital motion (so E = ∆ − 90◦). Then the angular speed of the asteroid ω as seen from Earth is the difference in the projected velocities perpendicular to the Earth–asteroid direction, divided by the Earth–asteroid distance: From the sine law applied to the triangle Sun–Earth–asteroid: ω = va sin φ − vE sin E d sin φ = s1 − E sin2 ∆ a2 a2 a The cosine rule applied to the same triangle, solving the resulting quadratic for d, gives: (1) (2) (3) d = aE cos ∆ + qa2 a − a2 E sin2 ∆ where we dropped the minus solution because d should always be positive. Kepler's third law (assuming Earth and asteroid masses very small compared with Sun's mass) implies where G = 6.673 × 10−11 m3 kg−1s−2 is the gravitational constant, MS = 1.989 × 1030 kg the Sun's mass and aE = 1.496 × 1011 m the Earth's semimajor axis. va = q GMS aa and vE = q GMS aE Substituting all these terms in Equation 1, and since sin E = − cos ∆, we obtain the following formula for the apparent angular speed of the asteroid as a function of ∆ and aa: ω = rGMS E sin2 ∆ a2 aa s1 − aE cos ∆ + qa2 a2 a + rGMS aE E sin2 ∆ a − a2 cos ∆ (4) Finally, the proper motion of the asteroid µ in arcsec per minute can be calculated from the angular speed ω in radians per second: µ = ω × 180 × 3600 π × 60 (5) We can use Equation 5 and the asteroid semimajor axis aa as a model to map the expected limits for the proper motion of MBAs (defined between aa = 2.0 and aa = 3.5 AU). In the context of this basic orbital model where we consider proper motions (as a function of elongation) only, we may also represent NEOs by considering orbits with aa < 1.3 AU. By plotting the values of µ between solar elongations 90◦ and 270◦ (e.g., using one degree step in ǫ), one can draw the limits corresponding to these populations. In Figure 5 we plot with dotted magenta lines the curves corresponding to these limits. The curves are symmetric about 180◦ (in fact symmetry properties of sky motions hold more generally in the non-coplanar case; see section 2.1 of Jedicke (1996). According to Figure 5, most of the 628 unknown objects agree with our model, in the sense that most are consistent with asteroids from the main belt. They are located at the bottom of the plot around µ = 0.2 − 0.8 ′′/min and between the two dotted curves corresponding to aa = 2.0 11 and aa = 3.5 AU. About 16 objects (2.5% of the total) marked with circles rise above the a = 1.3 NEO limit and above the main vertical group at the respective elongation. We mark these objects with "fit" or "best" in the last column of Table 4, treating them as potential NEOs. One can clearly distinguish three major outliers showing the fastest proper motions, namely VBVI213 at µ = 7.17 ′′/min and VBTU222 at µ = 2.12 ′′/min observed at ESO/MPG, and VTU021 at µ = 4.61 ′′/min observed at the INT. We mark them with "best" in the last column of Table 4, treating them as our best NEO candidates. The fastest object was recorded under the acronym VBVI213 and it moved about 10 times faster than all other MBAs observable close to opposition, so it represents our best NEO can- didate. This object was clearly visible on 8 CCD images, leaving a 20 pixel trail owing to the relatively long 2 minute exposures. It moved in the opposite direction and about 10 times faster than all other asteroids visible in the same field. In Figure 6 we include the image of this field (CCD#5 of WFI), presenting the corresponding 8 frame animation in the online electronic ver- sion of the paper. The image is displayed in normal sky orientation and the field of view is about 8′ × 16′ (one WFI CCD), with pixel size 0.714′′ (in 3 × 3 binning). Four MBAs marked with circles are visible moving to the upper right, while the NEA candidate is visible as a trail in the bottom part (enlarged twice in the left corner inset). The exposure time was 2 minutes and the cadence between frames was 3.3 minutes. Most of our fields were observed near opposition (150◦ < ǫ < 210◦) and all agree well with In Figure 5 there are about three fields located farther from opposition between our model. 130◦ < ǫ < 140◦ for which most unknown objects do not match our model which does not hold due to our basic (circular and coplanar) orbital assumptions (proper motions in the model are close to zero, whereas the real orbits give a small but noticeable component to the proper motions). Our second NEO search method uses the "NEO Rating" tool developed by the Minor Planet Center (MPC) which calculates a score for possible NEOs based on the expected proper motion of the MBA population distribution (MPC, 2011). Running this tool for all unknown ESO/MPG objects, we obtained NEO scores ("No-ID" probabilities) of 100% for three objects: VBTU197, VBTU222 and VBVI213 which confirm our findings using the first model. Running the NEO Rating for all unknown INT objects we confirm with 100% score two INT objects, VTU021 and VTD003, plus VITB01 with a relatively high score (77%). We plot with circles all these NEO candidates in Figure 5 and we include the scores in Table 4. All other objects from all runs received very low MPC rates (smaller than 5%), consistent with our MBA classification derived from our model. Our third NEO search method uses the calculated MOID derived from the preliminary orbits derived with FIND ORB. The results are included in Table 4 and they mostly agree with the other two methods, showing in most cases the reliability of the derived preliminary orbits, especially for the objects observed closed to opposition. Unfortunately most NEO candidates were observed in only one night and only one object was reobserved during a second night, namely VBTU203 = 2008 EN144. Two NEA candidates observed at INT were identified later with known objects, namely VTD003 = 2010 CJ33 and VIT005 = 2000 SN27. Based on their updated observed arcs, they are not NEAs, so we mark them by * in Table 4. Dropping them from the list, we count in total 16 NEO candidates; this number should be considered an upper bound. Four objects have MPC scores of 100%, small MOIDs (less than 0.1 AU) and they agree mostly with our model, so they represent our best 4 NEO candidates: VBVI213, VBTU222, VBTU197 (observed at ESO/MPG) and VTU021 (observed at the INT). We write their acronyms in bold in the first column of Table 4. There was 12 no NEO candidate observed with Swope, checking all three methods. We checked the remote possibility that some NEO candidates could be identified with Earth artificial satellites or space debris. In this sense, we checked all our observed fields against known Earth satellites by using the satellite identification server developed by Skvarc (2010b) based on the software SAT ID of Gray (2011b). No satellite with proper motion slower than 0.25′/min (corresponding to geostationary orbits) was found within one degree of any observed fields. Known and unknown space debris could also be studied statistically, according to Schildknecht (2007). Compared with NEO candidates, space debris move at very fast speed with angular velocities ranging from a few arc seconds per second (i.e., a few arcminutes per minute, at least 10 times faster than our fastest NEO candidate) to more than 1,000 arcseconds per second with respect to the stellar background. Thus, we drop any possibility that any of our unknown objects could be associated with artificial satellites or space debris. 4. Discussion 4.1. Comparison with the known asteroid population We compare the major orbital parameters of all unknown objects observed at ESO/MPG, Swope and INT with the entire known asteroid population at 10 Dec 2010 (541,260 orbits) based on the ASTORB database (Bowell, 2011). In Figure 7 we plot two classic asteroid orbital distributions, namely e versus a (left) and i versus a (right). We overlay in colours all the unknown objects observed in our survey, including official discoveries, later identifications (based on MPC orbits) and outstanding objects (based on FIND ORB orbits). One can easily observe that the majority of our objects fit both orbital distributions very well, marking the four major Kirkwood gaps and a few known families. Only 105 points represent our official discoveries and later identifications, while 523 points represent outstanding objects (five times more). Because outstanding objects have orbits calculated with FIND ORB, at least statistically this confirms the ability of FIND ORB to calculate preliminary orbits based on very small arcs. 4.2. Comparison between facilities In Figure 8 we plot the distribution of the observed apparent magnitude R (left panel) and the calculated absolute magnitude H versus semimajor axis a for all unknown asteroids observed with ESO/MPG, Swope and INT. The limiting magnitude of each system is evident by the levels in the R plot above which the regions become depleted of data. For each fixed limiting R, we expect a negative trend of H versus a, as seen in the right panel. We observed unknown MBAs up to a ∼ 3.3 AU, which is considered about half the outer main belt region (Yoshida & Nakamura, 2007). As expected, both 2m facilities sampled well the middle region (2.6 ≤ a ≤ 3.0) and the first half of the outer region (3.0 ≤ a ≤ 3.3), while well over half of unknown objects sampled with the Swope 1m are in the inner region of the main belt (2.0 ≤ a ≤ 2.6). Moreover, according to Table 2, both ESO/MPG and INT discovered about the same number of MBAs as the number of known MBAs. Thus, a 2m survey could bring an important contribution to knowledge of the main belt, being expected to double the present number of known MBAs to more than one million. According to the O–C plots for known MBAs (right panel of Figure 2, 3 and 4) and to the O–C standard deviation of 0.15′′ for both NEA and known MBA datasets, ESO/MPG appears to have the best astrometry required for accurate follow-up, recovery and discovery across the 13 whole WFI camera. With standard deviations of 0.18′′ for the known MBA dataset and 0.39′′ for the NEAs, the Swope telescope represents an adequate 1m facility for asteroid studies at limiting magnitude R ∼ 20.5 mag. The field of the INT/WFC appears the most distorted, these O–C positions showing the widest spread in Figure 4 and a standard deviation about 4 times larger compared with the other two facilities. Although the INT astrometry is acceptable around the centre of WFC and could be used to follow-up known objects expected to appear close to the centre, the INT field should be corrected in order to reach more accurate astrometry across the entire WFC field. Comparing the position of the centroid of the known bulk of MBAs with respect to the cal- culated positions, we conclude that the USNO-B1 catalog is less appropriate for astrometric reduction due to larger residuals, and we recommend instead UCAC-2 or UCAC-3 which appear to give more accurate results and therefore to be the current best representation of the Hipparcos frame up to magnitude 16. Nevertheless despite the superior accuracy from using the UCAC sys- tem, it still has shortcomings for the astrometry of fast moving objects such as the limited north declination coverage of UCAC-2 and the faulty northern proper motion system of UCAC-3. 4.3. Distribution of the unknown MBA and NEO candidates With the original 1m Spaceguard survey approaching its goal and limits, new 2m surveys such as Pan-STARRS will soon take over, increasing the detection limits in both size and depth in the Solar System. Based on our ESO/MPG and INT data, we briefly evaluate here the limits of such a 2m survey. The left panel of Figure 9 plots the histogram showing the observed apparent magnitude R for all the unknown objects observed at ESO/MPG (red colour), Swope (green), the INT (blue) and the total (black dots). Apparently, the dark time at ESO was most efficient to detect unknown objects at R ∼ 20.6 mag, allowing a limit R ∼ 21.5 mag. This detection limit is consistent with the actual 1.8m Pan-STARRS 1 which is expected to reach R ∼ 22 mag (Grav, 2009). The INT reached maximum detection at R ∼ 20.2 mag and a more shallow cutoff at R ∼ 21.2 mag, about 0.3 mag less faint than ESO/MPG despite its larger aperture, probably due to the worse observing conditions at the INT. The 1m Swope reached a maximum detection at R ∼ 19.8 mag and a limiting magnitude at R ∼ 20.4 mag, about one magnitude lower than other 2m facilities. We include these limits in Table 2. The right panel of Figure 9 plots the histogram showing the calculated absolute magnitude H for the unknown objects observed at the three facilities. Five objects fall outside the H range of the plot, namely the brightest object VB037 identified as the jovian Trojan 2001 TB234 (H = 13.3) and the faintest four objects VBTO016 (H = 22.7), VBTU222 (H = 24.7), VBVI213 (H = 25.3) and VTU021 (H = 26.4). The last four are visible as clear individual points in Fig- ure 8 (right) and the last three are among our best NEO candidates (Table 4). The H histograms are more evenly distributed, showing 2–3 maxima (possibly not all real) for each facility and an overall maximum at H ∼ 17.4 mag. This limit can be regarded as the limiting H giving com- pleteness for a 2m class facility for the entire main belt (including the outer region). According to Yoshida & Nakamura (2007), this limit corresponds to S-class asteroids about 1km in diame- ter, thus virtually all S-type MBAs larger than this limit should be accessible to a 2m telescope (including ESO/MPG and INT) in good weather conditions. As we saw in Figure 8 (left), R ∼ 21.6 mag represents the limiting apparent magnitude for the ESO/MPG. Most MBAs have absolute magnitudes between 15 < H < 21 mag, consistent with sizes between 170m and 6km, assuming albedos between 0.05 and 0.25 (NASA, 2011). The four best NEO candidates have the following H and sizes, assuming the same limits for their albedos: 14 VTU021: H = 26.4 mag, 13–31 m; VBVI213: H = 25.3 mag, 22–55 m; VBTU222: H = 24.7 mag, 30–70 m; VBTU197: H = 18.9 mag, 440–980 m. In the left panel of Figure 10 we plot the calculated MOID versus the elongation ǫ for the unknown objects observed at ESO/MPG (red), Swope (green) and INT (blue). With a dotted line we mark the MOID= 0.3 limit for NEAs below which all our NEO candidates appear. Based on this plot, there is no apparent favorable elongation to discover NEOs. In the right panel of Figure 10 we plot the calculated MOID versus the observed ecliptic lat- itude β for the unknown objects observed at ESO/MPG (red), Swope (green) and INT (blue). Most unknown objects and NEO candidates were observed at low latitudes, under 10◦. Most NEO candidates are observed at low latitudes and there is no particular favorable detection lati- tude with respect of the MBAs. 4.4. Survey statistics for 2m and 1m facilities Based on the statistics available from our ESO/MPG, Swope and INT surveys, we can evaluate the unknown MBA and NEA population observable at low latitudes ( β < 10◦) by 2m and 1m surveys. We include these results in Table 2. 4.4.1. Unknown MBA density Using data from our ESO/MPG survey (the best performing 2m facility) we observed in the 10◦ latitude range 347 known objects and 467 unknown objects scanning 13 square degrees. This gives an average of 27 known and 36 unknown MBAs per square degree visible to limiting magnitude R ∼ 21.5 mag in 2 min exposure time using the ESO/MPG. We include these findings at the end of Table 2. These numbers give for ESO/MPG a MBA ratio known:unknown = 0.7. We compare below our findings with other authors. Counting data from our Swope survey we observed 35 unknown objects and 65 known objects within 10◦ latitude range from the ecliptic scanning about 5 square degrees of sky. This gives an average of 11 known objects and 7 unknown objects per square degree visible to limiting magnitude R ∼ 20.4 mag in 2 min exposure time with a 1m facility. The total number agrees with earlier results from the Spacewatch 0.9m which detected for the whole survey 16 asteroids per square degree. Boattini et al. (2004) conducted a 3 night pilot search and follow-up program to detect NEAs using the ESO/MPG with WFI in 3 × 3 binning mode (i.e., same facility and setup as us). During the last two nights, the authors scanned in good weather conditions a total of 24 square degrees, counting an average of 10 known and 12 unknown asteroids (mostly MBAs) per square degree. This gives a ratio known:unknown = 0.8 which is consistent with our findings. Nevertheless, both their numbers are about three times less than our findings. Their survey strategy was a bit different than ours, namely they observed at small solar elongation during the first and last part of the night and observing near opposition during the middle part. Also, for identification they used mostly automated software, although some data were reduced with Astrometrica. It is well known that the eye and brain are better than computer software in detection of moving objects by a factor of 3/2 based on experience of Spacewatch II (Boattini et al., 2004) or by 1 − 1.5 mag according to other authors (Yoshida et al, 2003). Both these factors could explain the lower density of asteroids (mostly MBAs) found by Boattini et al. (2004) compared with our statistics. Wiegert et al (2007) searched 50 fields (50 square degrees) from the CFHT Legacy Survey (CFHTLS 3.6m) observed in r′ close to opposition and within a 2◦ latitude range from the eclip- tic. The moving objects were detected automatically by using the Sextractor software with a 15 threshold 3σ. The authors found an average of 70 asteroids per square degree up to r′ ∼ 21.5 mag (Wiegert, 2011), which agrees well with our findings. Using the Subaru 8.3m telescope equipped with the large field SuprimeCam with 7 sec expo- sures Yoshida et al (2003) surveyed 3 square degrees near opposition and the ecliptic (SMBAS I survey) and found 92 asteroids per square degree to limiting magnitude R = 21.5 mag. Us- ing the same facility to image 4 square degrees using 2 second exposures (SMBAS II survey), Yoshida & Nakamura (2007) found an average of 75 objects per square degree to the same limit. In these surveys, moving objects were detected by human inspection. Our findings using the ESO/MPG are very close to their densities, taking into account our lower S/N due to Subaru's larger aperture and their pointing at lower latitudes. Using one single 8.4m mirror of the Large Binocular Telescope (LBT), Ryan et al (2009) studied the asteroid distribution in the ecliptic, finding up to V = 22.3 mag (close to R ∼ 21.5 mag our limit) a density of 85 asteroids per square degree. Asteroid detection was performed visually using a three-color method. Their density found is very close to ours, counting our total number of objects (63 objects per square degree). 4.4.2. Unkown NEA density Counting the NEO candidates from Table 4, ESO/MPG produced 8 NEO candidates and 3 best NEO candidates scanning a field of 13 square degrees. This gives between 0.2 and 0.6 NEO candidates per square degree observable with this facility. The value 0.6 is an upper limit because we could not confirm our objects which were observed only in one night. Scanning 40 WFC fields (13 square degree) within 15◦ latitude in good weather at ESO/MPG, Boattini et al. (2004) discovered 3 NEA candidates (including 2 confirmed NEAs), which gives a density of 0.2 NEA candidates per square degree, matching our findings counting only the best candidates. Comparing their results with those from the 1.8m Spaceguard II survey, the authors conclude that on average one NEA per 10 square degrees could be discovered with ESO/MPG and the WFI. This is consistent with our findings if we count only one object, namely our best NEA candidate, VBVI213. Counting all our NEO candidates, our result is 6 times more optimistic than that of Boattini et al. (2004). According to Table 4, INT produced 8 unknown objects and only one best NEA candidate scanning a field of 10 square degrees. This gives between 0.1 and 0.8 NEO candidates per square degree observable with INT (mostly in bad conditions). These densities are similar with those found by ESO/MPG and consistent with any other 2m survey. 5. Conclusions We have analysed our observations taken with the ESO/MPG 2.2m in La Silla, the Swope 1m in Las Campanas and the INT 2.5m in La Palma. The total sky surveyed during 11 nights was about 29 square degrees, which allowed us to study statistics of MBAs and NEAs observable nowadays by other 1–2m facilities. Our main conclusions are: • These telescopes are successful at following up faint objects soon after discovery, prevent- ing their loss, recovering NEAs at their second or later opposition and eliminating NEA candidates and Virtual Impactors. • The majority of our unknown objects are consistent with MBAs, based on two evaluation methods. Up to 16 unknown objects could represent NEO candidates from which 4 repre- sent our best NEO candidates according to three evaluation methods. 16 • The O-C residuals for known MBAs and program NEAs amount to 0.15′′ for the ESO/MPG, 0.39′′ and 0.18′′ for Swope and 0.42′′ and 0.66′′ for the INT, whose prime focus field is the most distorted (especially the three non-central CCDs) and needs to be corrected in order to improve the astrometry. • The UCAC-2 catalog is better than USNO-B1 which shows an offset af 0.1′′ to the North, consistent with previous findings of other authors. • Published orbits (specifically a, e and i) of known asteroids are very similar to our calculated orbits using the FIND ORB software based on our observed very small arcs. • Based on statistics derived from our data, we could assess the observability of the unknown MBA and NEA populations using 1m and 2m class surveys. Employing a 1m facility one can observe today fewer unknown objects than known MBAs and virtually no new NEO. Using a 2m facility, a slightly larger number of unknown than known MBAs could be de- tected (up to about a = 3.2 AU), consistent with objects having sizes between 170m and 6km (taking into account the limits of the main belt and the albedo range). Between 0.1 and 0.8 new NEO candidates per square degree could be discovered using a 2m telescope. • A basic model assuming circular and coplanar orbits of the asteroids and Earth could be used in order to check any large all sky survey for potential NEO candidates. Employing the proper motion and Solar elongations, this model does not depend on calculated quantities such as orbital elements possibly subject to errors. Compared with other tools such as the MPC's NEO Rating and the calculated preliminary orbits, this model seems very accurate at small elongations (±30◦ from opposition) but, based on the residuals in our data, smaller elongations (around 120 − 140◦) need further study. 6. Acknowledgements This work was based on observations made with the ESO/MPG telescope at La Silla Obser- vatory under programme ID 080.C-2003(A), the Swope telescope at Las Campanas Observatory (CNTAC 2008), both granted under Chilean time, and the INT telescope in La Palma under Di- rector's Discretionary Time of Spain's Instituto de Astrofisica de Canarias (CAT DDT 2010). OV acknowledges ESO, LCO and IA/UCN for supporting the runs in Chile for himself and the students, and also to the ING, IAC and the IAA for supporting the run in La Palma for the stu- dents. OV and JL gratefully acknowledge support from the spanish "Ministerio de Ciencia e Innovaci´on" (MICINN) project AYA2008-06202-C03-02. This research has made intensive use of the Astrometrica software developed by Herbert Raab, very simple to install and use by stu- dents and amateur astronomers. We also used the image viewer SAOImage DS9, developed by Smithsonian Astrophysical Observatory and also IRAF, distributed by the National Optical As- tronomy Observatories, operated by the Association of Universities for Research in Astronomy, Inc. under cooperative agreement with the National Science Foundation. Special thanks are due to Bill Gray for providing FIND ORB and installation assistance. OV also acknowledges to Paul Wiegert for feedback necessary to compare MBA data observable with CFHT and the ESO/MPG telescopes. Thanks are due to Fumi Yoshida and Tsuko Nakamura for sharing their SMBAS data and interest. We also acknowledge to Jure Skvarc for his satellite identification software and to Lilian Dominguez for providing some references about space debris. Thanks are also due to Alain Maury who helped us to count NEA discoveries made in Europe. Acknowledgements are due to the referee whose constructive suggestions helped us to improve the paper. 17 References Assafin, M., et al., 2010, Astronomy & Astrophysics 515, A32 AstDyS, 2011, Asteroids Dynamic Site, http://hamilton.dm.unipi.it/astdys/ Benecchi, S.; Sheppard, S. S.; Vaduvescu, O.; Pozo, F.; Barr, A.; Tudorica, A.; Sonka, A., 2008, MPC 64096, 9 Benecchi, S.; Sheppard, S. S.; Vaduvescu, O.; et al., 2009, MPC 64484, 5 Benecchi, S.; Sheppard, S. S.; Vaduvescu, O.; et al., 2009, MPC 64753, 3 Benecchi, S.; Sheppard, S. S.; Vaduvescu, O.; et al., 2009, MPC 66688, 10 Benner, L., 2010 - private communication Birlan, M., et al, 2010, Astronomy & Astrophysics, 511, A40 Boattini, A., et al., 2003, Earth, Moon and Planets, 93, 239 Boattini, A., et al., 2004, Astronomy & Astrophysics, 418, 743 Bowell, E. and Muinonen, K., 1994, in Hazards due to Comets and Asteroids, 149 Bowell, E., 2011, ASTORB database, ftp://ftp.lowell.edu/pub/elgb/astorb.html Cavadore, C.; Vaduvescu, O.; Birlan, M.; Tudorica, A.; Toma, R.; Sonka, A.; Opriseanu, C.; Vidican, D. et al., 2008 MPC 63369, 9 Chapman, C. R., & Morrison, D., 1994, Nature, 367, 33 EARN, 2011 - Near Earth Asteroids Data-Base, http://earn.dlr.de/nea/ Elst, E. W.; Masi, G.; Lagerkvist, C.-I.; Boattini, A.; Behrend, R.; Vaduvescu, O., 2008, MPC 63591, 10 Elst, E. W.; Lagerkvist, C.-I.; Boattini, A.; Vaduvescu, O.; Greco, C.; Behrend, R., 2008, MPC 63129, 5 Elst, E. W.; Lagerkvist, C.-I.; Boattini, A.; Boehnhardt, H.; Vaduvescu, O., 2008, MPC 62871, 8 Elst, E. W.; Vaduvescu, O.; Birlan, M.; Tudorica, A.; Toma, R.; Sonka, A.; Opriseanu, C.; Vancea, C.; Vidican, D., 2008, MPC 62573, 3 Elst, E. W.; Lagerkvist, C.-I.; Boattini, A.; Vaduvescu, O., 2009, MPC 66195, 4 Elst, E. W.; Lagerkvist, C.-I.; Boattini, A.; Vaduvescu, O., 2009, MPC 65331, 2 Elst, E. W.; Masi, G.; Lagerkvist, C.-I.; Boattini, A.; Behrend, R.; Vaduvescu, O., 2009, MPC 65045, 7 EURONEAR, 2011 - Observing Tools - Planning NEA observations, http://euronear.imcce.fr/tiki-index.php?page=Planning Fitzsimmons, A., Vaduvescu, O. & Asher, D., 2009, MPC 66457, 4 Fitzsimmons, A., Vaduvescu, O., Tudorica, A. & Badea, M., 2009, MPC 65927, 9 Fitzsimmons, A., Vaduvescu, O. & Tudorica, A., 2009, MPC 65332, 1 Grav, T., 2009 - communication on the MPML discussion list Gray, B., 2011a - FIND ORB, http://www.projectpluto.com/find orb.htm Gray, B., 2011b - SAT ID software, http://www.projectpluto.com/sat id.htm Holman, M., Fitzsimmons, A., Grav, T., Vaduvescu, O., 2009, MPC 66196, 2 Holvorcem, P. R., Schwartz, M., Vaduvescu, O., Tudorica, A., Dumitru, D., 2010, MPEC 2010-E39 IMCCE, 2011 - Ephemerides, http://www.imcce.fr/en/ephemerides/ Jedicke, R., 1996, Astronomical Journal, 111, 970 Kern, S. D.; Sheppard, S. S.; Vaduvescu, O.; Schechter, P. L., 2008, MPC 63365, 10 Kern, S. D.; Sheppard, S. S.; Vaduvescu, O.; Schechter, P. L., 2009, MPC 66190, 11 Kolena, Proper Motion to Distance of Asteroid from the Sun, J., 1999 - Conversion http://www.phy.duke.edu/∼kolena/asteroid.html Milani, A. and Gronchi, G. F., 2009, Theory of Orbit Determination Minor Planet Center (MPC), 2011, http://minorplanetcenter.net/ Morbidelli A. et al., 2002, in Asteroids III, 409 Morbidelli A., 2005, Asteroid Population Models, in Dynamics of Population of Planetary Systems, Proc. IAU Colloq. 197, 229 NEODyS, 2011, Near Earth Objects - Dynamic Site, http://newton.dm.unipi.it/neodys/ NASA & JPL, 2011, Near Earth Object Program, http://neo.jpl.nasa.gov/ O'Brien, D. P. & Greenberg, R., 2005, Icarus, 178, 179 Pozo, F.; Barr, A.; Vaduvescu, O. et al., MPEC, 2008-U48 Raab, H., 2011, Astrometrica software, http://www.astrometrica.at Ryan E. R., et al., 2009, The Astronomical Journal, 137, 5134 Smithsonian Astrophysical Observatory, 2011, SAOImage DS9, http://hea-www.harvard.edu/RD/ds9/ Schildknecht, T., 2007, Optical surveys for space debris, Astron Astrophys Rev, 14, 41-111 Scotti, J. V.; Pozo, F.; Barr, A.; Vaduvescu, O. et al., 2008, MPEC, 2008-U45 Scotti, J. V.; Pozo, F.; Barr, A.; Vaduvescu, O. et al., 2008, MPEC, 2008-U46 Sheppard, S. S.; Vaduvescu, O.; Galad, A.; Tudorica, A.; Nedelcu, A.; Toma, R.; Opriseanu, C., 2008, MPC 62258, 1 Skvarc, J, 2010a, Asteroid Residual Calculator http://www.fitsblink.net/residuals/ Skvarc, J, 2010b, Identification of satellites from astrometric positions http://www.fitsblink.net/satellites/ 18 Spaceguard Foundation, 2011, http://spaceguard.esa.int Tholen, D. L., Bernardi, F., & Micheli, M., 2008, AAS DPS Meeting 40, 434 Tholen, D. J.; Vaduvescu, O.; Birlan, M.; Tudorica, A.; Toma, R.; Nedelcu, A.; Sonka, A.; Opriseanu, C. et al., 2008 MPC 62262, 5 Tholen, D. J.; Elst, E. W.; Lagerkvist, C.-I.; Boattini, A.; Behrend, R.; Vaduvescu, O., 2009, MPC 65636, 2 Tubbiolo, A. F.; Vaduvescu, O.; Tudorica, A.; et al., 2008, MPEC, 2008-K66 Vaduvescu, O. & Tudorica, A., 2008, MPC 63125, 6 Vaduvescu, O., 2009, How we discovered between 56 and 483 asteroids in 3 nights, Vega 129, iul 2009 (electronic magazine), Bucharest Astroclub (in Romanian) http://www.astroclubul.ro/index.php/revista-vega/arhiva-vega Wiegert, P. et al, 2007, Astronomical Journal, 133, 1609 Wiegert, P., 2011 - private communication Yoshida, F. et al, 2003, Publ. Astron. Soc. Japan, 55, 701 Yoshida, F. & Nakamura, T., 2007, Planetary and Space Science, 55, 1113 Young, J.; Vaduvescu, O.; Tudorica, A., et al., 2008, MPEC, 2008-K63 19 Figure 1: Pairing the unknown objects observed in multiple nights based on positions derived from the extrapolated arc. Cyan points stand for extrapolated positions of objects observed in the first night and magenta points mark objects observed in the second night in the follow-up field. Figure 2: O–C (observed minus calculated) residuals for program NEAs and known MBAs observed at ESO/MPG. 20 Figure 3: O–C (observed minus calculated) residuals for program NEAs and known MBAs observed with Swope tele- scope. Figure 4: O–C (observed minus calculated) residuals for program NEAs and known MBAs observed with the INT. 21 1.3 2.0 3.5 1.3 2.0 3.5 Figure 5: Basic orbital model using the asteroid observed proper motion µ and the Solar elongation ǫ. We plot all unknown objects observed at ESO/MPG (red), Swope (green) and INT (blue). The three overlaid dotted magenta curves correspond to asteroids orbiting between a = 2.0 and a = 3.5 AU (Main Belt) and a = 1.3 (Near Earth Objects limit). The model allows us to easily flag NEO candidates in a survey. We mark with circles our NEO candidates and we include their properties in Table 4. 22 Figure 6: VBVI213, our best NEO candidate, at the bottom of the 8′ × 16′ ESO/MPG WFI CCD#5, moving in the opposite direction to the four MBAs marked above, and about 10 times faster. The image is displayed in normal sky orientation (N up, E left) and the inset zooms in on the NEO candidate. An animation including all 8 available frames is available online. 23 Figure 7: Orbital distributions of 628 unknown objects observed at ESO/MPG (red points), Swope (green) and INT (blue) compared with the entire known asteroid population (ASTORB - 541,260 fine black points). Although our preliminary orbits were derived using mostly short arcs, the distributions are consistent with the known MBA population, showing the usefulness of the FIND ORB orbital fit in a, e and i. Figure 8: The observed apparent R magnitude (left) and calculated absolute magnitude H (right) versus the semimajor axis a for the ESO/MPG unknown asteroids dataset (red points), Swope (green) and INT (blue). The three objects having the faintest H are among the best NEO candidates. 24 Figure 9: Histograms showing number of unknown objects as function of observed apparent R magnitude (left) and calculated absolute magnitude H (right) for the ESO/MPG dataset (red), Swope (green), INT (blue) and the total number of objects (black dots). Figure 10: Minimal Orbital Intersection Distance (MOID) versus Solar elongation ǫ (left panel), and versus ecliptic latitude β (right panel) for the unknown objects observed at ESO/MPG (red), Swope (green) and INT (blue). The dotted line at MOID < 0.3 marks the NEO region under which all our NEO candidates appear. 25 Table 1: The observing log for NEA observations at INT. We list the name of the asteroid, its classification at time of observation, date of observation, expected apparent magnitude V, exposure time (seconds), number of observed positions, apparent motion µ (′′/min), ephemeris uncertainty (arcsec) and observed orbital arc since discovery (d-days, m-months, y-years). Objects marked with * represent special cases discussed in the paper. Asteroid Class Date (UT) V (mag) Exp (s) Nr pos µ (′′/min) 3σ Obs (′′) arc ? NEA 2009 Feb 12 2009 CW1 2009 Feb 12 2009 CB2* 2009 Feb 12 NEA 2009 CA2 2009 Feb 12 NEA 2003 SJ84* 2009 Apr 15 PHA 2009 FF 2009 Apr 15 PHA 2009 DZ 2009 Apr 15 NEA 2009 FJ30 2009 Apr 15 NEA 2009 FT 2009 Apr 15 PHA 2009 FJ44 2009 Apr 15 PHA 2009 FG19 2009 Apr 15 NEA 2009 FH44 2009 Apr 15 PHA 2009 FY4 2009 Apr 15 NEA 2009 FV4 2009 Apr 15 NEA 2009 FR30 2009 Apr 15 NEA 2009 FT32 2009 Nov 13 VI 2009 VR25* NEA 2009 Nov 14 2009 VN1 NEA 2009 Nov 14 2009 VQ PHA 2009 Nov 14 2009 KK 2010 Feb 20 NEA 2010 DJ1* NEA 2010 Feb 20 2010 CF55 2010 Feb 20 NEA 2010 CF12* 2010 Feb 23 NEA 2010 DX1 2010 Feb 23 2010 DF1 PHA 2007 RM133* NEA 2010 Feb 23 2000 SV20* NEA 2010 Feb 25 NEA 2010 Mar 03 2010 DC* 2010 Mar 03 PHA 2009 FY4 2010 Mar 03 PHA 2007 EF* PHA 2000 CO101 2010 Mar 03 2008 EE NEA 2010 Mar 03 2010 Mar 03 PHA 2006 SS134 PHA 2010 DJ56 2010 Mar 03 NEA 2010 Mar 03 2007 JZ20 2008 TZ3 PHA 2010 Mar 03 3 3 1 2 2 1 5 4 1 2 2 1 1 1 11 5 7 7 2 10 3 10 8 3 1 1 1 12 20 6 6 12 2 1 2 40 1000 30 81 38 3 9 1 2 24 1 12 26 56 42 1400 932 3 270 190 39 ? 24 94 490 2900 58 1 1 1 1 2 3 1 1 9d 6d 9d 6y 1m 1m 18d 18d 16d 25d 16d 26d 26d 20d 24d 2d 5d 6d 6m 3d 5d 8d 4d 6d 3y 10y 17d 1y 3y 10y 2y 4y 11d 3y 2y 20 60 30 120 60 120 30 30 60 90 60 120 90 120 30 30 30 30 60 30 90 90 30 60 120 120 120 10 10 20 20 20 180 180 120 10 5 8 5 5 6 5 6 5 5 5 10 5 4 5 4 5 5 5 5 3 2 5 4 5 4 8 5 8 6 7 3 4 5 7 19.9 20.1 20.0 21.4 20.8 21.1 17.8 19.2 19.5 20.5 19.8 21.0 21.1 21.2 20.8 20.7 21.0 20.9 21.2 19.3 20.8 19.8 17.9 20.5 21.4 21.0 20.1 16.8 18.2 16.3 16.5 20.5 20.9 20.8 21.1 26 Table 2: Summary of the ESO/MPG, Swope and the INT runs. Observations ESO/MPG Swope INT Total Program NEAs Nr of positions O–C standard deviation (′′) Known MBAs Nr of positions O–C standard deviation (′′) Unknown objects Nr of positions Official discoveries Later identifications Outstanding objects Nr. of nights Nr. of observed fields Sky coverage (square degrees) Limiting magnitude (R) Total nr. of objects Total nr. of positions Known MBAs density (obj/sq.deg) Unknown MBAs density Unknown NEOs density 15 156 0.15 347 2,976 0.15 467 4,183 58 17 392 3 42 13 21.5 829 7,315 27 36 0.2-0.6 50 506 0.39 68 680 0.18 41 261 0 8 33 5 50 6 20.4 159 1,447 11 7 0 35 169 0.42 143 733 0.66 120 574 0 22 98 3 35 10 21.2 298 1,476 14 12 0.1-0.8 100 831 - 558 4,389 - 628 5,019 58 47 523 11 127 29 - 1,286 10,239 - - - 27 Table 3: Minor Planet Circulars (MPC) and Minor Planet Electronic Circulars (MPEC) publishing our NEA observations Telescope MP(E)C Reference Swope Swope Swope Swope Swope Swope Swope Swope Swope Swope Swope Swope MPC 64484, 5 MPC 64096, 9 MPC 66688, 10 MPC 64753, 3 MPC 63365, 10 MPC 66190, 11 MPEC 2008-U48 MPEC 2008-U46 MPEC 2008-U45 MPEC 2008-K66 MPC 63125, 6 MPEC 2008-K63 Benecchi et al (2008a) Benecchi et al (2008b) Benecchi et al (2009a) Benecchi et al (2009b) Kern et al (2008) Kern et al (2009) Pozo et al (2008) Scotti et al (2008a) Scotti et al (2008b) Tubbiolo et al (2008) Vaduvescu & Tudorica (2008) Young et al (2009) ESO/MPG MPC 63369, 9 ESO/MPG MPC 63591, 10 ESO/MPG MPC 63129, 5 ESO/MPG MPC 62871, 8 ESO/MPG MPC 62573, 3 ESO/MPG MPC 66195, 4 ESO/MPG MPC 65331, 2 ESO/MPG MPC 65045, 7 ESO/MPG MPC 62258, 1 ESO/MPG MPC 62262, 5 ESO/MPG MPC 65636, 2 Cavadore et al (2009) Elst et al (2008a) Elst et al (2008b) Elst et al (2008c) Elst et al (2008d) Elst et al (2009a) Elst et al (2009b) Elst et al (2009c) Sheppard et al (2008) Tholen et al (2008) Tholen et al (2009) INT INT INT INT INT MPC 66196, 2 MPEC 2010-E39 MPC 66457, 4 MPC 65927, 9 MPC 65332, 1 Holman et al (2009) Holvorcem et al (2010) Fitzsimmons et al. (2009a) Fitzsimmons et al. (2009b) Fitzsimmons et al. (2009c) Table 4: Near Earth Object (NEO) candidates – fast unknown objects observed at ESO/MPG (first group) and the INT (second group). We list in bold 4 objects which qualify as the best NEO candidates, having small MOIDs, 100% NEO Rating score and fitting best the ǫ − µ model. The two one-nighter objects marked with * are not NEAs according to their later identifications. Acronym VB011 VBTU203 VBTU197 VBVA002 VBTU222 VBTU226 VBSO072 VBVI213 VITB01 VTD069 VIT005 * VTD010 VTU021 VTD003 * VTU031 VTU028 VBA003 VTU020 µ (′′/min) 0.55 0.65 0.68 0.60 2.12 0.89 1.17 7.17 0.43 0.60 0.95 1.02 4.61 1.15 0.90 0.72 0.45 0.67 ǫ (◦) 125 132 135 154 162 162 184 191 92 151 167 177 173 193 202 202 219 228 R MOID (AU) a (AU) 20.8 20.8 21.4 19.8 20.4 21.0 20.0 ∼20 20.1 20.5 19.0 20.2 18.9 19.3 20.5 20.6 21.1 20.3 0.19 0.75 0.09 0.57 0.03 0.56 0.70 0.07 2.25 0.51 0.38 0.97 0.02 0.35 0.92 0.30 0.79 0.58 2.28 1.84 3.02 2.50 1.01 1.94 1.95 2.46 3.28 3.01 2.66 2.63 1.03 2.43 1.99 2.39 1.80 1.84 28 e 0.49 0.05 0.68 0.39 0.08 0.20 0.13 0.57 0.01 0.52 0.20 0.26 0.02 0.14 0.03 0.46 0.02 0.14 i (◦) 6.7 23.5 8.2 13.8 2.2 25.7 17.5 2.0 21.1 10.2 14.0 21.3 2.1 3.3 19.6 7.2 23.5 22.0 Nr pos 8 16 7 3 5 5 8 7 5 4 4 5 3 4 5 3 5 6 σ MPC Model (′′) score 0.13 0.26 0.38 0.08 0.16 0.52 0.12 0.31 0.26 0.51 0.38 0.30 0.43 0.35 0.14 0.26 0.08 0.31 14 10 100 22 100 21 24 100 77 6 28 5 100 100 10 15 18 25 close fit best fit best fit fit best bad fit fit fit best fit fit fit fit fit
1012.1780
1
1012
2010-12-08T15:07:07
A new view on planet formation
[ "astro-ph.EP" ]
The standard picture of planet formation posits that giant gas planets are over-grown rocky planets massive enough to attract enormous gas atmospheres. It has been shown recently that the opposite point of view is physically plausible: the rocky terrestrial planets are former giant planet embryos dried of their gas "to the bone" by the influences of the parent star. Here we provide a brief overview of this "Tidal Downsizing" hypothesis in the context of the Solar System structure.
astro-ph.EP
astro-ph
**FULL TITLE** ASP Conference Series, Vol. **VOLUME**, **YEAR OF PUBLICATION** **NAMES OF EDITORS** A new view on planet formation Sergei Nayakshin Department of Physics & Astronomy, University of Leicester, Leicester, LE1 7RH, UK The standard picture of planet formation posits that giant gas Abstract. planets are over-grown rocky planets massive enough to attract enormous gas atmospheres. It has been shown recently that the opposite point of view is physically plausible: the rocky terrestrial planets are former giant planet em- bryos dried of their gas "to the bone" by the influences of the parent star. Here we provide a brief overview of this "Tidal Downsizing" hypothesis in the context of the Solar System structure. 1. Introduction In the popular "core accretion" scenario (CA model hereafter; e.g., Safronov 1969; Wetherill 1990; Pollack et al. 1996), the terrestrial planet cores form first from much smaller solid constituents. A massive gas atmosphere builds up around the rocky core if it reaches a critical mass of about 10 M⊕ (e.g., Mizuno 1980). The CA model's main theoretical difficulty is in the very beginning of the growth: it is not clear how metre-sized rocks would stick together while colliding at high speeds, subject to high radial drifts into the parent star (Weidenschilling 1977, 1980), although gas-dust dynamical instabilities are suggested to help (e.g., Youdin & Goodman 2005; Johansen et al. 2007). Nevertheless, believed to be the only viable model for terrestrial planet formation, the model has enjoyed an almost universal support (e.g., Ida & Lin 2008). This strongest asset of the theory – a "monopoly" on making terrestrial planets – is actually void. Recently, it has been proposed by Boley et al. (2010); Nayakshin (2010a,b,c) that a modified version of the gravitational disc insta- bility model for giant planet formation(Kuiper 1951; Boss 1998) may account for terrestrial planets as well, if gas clump migration (Goldreich & Tremaine 1980) and clump disruption due to tidal forces (McCrea & Williams 1965) are taken into account. This new scheme addresses (Nayakshin 2010c) all of the well known objections (Wetherill 1990; Rafikov 2005) to forming Jupiter in the Solar System via disc fragmentation. The TD hypothesis is a new combination of earlier ideas and contains four important stages (Figure 1): (1) Formation of gas clumps (which we also call giant planet embryos; GEs). As the protoplanetary disc cannot fragment inside R ∼ 50 AU (Rafikov 2005; Boley et al. 2006), GEs are formed at somewhat larger radii. The mass of the clumps is estimated at MGE ∼ 10MJ (10 Jupiter masses) (Boley et al. 2010; Nayakshin 2010a); they are intially fluffy and cool 1 2 (T ∼ 100 K), but contract with time and become much hotter (Nayakshin 2010a). (2) Inward radial migration of the clumps due to gravitational interactions with the surrounding gas disc (Goldreich & Tremaine 1980; Vorobyov & Basu 2010; Boley et al. 2010; Cha & Nayakshin 2010). (3) Grain growth and sedimentation inside the clumps (McCrea & Williams 1965; Boss 1998; Boss et al. 2002). If the clump temperature remains below 1400 − 2000K, massive terrestrial planet cores may form (Nayakshin 2010b), with masses up to the total high Z element content of the clump (e.g., ∼ 60 Earth masses for a Solar metalicity clump of 10MJ ). (4) A disruption of GEs in the inner few AU due to tidal forces (McCrea 1960; McCrea & Williams 1965; Boley et al. 2010; Nayakshin 2010c) or due to irradiation from the star (Nayakshin 2010c) can result in (a) a smallish solid core and a complete gas envelope removal – a terrestrial planet; (b) a massive solid core, with most of the gas removed – a Uranus-like planet; (c) a partial envelope removal leaves a gas giant planet like Jupiter or Saturn. For (b), an internal energy release due to a massive core formation removes the envelope (Handbury & Williams 1975; Nayakshin 2010b). It is interesting to note that it is the proper placement of step (1) into the outer reaches of the System and then the introduction of the radial migration (step 2) that makes this model physically viable. The theory based on elements (3,4) from an earlier 1960-ies scenario for terrestrial planet formation by McCrea (1960); McCrea & Williams (1965) were rejected by Donnison & Williams (1975) because step (1) is not possible in the inner Solar System. Similarly, the giant disc instability (Kuiper 1951; Boss 1998) cannot operate at R ∼ 5 AU to make Jupiter (Rafikov 2005). It is therefore the proper placement of step (1) into the outer reaches of the System and then the introduction of the radial migration (step 2) that makes this model physically viable. The new hypothesis resolves (Nayakshin 2010d) an old mystery of the Solar System: the mainly coherent and prograde rotation of planets, which is unexpected if planets are built by randomly oriented impacts. 2. Solar System structure The gross structure of the Solar System planets is naturally accounted for by the TD model. The innermost terrestrial planets are located within the tidal disruption radius of rt ∼ 2 − 3 AU (Nayakshin 2010c), so these are indeed expected to have no massive atmospheres. The asteroid belt in this scheme are the solids that grew inside the giant planet embryos but not made into the central core, and which were then left around the rt. The gas giant planets are somewhat outside the tidal disruption radius, and thus have been only partially affected by tidal disruption/Solar irradiation. The outer icy giant planets are too far from the Sun to have been affected strongly by it, so they are interesting cases of self-disruption in the TD model. In particular, 35 years ago, Handbury & Williams (1975) suggested that the mas- sive core formation in Uranus and Neptune evaporated most of their hydrogen 3 envelopes. To appreciate the argument, compare the binding energy of the solid core with that of the GE. We expect the core of high-Z elements to have a density ρc ∼ a few g cm−3. The radial size of the solid core, Rcore ∼ (3Mcore/4πρc)1/3. The binding energy of the solid core is Ebind,c ∼ 3 5 GM 2 core Rcore ≈ 1041 erg (cid:18) Mc 10 M⊕(cid:19)5/3 . (1) The clump radius RGE ≈ 0.8 AU at the age of t = 104 years, independently of its massNayakshin (2010c), MGE. Thus, the GE binding energy at that age is Ebind,GE ∼ 3 10 GM 2 GE RGE 3MJ (cid:19)2 ≈ 1041 erg (cid:18) MGE . (2) The two are comparable for Mcore ∼ 10 M⊕. Radiation hydrodynamics sim- ulations confirm such internal disruption events: the run labelled M0α3 in Nayakshin (2010b) made a ∼ 20 M⊕ solid core that unbound all but 0.03 M⊕ of the gaseous material of the original 10MJ gas clump. Future work on the TD hypothesis should address the outer Solar Sys- tem structure (Kuiper belt; comet compositions, etc.). Detailed predictions for exo-planet observations are difficult as the model dependencies are non-linear (Nayakshin 2010b), but some predictions distinctively different from the CA scenario may be possible as planets loose rather than gain mass as they migrate inwards. Acknowledgments. The author acknowledges the support of the STFC research council and the IAU travel grant to attend this exciting meeting. References Boley, A. C., Mej´ıa, A. C., Durisen, R. H., Cai, K., Pickett, M. K., & D'Alessio, P. 2006, ApJ, 651, 517 Boley A. C., Hayfield T., Mayer L., Durisen R. H., 2010, Icarus, 207, 509 Boss A. P., 1998, ApJ, 503, 923 Boss A. P., Wetherill G. W., Haghighipour N., 2002, Icarus, 156, 291 Cha S.-H., Nayakshin S., 2010, submitted to MNRAS Donnison, J. R., & Williams, I. P. 1975, MNRAS, 172, 257 Goldreich P., Tremaine S., 1980, ApJ, 241, 425 Handbury M. J., Williams I. P., 1975, AP&SS, 38, 29 Ida S., Lin D. N. C., 2008, ApJ, 685, 584 Johansen A., Oishi J. S., Low M., Klahr H., Henning T., Youdin A., 2007, Nat, 448, 1022 Kuiper G. P., 1951, in 50th Anniversary of the Yerkes Observatory and Half a Century of Progress in Astrophysics, edited by J. A. Hynek, 357–+ McCrea W. H., 1960, Royal Society of London Proceedings Series A, 256, 245 McCrea W. H., Williams I. P., 1965, Royal Society of London Proceedings Series A, 287, 143 Mizuno H., 1980, Progress of Theoretical Physics, 64, 544 Nayakshin, S. 2010, MNRAS, 408, 2381 Nayakshin S., 2010b, ArXiv e-prints, 1007.4165 Nayakshin, S. 2010, MNRAS, 408, L36 Nayakshin, S. 2010, MNRAS, L167 4 Figure 1. A cartoon of the Tidal Downsizing hypothesis. A protostar (the central Sun symbol) is surrounded by a massive R > ∼ 100 gas disc (the larger grey oval). The four planet formation stages are schematically marked by numbers: (1) The formation of massive gas clumps (embryos) in the outer disc; (2) migration of the clumps closer in to the star, occurring simultaneously with (3) dust grains growth and (possibly) sedimentation into a massive solid core in the centre. The core is shown as a small brown sphere inside the larger gas embryo; (4) disruption of the embryo by tidal forces, irradiation or internal heat liberation. The brown pattern-filled donut-shaped area shows the solid debris ring left from an embryo disruption. The most inward orbit in the diagram shows a terrestrial-like planet, e.g., a solitary solid core whose gas envelope was completely removed. The planet on the next smallest orbit is a giant-like planet with a solid core that retained some of its gas envelope. Pollack J. B., Hubickyj O., Bodenheimer P., Lissauer J. J., Podolak M., Greenzweig Y., 1996, Icarus, 124, 62 Rafikov R. R., 2005, ApJ, 621, L69 Safronov V. S., 1969, Evoliutsiia doplanetnogo oblaka. Shakura N. I., Sunyaev R. A., 1973, A&A, 24, 337 Vorobyov E. I., Basu S., 2005, ApJ, 633, L137 Vorobyov E. I., Basu S., 2006, ApJ, 650, 956 Vorobyov E. I., Basu S., 2010, ArXiv e-prints Weidenschilling S. J., 1977, MNRAS, 180, 57 Weidenschilling S. J., 1980, Icarus, 44, 172 Wetherill G. W., 1990, Annual Review of Earth and Planetary Sciences, 18, 205 Youdin A. N., Goodman J., 2005, ApJ, 620, 459
1310.5630
1
1310
2013-10-21T16:28:50
Transiting hot Jupiters from WASP-South, Euler and TRAPPIST: WASP-95b to WASP-101b
[ "astro-ph.EP" ]
We report the discovery of the transiting exoplanets WASP-95b, WASP-96b, WASP-97b, WASP-98b, WASP-99b, WASP-100b and WASP-101b. All are hot Jupiters with orbital periods in the range 2.1 to 5.7 d, masses of 0.5 to 2.8 Mjup, and radii of 1.1 to 1.4 Rjup. The orbits of all the planets are compatible with zero eccentricity. WASP-99b shows the shallowest transit yet found by WASP-South, at 0.4%. The host stars are of spectral type F2 to G8. Five have metallicities of [Fe/H] from -0.03 to +0.23, while WASP-98 has a metallicity of -0.60, exceptionally low for a star with a transiting exoplanet. Five of the host stars are brighter than V = 10.8, which significantly extends the number of bright transiting systems available for follow-up studies. WASP-95 shows a possible rotational modulation at a period of 20.7 d. We discuss the completeness of WASP survey techniques by comparing to the HAT project.
astro-ph.EP
astro-ph
Mon. Not. R. Astron. Soc. 000, range (0000) Printed 27 June 2018 (MN LATEX style file v2.2) Transiting hot Jupiters from WASP-South, Euler and TRAPPIST: WASP-95b to WASP-101b Coel Hellier1, D.R. Anderson1, A. Collier Cameron2, L. Delrez3, M. Gillon3, E. Jehin3, M. Lendl4, P.F.L. Maxted1, F. Pepe4, D. Pollacco5, D. Queloz4,6, D. S´egransan4, B. Smalley1, A.M.S. Smith1,7, J. Southworth1, A.H.M.J. Triaud4,8⋆, S. Udry4 & R.G. West5 1Astrophysics Group, Keele University, Staffordshire, ST5 5BG, UK 2SUPA, School of Physics and Astronomy, University of St. Andrews, North Haugh, Fife, KY16 9SS, UK 3Institut d'Astrophysique et de G´eophysique, Universit´e de Li`ege, All´ee du 6 Aout, 17, Bat. B5C, Li`ege 1, Belgium 4Observatoire astronomique de l'Universit´e de Gen`eve 51 ch. des Maillettes, 1290 Sauverny, Switzerland 5Department of Physics, University of Warwick, Gibbet Hill Road, Coventry CV4 7AL, UK 6Cavendish Laboratory, J J Thomson Avenue, Cambridge, CB3 0HE, UK 7N. Copernicus Astronomical Centre, Polish Academy of Sciences, Bartycka 18, 00-716 Warsaw, Poland 8Department of Physics and Kavli Institute for Astrophysics & Space Research, Massachusetts Institute of Technology, Cambridge, MA 02139, USA date ABSTRACT We report the discovery of the transiting exoplanets WASP-95b, WASP-96b, WASP- 97b, WASP-98b, WASP-99b, WASP-100b and WASP-101b. All are hot Jupiters with orbital periods in the range 2.1 to 5.7 d, masses of 0.5 to 2.8 MJup, and radii of 1.1 to 1.4 RJup. The orbits of all the planets are compatible with zero eccentricity. WASP-99b shows the shallowest transit yet found by WASP-South, at 0.4%. The host stars are of spectral type F2 to G8. Five have metallicities of [Fe/H] from –0.03 to +0.23, while WASP-98 has a metallicity of –0.60, exceptionally low for a star with a transiting exoplanet. Five of the host stars are brighter than V = 10.8, which significantly extends the number of bright transiting systems available for follow-up studies. WASP-95 shows a possible rotational modulation at a period of 20.7 d. We discuss the completeness of WASP survey techniques by comparing to the HAT project. Key words: planetary systems 1 INTRODUCTION The WASP-South survey has dominated the discovery of transiting hot-Jupiter exoplanets in the Southern hemi- sphere. WASP-South is well matched to the capabilities of the Euler/CORALIE spectrograph and the robotic TRAP- PIST telescope, with the combination of all three proving efficient for discovering transiting exoplanets in the range V = 9–13. WASP-South has now been running nearly continuously for 7 years. Approximately 1000 candidates have been ob- served with Euler/CORALIE, while, since December 2010, TRAPPIST has observed 1000 lightcurves of WASP candi- ⋆ Fellow of the Swiss National Science Foundation c(cid:13) 0000 RAS dates and planets. Here we present new WASP-South plan- ets which take WASP numbering above 100. Since WASP host stars are generally brighter than host stars of Kepler exoplanets, ongoing WASP-South discover- ies are important for detailed study of exoplanets and will be prime targets for future missions such as CHEOPS and JWST, and proposed missions such as EChO and FINESSE. 2 OBSERVATIONS The observational and analysis techniques used here are the same as in recent WASP discovery papers (e.g. Hellier et al. 2012), and thus are described briefly. For detailed accounts see the early papers including Pollacco et al. (2006), Collier- Cameron et al. (2007a) and Pollacco et al. (2007). In outline, WASP-South surveys the visible sky each 2 Hellier et al. Table 1. Observations Facility Date WASP-95: WASP-South Euler/CORALIE TRAPPIST EulerCAM TRAPPIST EulerCAM EulerCAM WASP-96: WASP-South Euler/CORALIE TRAPPIST EulerCAM TRAPPIST EulerCAM WASP-97: WASP-South Euler/CORALIE EulerCAM TRAPPIST EulerCAM EulerCAM TRAPPIST WASP-98: WASP-South Euler/CORALIE EulerCAM TRAPPIST EulerCAM TRAPPIST TRAPPIST EulerCAM WASP-99: WASP-South Euler/CORALIE EulerCAM WASP-100: WASP-South Euler/CORALIE TRAPPIST TRAPPIST TRAPPIST WASP-101: WASP-South Euler/CORALIE TRAPPIST TRAPPIST TRAPPIST TRAPPIST EulerCAM TRAPPIST 2010 May–2011 Nov 2012 Jul–2013 May 2012 Sep 14 2012 Sep 16 2012 Sep 27 2013 May 23 2013 Aug 01 2010 Jun–2011 Dec 2011 Oct–2012 Oct 2011 Nov 11 2012 Jul 01 2012 Dec 11 2013 Jun 29 2010 Jun–2012 Jan 2012 Sep–2012 Nov 2012 Nov 23 2012 Nov 23 2013 Jul 11 2013 Aug 09 2013 Aug 09 2006 Aug–2012 Jan 2011 Nov–2012 Nov 2012 Oct 28 2012 Oct 28 2012 Oct 31 2012 Nov 03 2013 Aug 29 2013 Sep 07 23 200 points 14 radial velocities z band Gunn r filter z band Gunn r filter Gunn r filter 13 100 points 21 radial velocities Blue-block filter Gunn r filter Blue-block filter Gunn r filter 23 900 points 12 radial velocities Gunn r filter z-band Gunn r filter Gunn r filter I + z filter 12 700 points 15 radial velocities IC band blue-block filter IC band blue-block filter blue-block filter IC band 2010 Jul–2012 Jan 2012 Feb–2013 Jan 2012 Nov 21 9000 points 20 radial velocities Gunn r filter 2010 Aug–2012 Jan 2012 Sep–2013 Mar 2012 Dec 04 2013 Jan 10 2013 Sep 29 13 500 points 19 radial velocities I + z filter I + z filter z filter 2009 Jan–2012 Mar 2011 Jan–2013 Jun 2012 Jan 11 2012 Jan 22 2012 Jan 29 2012 Mar 05 2012 Dec 06 2013 Feb 23 16 100 points 21 radial velocities z band z band z band z band Gunn r filter z band clear night using an array of 200mm f/1.8 lenses and a cadence of ∼ 10 mins. Transit searching of accumulated lightcurves leads to candidates that are passed to TRAP- PIST (Jehin et al. 2011), a robotic 0.6-m photometric tele- scope, which can resolve blends and check that the candidate transits are planet-like, and to the 1.2-m Euler/CORALIE spectrograph, for radial-velocity observations. About 1 in 12 candidates turns out to be a planet. Higher-quality transit lightcurves are then obtained with TRAPPIST and Euler- CAM (Lendl et al. 2012). A list of the observations reported here is given in Table 1 while the CORALIE radial velocities are listed in Table A1. 3 THE HOST STARS For each star, the individual CORALIE spectra were co- added to produce a single spectrum with typical S/N of ∼ 100:1 (though the fainter WASP-98, V = 13, had a S/N of only 40:1). Our methods for spectral analysis are described in Doyle et al. (2013). The excitation balance of the Fe i lines was used to determine the effective temperature (Teff ), and spectral type was estimated from Teff using the table in Gray (2008). The surface gravity (log g) was determined from the ionisation balance of Fe i and Fe ii. The Ca i line at 6439A and the Na i D lines were also used as log g di- agnostics. The metallicity was determined from equivalent width measurements of several unblended lines. The projected stellar rotation velocity (v sin i) was de- termined by fitting the profiles of several unblended Fe i lines. An instrumental FWHM of 0.11 ± 0.01 A was deter- mined from the telluric lines around 6300A. For age estimates we use the lithium abundance and the Sestito & Randlich (2005) calibration. We also give a gyrochronological age, from the measured v sin i and assum- ing that the star's spin is perpendicular to us, so that this is the true equatorial speed. This is then combined with the stellar radius to give a rotational period, to compare with the results of Barnes (2007). The results for each star are listed in the Tables 2 to 10. We also list proper motions from the UCAC4 catalogue of Zacharias et al. (2013). The motions and metallicities are all compatible with the stars being local thin-disc stars. We searched the WASP photometry of each star for rotational modulations by using a sine-wave fitting algo- rithm as described by Maxted et al. (2011). We estimated the significance of periodicities by subtracting the fitted tran- sit lightcurve and then repeatedly and randomly permuting the nights of observation. We found a marginally significant modulation in WASP-95 (see Section 4.2) but not in any of the other stars (with 95%-confidence upper limits being typically 1 mmag). 4 SYSTEM PARAMETERS The CORALIE radial-velocity measurements were com- bined with the WASP, EulerCAM and TRAPPIST photom- etry in a simultaneous Markov-chain Monte-Carlo (MCMC) analysis to find the system parameters. For details of our methods see Collier Cameron et al. (2007b). The limb- darkening parameters are noted in each Table, and are taken from the 4-parameter non-linear law of Claret (2000). For all of our planets the data are compatible with zero eccentricity and hence we imposed a circular orbit (see An- derson et al. 2012 for the rationale for this). The upper limits on the eccentricity range from 0.02 (for the V = 9.5 WASP- 99) to 0.27 (for the fainter, V = 12.2, WASP-96). c(cid:13) 0000 RAS, MNRAS 000, range Table 2. System parameters for WASP-95. 1SWASP J222949.73–480011.0 2MASS 22294972–4800111 RA = 22h29m49.73s, Dec = –48◦00 V mag = 10.1 Rotational modulation 2 mmag at ∼ 20.7 d pm (RA) 94.1 ± 0.9 (Dec) –8.5 ± 1.0 mas/yr 11.0 ′′ ′ (J2000) Stellar parameters from spectroscopic analysis. Spectral type Teff (K) log g v sin I (km s−1) [Fe/H] log A(Li) Age (Lithium) [Gy] Age (Gyro) [Gy] G2 5830 ± 140 4.36 ± 0.07 3.1 ± 0.6 +0.14 ± 0.16 <0.2 >3 2.4+1.7 −1.0 Parameters from MCMC analysis. ∗ P (d) Tc (HJD) (UTC) T14 (d) T12 = T34 (d) ∆F = R2 P/R2 b i (◦) K1 (km s−1) γ (km s−1) e M∗ (M⊙) R∗ (R⊙) log g∗ (cgs) ρ∗ (ρ⊙) Teff (K) MP (MJup) RP (RJup) log gP (cgs) ρP (ρJ) a (AU) TP,A=0 (K) 2.1846730 ± 0.0000014 245 6338.45851 ± 0.00024 0.116 ± 0.001 0.011 ± 0.001 0.0105 ± 0.0003 0.19 +0.21 −0.13 88.4 +1.2 −2.1 0.1757 ± 0.0017 6.2845 ± 0.0004 0 (adopted) (< 0.04 at 3σ) 1.11 ± 0.09 1.13 +0.08 −0.04 4.38 +0.02 −0.04 0.78 +0.04 −0.13 5630 ± 130 1.13 +0.1 −0.04 1.21 ± 0.06 3.34 +0.02 −0.07 0.85 +0..07 −0.2 0.03416 ± 0.00083 1570 ± 50 Errors are 1σ; Limb-darkening coefficients were: (Euler r) a1 = 0.701, a2 = –0.490, a3 = 1.077, a4 = –0.516 (Trap z) a1 = 0.780, a2 = –0.718, a3 = 1.082, a4 = –0.487 WASP-South hot Jupiters 3 0.6 0.8 1 1.2 1.4 WASP TRAPPIST EulerCAM TRAPPIST EulerCAM EulerCAM 0.96 0.98 1 1.02 1.04 CORALIE g a m . l e R -0.06 0 0.06 1 0.99 0.98 0.97 x u l f e v i t l a e R 0.96 0.95 0.94 0.93 0.92 0.91 200 100 0 -100 -200 60 0 -60 ) 1 - s m ( V R . l e R n a p s . s B i 0.6 0.8 1 1.2 1.4 Orbital phase Figure 1. WASP-95b discovery data: (Top) The WASP data folded on the transit period. (Second panel) The binned WASP data with (offset) the follow-up transit lightcurves (ordered from the top as in Table 1) together with the fitted MCMC model. (Third) The CORALIE radial velocities with the fitted model. (Lowest) The bisector spans; the absence of any correlation with radial velocity is a check against transit mimics. c(cid:13) 0000 RAS, MNRAS 000, range 4 Hellier et al. Table 3. System parameters for WASP-96. 1SWASP J000411.14–472138.2 2MASS 00041112–4721382 RA = 00h04m11.14s, Dec = –47◦21 V mag = 12.2 Rotational modulation < 1 mmag (95%) pm (RA) 23.1 ± 1.0 (Dec) 3.7 ± 1.0 mas/yr 38.2 ′′ ′ (J2000) Stellar parameters from spectroscopic analysis. Spectral type Teff (K) log g v sin I (km s−1) [Fe/H] log A(Li) Age (Lithium) [Gy] Age (Gyro) [Gy] G8 5500 ± 150 4.25 ± 0.15 1.5 ± 1.3 +0.14 ± 0.19 1.48 ± 0.15 2 ∼ 5 8+26 −8 Parameters from MCMC analysis. ∗ P (d) Tc (HJD) (UTC) T14 (d) T12 = T34 (d) ∆F = R2 P/R2 b i (◦) K1 (km s−1) γ (km s−1) e M∗ (M⊙) R∗ (R⊙) log g∗ (cgs) ρ∗ (ρ⊙) Teff (K) MP (MJup) RP (RJup) log gP (cgs) ρP (ρJ) a (AU) TP,A=0 (K) 3.4252602 ± 0.0000027 245 6258.0621 ± 0.0002 0.1011 ± 0.0011 0.0200 ± 0.0014 0.0138 ± 0.0003 0.710 ± 0.019 85.6 ± 0.2 0.062 ± 0.004 –1.23300 ± 0.00009 0 (adopted) (< 0.27 at 3σ) 1.06 ± 0.09 1.05 ± 0.05 4.42 ± 0.02 0.922 ± 0.073 5540 ± 140 0.48 ± 0.03 1.20 ± 0.06 2.88 ± 0.04 0.28 ± 0.04 0.0453 ± 0.0013 1285 ± 40 Errors are 1σ; Limb-darkening coefficients were: (Euler r) a1 = 0.722, a2 = –0.581, a3 = 1.203, a4 = –0.561 (Trap BB) a1 = 0.722, a2 = –0.581, a3 = 1.203, a4 = –0.561 g a m -0.06 0 0.06 . l e R 1 0.99 0.98 x u 0.97 l f e v i t l a e R 0.96 0.95 0.94 0.93 100 0 ) 1 - s m ( V R . l e R -100 100 0 -100 -200 n a p s . s B i 0.6 0.8 1 1.2 1.4 WASP TRAPPIST EulerCAM TRAPPIST EulerCAM 0.98 1 1.02 CORALIE 0.6 0.8 1 1.2 1.4 Orbital phase Figure 2. WASP-96b discovery data (as in Fig. 1). c(cid:13) 0000 RAS, MNRAS 000, range Table 4. System parameters for WASP-97. 1SWASP J013825.04–554619.4 2MASS 01382504–5546194 RA = 01h38m25.04s, Dec = –55◦46 V mag = 10.6 Rotational modulation < 1 mmag (95%) pm (RA) 94.6 ± 1.1 (Dec) 20.8 ± 1.0 mas/yr 19.4 ′′ ′ (J2000) Stellar parameters from spectroscopic analysis. Spectral type Teff (K) log g v sin I (km s−1) [Fe/H] log A(Li) Age (Lithium) [Gy] Age (Gyro) [Gy] G5 5670 ± 110 4.45 ± 0.08 1.1 ± 0.5 +0.23 ± 0.11 < 0.85 > 5 11.9+16.0 −8.3 Parameters from MCMC analysis. ∗ P (d) Tc (HJD) (UTC) T14 (d) T12 = T34 (d) ∆F = R2 P/R2 b i (◦) K1 (km s−1) γ (km s−1) e M∗ (M⊙) R∗ (R⊙) log g∗ (cgs) ρ∗ (ρ⊙) Teff (K) MP (MJup) RP (RJup) log gP (cgs) ρP (ρJ) a (AU) TP,A=0 (K) 2.072760 ± 0.000001 245 6438.18683 ± 0.00018 0.1076 ± 0.0008 0.011 ± 0.001 0.0119 ± 0.0002 0.23 +0.11 −0.15 88.0 +1.3 −1.0 0.1945 ± 0.0023 6.80877 ± 0.00029 0 (adopted) (< 0.05 at 3σ) 1.12 ± 0.06 1.06 ± 0.04 4.43 ± 0.03 0.93 ± 0.09 5640 ± 100 1.32 ± 0.05 1.13 ± 0.06 3.37 ± 0.04 0.91 ± 0.11 0.03303 ± 0.00056 1555 ± 40 Errors are 1σ; Limb-darkening coefficients were: (Euler r) a1 = 0.700, a2 = –0.489, a3 = 1.077, a4 = –0.516 (Trap Iz) a1 = 0.778, a2 = –0.713, a3 = 1.077, a4 = –0.484 g a m . l e R -0.06 0 0.06 1 0.99 0.98 0.97 0.96 0.95 x u l f e v i t l a e R 0.94 0.93 0.92 0.91 200 100 0 -100 -200 50 0 -50 ) 1 - s m ( V R . l e R n a p s . s B i -100 WASP-South hot Jupiters 5 0.6 0.8 1 1.2 1.4 WASP EulerCAM TRAPPIST EulerCAM EulerCAM TRAPPIST 0.96 0.98 1 1.02 1.04 CORALIE 0.6 0.8 1 1.2 1.4 Orbital phase Figure 3. WASP-97b discovery data (as in Fig. 1). c(cid:13) 0000 RAS, MNRAS 000, range 6 Hellier et al. Table 5. System parameters for WASP-98. 1SWASP J035342.90–341941.7 2MASS 03534291–3419414 RA = 03h53m42.90s, Dec = –34◦19 V mag = 13.0 Rotational modulation < 2 mmag (95%) pm (RA) 32.0 ± 1.1 (Dec) –13.4 ± 1.1 mas/yr 41.7 ′′ ′ (J2000) Stellar parameters from spectroscopic analysis. Spectral type Teff (K) log g v sin I (km s−1) [Fe/H] log A(Li) Age (Lithium) [Gy] Age (Gyro) [Gy] G7 5550 ± 140 4.40 ± 0.15 < 0.5 −0.60 ± 0.19 < 0.91 > 3 > 8 Parameters from MCMC analysis. ∗ P (d) Tc (HJD) (UTC) T14 (d) T12 = T34 (d) P/R2 ∆F = R2 b i (◦) K1 (km s−1) γ (km s−1) e M∗ (M⊙) R∗ (R⊙) log g∗ (cgs) ρ∗ (ρ⊙) Teff (K) MP (MJup) RP (RJup) log gP (cgs) ρP (ρJ) a (AU) TP,A=0 (K) 2.9626400 ± 0.0000013 245 6333.3913 ± 0.0001 0.0795 ± 0.0005 0.0202 ± 0.0006 0.02570 ± 0.00025 0.71 ± 0.01 86.3 ± 0.1 0.15 ± 0.01 –38.2882 ± 0.0003 0 (adopted) (< 0.24 at 3σ) 0.69 ± 0.06 0.70 ± 0.02 4.583 ± 0.014 1.99 ± 0.07 5525 ± 130 0.83 ± 0.07 1.10 ± 0.04 3.20 ± 0.03 0.63 ± 0.06 0.036 ± 0.001 1180 ± 30 Errors are 1σ; Limb-darkening coefficients were: (Euler I) a1 = 0.558, a2 = –0.157, a3 = 0.576, a4 = –0.325 (Trap BB) a1 = 0.486, a2 = 0.090, a3 = 0.444, a4 = –0.294 0.6 0.8 1 1.2 1.4 WASP EulerCAM TRAPPIST EulerCAM TRAPPIST TRAPPIST EulerCAM 0.98 1 1.02 CORALIE g a m . l e R -0.06 0 0.06 1 0.99 0.98 0.97 0.96 0.95 0.94 x u l f e v i t l a e R 0.93 0.92 0.91 0.9 0.89 0.88 200 100 0 -100 -200 ) 1 - s m ( V R . l e R n a p s . s B i 100 0 -100 -200 0.6 0.8 1 1.2 1.4 Orbital phase Figure 4. WASP-98b discovery data (as in Fig. 1). c(cid:13) 0000 RAS, MNRAS 000, range WASP-South hot Jupiters 7 0.6 0.8 1 1.2 1.4 WASP EulerCAM 0.96 0.98 1 1.02 1.04 CORALIE g a m . l e R -0.06 0 0.06 1 x u l f 0.99 e v i t l a e R 0.98 300 200 100 0 -100 -200 ) 1 - s m ( V R . l e R -300 100 0 n a p s . s B i -100 0.6 0.8 1 1.2 1.4 Orbital phase Figure 5. WASP-99b discovery data (as in Fig. 1). Table 6. System parameters for WASP-99. 1SWASP J023935.44–500028.8 2MASS 02393544–5000288 RA = 02h39m35.44s, Dec = –50◦00 V mag = 9.5 Rotational modulation < 2 mmag (95%) pm (RA) –2.7 ± 1.1 (Dec) –38.1 ± 0.9 mas/yr 28.8 ′′ ′ (J2000) Stellar parameters from spectroscopic analysis. Spectral type Teff (K) log g v sin I (km s−1) [Fe/H] log A(Li) Age (Lithium) [Gy] Age (Gyro) [Gy] F8 6150 ± 100 4.3 ± 0.1 6.8 ± 0.5 +0.21 ± 0.15 2.52 ± 0.08 1 ∼ 3 1.4+1.1 −0.6 Parameters from MCMC analysis. ∗ P (d) Tc (HJD) (UTC) T14 (d) T12 = T34 (d) P/R2 ∆F = R2 b i (◦) K1 (km s−1) γ (km s−1) e M∗ (M⊙) R∗ (R⊙) log g∗ (cgs) ρ∗ (ρ⊙) Teff (K) MP (MJup) RP (RJup) log gP (cgs) ρP (ρJ) a (AU) TP,A=0 (K) 5.75251 ± 0.00004 245 6224.9824 ± 0.0014 0.219 ± 0.003 0.0137 +0.0017 −0.0006 0.0041 ± 0.0002 0.18 ± 0.17 88.8 ± 1.1 0.2422 ± 0.0017 24.9610 ± 0.0002 0 (adopted) (< 0.02 at 3σ) 1.48 ± 0.10 1.76 +0.11 −0.06 4.12 +0.02 −0.04 0.27 +0.02 −0.04 6180 ± 100 2.78 ± 0.13 1.10 +0.08 −0.05 3.72 +0.03 −0.06 2.1 ± 0.3 0.0717 ± 0.0016 1480 ± 40 Errors are 1σ; Limb-darkening coefficients were: (Euler r) a1 = 0.590, a2 = 0.036, a3 = 0.306, a4 = –0.205 c(cid:13) 0000 RAS, MNRAS 000, range 8 Hellier et al. Table 7. System parameters for WASP-100. 1SWASP J043550.32–640137.3 2MASS 04355033–6401373 RA = 04h35m50.32s, Dec = –64◦01 V mag = 10.8 Rotational modulation < 1 mmag (95%) pm (RA) 11.9 ± 1.0 (Dec) –2.2 ± 2.3 mas/yr 37.3 ′′ ′ (J2000) Stellar parameters from spectroscopic analysis. Spectral type Teff (K) log g v sin I (km s−1) [Fe/H] log A(Li) Age (Lithium) [Gy] Age (Gyro) [Gy] F2 6900 ± 120 4.35 ± 0.17 12.8 ± 0.8 −0.03 ± 0.10 < 1.80 too hot too hot Parameters from MCMC analysis. ∗ P (d) Tc (HJD) (UTC) T14 (d) T12 = T34 (d) P/R2 ∆F = R2 b i (◦) K1 (km s−1) γ (km s−1) e M∗ (M⊙) R∗ (R⊙) log g∗ (cgs) ρ∗ (ρ⊙) Teff (K) MP (MJup) RP (RJup) log gP (cgs) ρP (ρJ) a (AU) TP,A=0 (K) 2.849375 ± 0.000008 245 6272.3395 ± 0.0009 0.160 ± 0.005 0.021 ± 0.005 0.0076 ± 0.0005 0.64 +0.08 −0.16 82.6 +2.6 −1.7 0.213 ± 0.008 29.9650 ± 0.0002 0 (adopted) (< 0.10 at 3σ) 1.57 ± 0.10 2.0 ± 0.3 4.04 ± 0.11 0.20 +0.10 −0.05 6900 ± 120 2.03 ± 0.12 1.69 ± 0.29 3.21 ± 0.15 0.4 ± 0.2 0.0457 ± 0.0010 2190 ± 140 Errors are 1σ; Limb-darkening coefficients were: (Trap Iz) a1 = 0.542, a2 = 0.086, a3 = 0.001, a4 = –0.055 g a m . l e R -0.06 0 0.06 1 0.99 0.98 x u l f e v i t 0.97 l a e R 0.96 0.95 0.94 ) 1 - s m ( V R . l e R 200 100 0 -100 -200 n a p s . s B i 200 100 0 -100 -200 0.6 0.8 1 1.2 1.4 WASP TRAPPIST TRAPPIST TRAPPIST 0.94 0.96 0.98 1 1.02 1.04 1.06 CORALIE 0.6 0.8 1 1.2 1.4 Orbital phase Figure 6. WASP-100b discovery data (as in Fig. 1). c(cid:13) 0000 RAS, MNRAS 000, range Table 8. System parameters for WASP-101. 1SWASP J063324.26–232910.2 2MASS 06332426–2329103 RA = 06h33m24.26s, Dec = –23◦29 V mag = 10.3 Rotational modulation < 1 mmag (95%) pm (RA) –2.3 ± 0.9 (Dec) 23.3 ± 1.2 mas/yr 10.2 ′′ ′ (J2000) Stellar parameters from spectroscopic analysis. Spectral type Teff (K) log g v sin I (km s−1) [Fe/H] log A(Li) Age (Lithium) [Gy] Age (Gyro) [Gy] F6 6380 ± 120 4.31 ± 0.08 12.4 ± 0.5 +0.20 ± 0.12 2.80 ± 0.09 0.5 ∼ 2 0.9+1.3 −0.4 Parameters from MCMC analysis. ∗ P (d) Tc (HJD) (UTC) T14 (d) T12 = T34 (d) ∆F = R2 P/R2 b i (◦) K1 (km s−1) γ (km s−1) e M∗ (M⊙) R∗ (R⊙) log g∗ (cgs) ρ∗ (ρ⊙) Teff (K) MP (MJup) RP (RJup) log gP (cgs) ρP (ρJ) a (AU) TP,A=0 (K) 3.585722 ± 0.000004 245 6164.6934 ± 0.0002 0.113 ± 0.001 0.023 ± 0.001 0.0126 ± 0.0002 0.736 ± 0.013 85.0 ± 0.2 0.054 ± 0.004 42.6373 ± 0.0006 0 (adopted) (< 0.03 at 3σ) 1.34 ± 0.07 1.29 ± 0.04 4.345 ± 0.019 0.626 ± 0.043 6400 ± 110 0.50 ± 0.04 1.41 ± 0.05 2.76 ± 0.04 0.18 ± 0.02 0.0506 ± 0.0009 1560 ± 35 Errors are 1σ; Limb-darkening coefficients were: (Trap z) a1 = 0.640, a2 = –0.172, a3 = 0.302, a4 = –0.174 (Euler r) a1 = 0.548, a2 = 0.238, a3 = –0.010, a4 = –0.067 WASP-South hot Jupiters 9 0.6 0.8 1 1.2 1.4 WASP TRAPPIST TRAPPIST TRAPPIST TRAPPIST EulerCAM TRAPPIST 0.98 1 1.02 CORALIE g a m . l e R -0.06 0 0.06 1 0.99 0.98 0.97 0.96 x u 0.95 l f e v i t l a e R 0.94 0.93 0.92 0.91 0.9 0.89 100 0 ) 1 - s m ( V R . l e R -100 100 0 -100 n a p s . s B i 0.6 0.8 1 1.2 1.4 Orbital phase Figure 7. WASP-101b discovery data (as in Fig. 1). c(cid:13) 0000 RAS, MNRAS 000, range 10 Hellier et al. Figure 8. Evolutionary tracks on a modified H–R diagram −1/3 (ρ versus Teff ). The red lines are for solar metallicity, [Fe/H] ∗ = 0, showing (solid lines) mass tracks with the labelled mass, and (dashed lines) age tracks for log(age) = 7.85 & 9.15 yrs. The green lines are the same but for a higher metallicity of [Fe/H] = +0.19 and log(age) = 7.85, 9.4 & 9.8. The blue lines are the same for a lower metallicity of [Fe/H] = –0.6, and log(age) = 7.85 & 9.8 yrs. Stars are colour coded to the nearest of these metallicites. The models are from Girardi et al. (2000). The fitted parameters were Tc, P , ∆F , T14, b, K1, where Tc is the epoch of mid-transit, P is the orbital period, ∆F is the fractional flux-deficit that would be observed during transit in the absence of limb-darkening, T14 is the total transit duration (from first to fourth contact), b is the impact parameter of the planet's path across the stellar disc, and K1 is the stellar reflex velocity semi-amplitude. The transit lightcurves lead directly to stellar density but one additional constraint is required to obtain stellar masses and radii, and hence full parametrisation of the sys- tem. Here we use the calibrations presented by Southworth (2011), based on masses and radii of eclipsing binaries. For each system we list the resulting parameters in Ta- bles 2 to 8, and plot the resulting data and models in Fig- ures 1 to 7. We also refer the reader to Smith et al. (2012) who present an extensive analysis of the effect of red noise in the transit lightcurves on the resulting system parameters. As in past WASP papers we plot the spectroscopic Teff , and the stellar density from fitting the transit, against the evolutionary tracks from Girardi et al. (2000), as shown in Fig. 8. 0.06 0.05 0.04 n P 0.03 0.02 0.01 0.00 e d u t i n g a M −0.04 −0.02 0.00 0.02 0.04 1 10 Period [d] 100 −0.2 0.0 0.2 0.4 0.6 0.8 1.0 1.2 Phase Figure 9. The possible rotational modulation in WASP-95 at a period of 20.7 d. The horizontal lines are the 10% (dashed) and 1% (dot-dashed) false-alarm probabilities. The upper plot is the periodogram (the y-scale being the fraction of the scatter in the data that is modelled by a sinusoidal variation, weighted by the standard errors); the lower plot is the data folded on the 20.7-d period. 4.1 WASP-95 WASP-95 is a V = 10.1, G2 star with an [Fe/H] of +0.14 ± 0.16. It may be slightly evolved, with an age of several billion years. WASP-95 shows a possible rotational modulation at a period of 20.7 d and an amplitude of 2 mmag in the WASP data (Fig. 9), though this is seen in only one of the two years of data. The values of v sin i from the spectroscopic analysis (assuming that the spin axis is perpendicular to us) and the stellar radius from the transit analysis combine to a rotation period of 19.7 ± 3.9 d, which is compatible with the possible rotational modulation. WASP-95b is a typical hot Jupiter (Porb = 2.18 d, M = 1.2 MJup, R = 1.2 RJup). c(cid:13) 0000 RAS, MNRAS 000, range WASP-South hot Jupiters 11 20000 37 1 4.2 WASP-96 WASP-96 is fainter G8 star, at V = 12.2, with an [Fe/H] of +0.14 ± 0.19. The planet is a typical hot Jupiter (Porb = 3.4 d, M = 0.5 MJup, R = 1.2 RJup). 4.3 WASP-97 WASP-97 is a V = 10.6, G5 star, with an above-solar metal- licity of [Fe/H] of +0.23 ± 0.11. The planet is again a typical hot Jupiter (Porb = 2.1 d, M = 1.3 MJup, R = 1.1 RJup). 4.4 WASP-98 At V = 13.0, WASP-98 is at the faint end of the WASP- South/CORALIE survey. The spectral analysis has a lower S/N of only 40 and so is less reliable. It appears to be excep- tionally metal poor ([Fe/H] = –0.6 ± 0.19) for a transit host star. The spectral analysis suggests G7, though the transit parameters lead to a lower mass of 0.7 M⊙. The planet is a typical hot Jupiter having a high-impact (b = 0.7) transit. 10000 5000 s t 15000 i n o p P S A W 19 9 38 15 4.5 WASP-99 WASP-99 is the brightest host star reported here, at V = 9.5. Spectral analysis suggests that it is an F8 star with [Fe/H] of +0.21 ± 0.15, though the transit analysis gives a higher mass (M = 1.5 M⊙) and an expanded radius (R = 1.8 R⊙). The planet is relatively massive (M = 2.8 MJup), though its mass and radius are similar to those of many known planets. The transit of WASP-99b is the shallowest yet found by WASP-South, at 0.0041 ± 0.0002, along with that of WASP-72b at 0.0043 ± 0.0004 (Gillon et al. 2013). 4.6 WASP-100 WASP-100 is a V = 10.8, F2 star of solar metallicity. The planet is a typical bloated hot Jupiter (Porb = 2.8 d, M = 2.0 MJup, R = 1.7 RJup) with a high irradiation (see, e.g., West et al. 2013). 4.7 WASP-101 WASP-101 is a V = 10.3, F6 star with [Fe/H] = +0.20 ± 0.12. The analysis is compatible with an unevolved main- sequence star. The planet is a bloated, low-mass planet (Porb = 3.6 d, M = 0.50 MJup, R = 1.4 RJup). 5 COMPLETENESS With the WASP survey now passing WASP-100 we can ask how complete the survey methods are and how many planets we are missing. A full discussion of selection effects is a large topic and is not attempted here, but we can use the HAT project as a straightforward check on our techniques. The HAT project (Bakos et al. 2004) is very similar in conception and hardware to WASP, and thus whether a HAT planet is also detected by us gives an indication of how many we are overlooking. c(cid:13) 0000 RAS, MNRAS 000, range 0 13 12 11 V mag 10 9 Figure 10. The number of WASP data points on each HAT planet as a function of magnitude. Red circles denote that the planet is detected by WASP, blue triangles are marginal detec- tions, black squares are non-detections. Some planets are labelled by their HAT number. Fig. 11 shows, for each HAT planet (up to HAT-P- 46 and HATS-3b), the number of data points it has in the WASP survey and whether we also detect it. There is some level of subjectivity in claiming a detection, since all WASP candidates are scrutinised by eye, and thus two plan- ets are shown as marginal detections. Most non-detections result from there being little or no WASP data (< 6000 data points), but it is worthwhile to review the remaining 7 cases. HAT-P-1 (Bakos et al. 2007) is bright and has 14 000 points in WASP data. It isn't detected owing to some ex- cess noise in some parts of our automated photometry, which prevents the transit-search algorithm from finding the tran- sits. HAT-P-9b (Shporer et al. 2009) has V = 12.3 and there are 11 000 WASP points. There is excess noise in some of the WASP photometry, preventing the search algorithm finding the transit. These two are the worst failures of the WASP techniques applied to the HAT sample. HAT-P-37 (Bakos et al. 2012) at V = 13.2 is fainter than any WASP planet, and also has a brighter eclipsing bi- nary 68′′ away, which misleads the WASP search algorithms. HAT-P-38 (Sato et al. 2012) is relatively faint (V = 12.6) and there is relatively limited WASP data (9000 points). Of the two marginal detections, HAT-P-19b (Hartman et al. 2011) has 12 000 WASP points but is faint (V = 12.9) and has an orbital period of 4.008 d. As a single-longitude survey, WASP's sampling hampers the finding of integer-day periods. HAT-P-15b (Kov´acs et al. 2010) is also below 12th magnitude (V = 12.2), and has relatively sparse coverage 12 Hellier et al. by WASP (9000 points), and also has a long orbital of 10.86 d, giving fewer transits than the typical 2–5-d hot Jupiters. Thus, in summary, at brighter than V = 12.8 and given 10 000 WASP points (roughly two seasons of data) we fail to detect only 2 out of 23 HAT-discovered planets (note that the number 23 doesn't include independent discover- ies by HAT of WASP planets). Thus we conclude that, in well-covered regions of sky, the WASP-South survey tech- niques give fairly complete discoveries for the sort of planets findable by WASP-like surveys. That means hot Jupiters around stars in the magnitude range V = 9–13, and exclud- ing crowded areas of sky such as the galactic plane. It also applies only to stars later than mid F, since hotter stars don't give good RV signals. The lower period limit of 0.8 d is likely a real cut-off in the distribution of hot Jupiters (e.g., Hellier et al. 2011), since we routinely look for candidates down to periods of 0.5 d. The longer period limit is primarily set by the amount of observational coverage, and our completeness will drop off above Porb ∼ 7 d, though this could increase to 'warm' Jupiters beyond ∼ 10 d as WASP continues to accumulate data. The upper limit in planetary size is also likely a real fea- ture of hot Jupiters, given the number of systems found with R = 1.8–2.0 RJup, but not above that range. The lower size limit of WASP discoveries is set primarily by red noise and so is harder to evaluate. Fig. 12 shows the transit depths of WASP planets, suggesting that we have a good discov- ery probability down to depths of ∼ 0.5% in the magnitude range V = 9–12. This is sufficient for R ∼> 0.9 RJup given R∗ < 1.3 R⊙, and ∼> 0.7 RJup given < 1.0 R⊙, and thus is sufficient for most hot Jupiters. ACKNOWLEDGEMENTS WASP-South is hosted by the South African Astronomical Observatory and we are grateful for their ongoing support and assistance. Funding for WASP comes from consortium universities and from the UK's Science and Technology Fa- cilities Council. TRAPPIST is funded by the Belgian Fund for Scientific Research (Fond National de la Recherche Sci- entifique, FNRS) under the grant FRFC 2.5.594.09.F, with the participation of the Swiss National Science Fundation (SNF). M. Gillon and E. Jehin are FNRS Research Asso- ciates. A.H.M.J. Triaud is a Swiss National Science Founda- tion Fellow under grant PBGEP2-145594 REFERENCES Anderson D. R. et al., 2012, MNRAS, 422, 1988 Bakos G. `A., Noyes R. W., Kov´acs G., Stanek K. Z., Sasselov, D. D., Domsa, I., 2004, PASP, 116, 266 Bakos, G. ´A. et al., 2007, ApJ, 656, 552 Bakos, G. ´A. et al., 2012, AJ, 144, 19 Barnes, S.A. 2007, ApJ, 669, 1167 Claret, A., 2000, A&A, 363, 1081 Collier Cameron, A., et al., 2007a, MNRAS, 375, 951 Collier Cameron, A., et al., 2007b, MNRAS, 380, 1230 Doyle, A.P. et al. 2013, MNRAS,, 428, 3164 Gillon, M., et al., 2013, A&A, 552, A82 Girardi, L., Bressan, A., Bertelli, G., Chiosi, C. 2000, A&AS, 141, 371 0.03 0.02 h t p e d l a n o i t c a r F 0.01 0 13 12 11 V mag 10 9 8 Figure 11. The transits depths of WASP planets as a function of host-star magnitude. Gray D.F., 2008, The observation and analysis of stellar photo- spheres, 3rd Edition (Cambridge University Press) Hartman, J. D. et al., 2011, ApJ, 726, 52 Hellier, C. et al., 2012, MNRAS, 426, 739 Hellier, C. et al., 2011, A&A, 535, L7 Jehin, E. et al., 2011, Messenger, 145, 2 Kov´acs, G. et al., 2010, ApJ, 724, 866 Lendl, M. et al., 2012, A&A, 544, A72 Maxted, P.F.L. et al. 2011, PASP, 123, 547 Pollacco, D., et al., 2006, PASP, 118, 1407 Pollacco, D., et al., 2008, MNRAS, 385, 1576 Sato, B. et al., 2012, PASJ, 64, 97 Sestito, P. & Randlich, S., 2005, A&A, 442, 615 Shporer, A., et al., 2009, ApJ, 690, 1393 Smith A. M. S. et al. 2012, AJ, 143, 81 Southworth J., 2011, MNRAS, 417, 2166 Zacharias, N., Finch, C. T., Girard, T. M., Henden, A., Bartlett, J. L., Monet, D. G., Zacharias, M. I. 2013, AJ, 145, 44 West, R. G. et al. 2013, A&A, submitted c(cid:13) 0000 RAS, MNRAS 000, range
1805.05016
1
1805
2018-05-14T05:55:32
Modeling Martian Atmospheric Losses over Time: Implications for Exoplanetary Climate Evolution and Habitability
[ "astro-ph.EP", "astro-ph.SR", "physics.space-ph" ]
In this Letter, we make use of sophisticated 3D numerical simulations to assess the extent of atmospheric ion and photochemical losses from Mars over time. We demonstrate that the atmospheric ion escape rates were significantly higher (by more than two orders of magnitude) in the past at $\sim 4$ Ga compared to the present-day value owing to the stronger solar wind and higher ultraviolet fluxes from the young Sun. We found that the photochemical loss of atomic hot oxygen dominates over the total ion loss at the current epoch whilst the atmospheric ion loss is likely much more important at ancient times. We briefly discuss the ensuing implications of high atmospheric ion escape rates in the context of ancient Mars, and exoplanets with similar atmospheric compositions around young solar-type stars and M-dwarfs.
astro-ph.EP
astro-ph
Draft version May 15, 2018 Typeset using LATEX twocolumn style in AASTeX62 8 1 0 2 y a M 4 1 . ] P E h p - o r t s a [ 1 v 6 1 0 5 0 . 5 0 8 1 : v i X r a Modeling Martian Atmospheric Losses over Time: Implications for Exoplanetary Climate Evolution and Habitability Chuanfei Dong,1, 2 Yuni Lee,3 Yingjuan Ma,4 Manasvi Lingam,5 Stephen Bougher,6 Janet Luhmann,7 Shannon Curry,7 Gabor Toth,6 Andrew Nagy,6 Valeriy Tenishev,6 Xiaohua Fang,8 David Mitchell,7 David Brain,8 and Bruce Jakosky8 1Department of Astrophysical Sciences, Princeton University, Princeton, NJ 08544, USA 2Princeton Center for Heliophysics, Princeton Plasma Physics Laboratory, Princeton University, Princeton, NJ 08540, USA 3NASA Goddard Space Flight Center, Greenbelt, MD 20771, USA 4Institute of Geophysics and Planetary Physics, University of California, Los Angeles, CA 90095, USA 5Institute for Theory and Computation, Harvard University, Cambridge MA 02138, USA 6Department of Climate and Space Sciences and Engineering, University of Michigan, Ann Arbor, MI 48109, USA 7Space Sciences Laboratory, University of California, Berkeley, CA 94720, USA 8Laboratory for Atmospheric and Space Physics, University of Colorado, Boulder, CO 80303, USA ABSTRACT In this Letter, we make use of sophisticated 3D numerical simulations to assess the extent of atmo- spheric ion and photochemical losses from Mars over time. We demonstrate that the atmospheric ion escape rates were significantly higher (by more than two orders of magnitude) in the past at ∼ 4 Ga compared to the present-day value owing to the stronger solar wind and higher ultraviolet fluxes from the young Sun. We found that the photochemical loss of atomic hot oxygen dominates over the total ion loss at the current epoch whilst the atmospheric ion loss is likely much more important at ancient times. We briefly discuss the ensuing implications of high atmospheric ion escape rates in the context of ancient Mars, and exoplanets with similar atmospheric compositions around young solar-type stars and M-dwarfs. 1. INTRODUCTION Mars has always represented an important target from the standpoint of planetary science (de Pater & Lissauer 2001), especially on account of its past, and perhaps even its current, biological potential (Jakosky & Phillips 2001; Ehlmann et al. 2016). In particular, ancient Mars (∼ 4 Ga) has attracted a great deal of atten- tion (Wordsworth 2016) because it may have possessed aqueous environments with water-rock interactions (Hurowitz et al. 2017), minerals (Ehlmann & Edwards 2014), biogenic elements (Fair´en et al. 2016), suitable energy sources for prebiotic chemistry (Lingam et al. 2018) and possibly oceans (di Achille & Hynek 2010); all of these factors could have enhanced the prospects for its habitability. Furthermore, some authors have sug- gested that the atmospheric composition and conditions of Noachian Mars were fairly similar to Hadean-Archean Earth (McKay 2010) and there is also a non-negligible probability that life could have been transferred from the former to the latter via lithopanspermia. However, one of the most striking differences between ancient and current Mars is that the former had a Corresponding author: Chuanfei Dong [email protected] thicker atmosphere compared to the present-day value (Jakosky & Phillips 2001), thereby making Noachian Mars potentially more conducive to hosting life. This discrepancy immediately raises the question of how and when the majority of the Martian atmosphere was lost, as well as the channels through which it occurred (Brain et al. 2016). There are compelling observational and theoretical reasons to believe that the majority of atmospheric escape must have occurred early in the planet's geological history (Lammer 2013; Jakosky et al. 2017), when the extreme ultraviolet (EUV) flux and the solar wind from the Sun were much stronger than today (Ribas et al. 2005; Boesswetter et al. 2010; Dong et al. 2017a). Moreover, the Martian dynamo shut down ∼ 4.1 Ga and Mars currently has only weak crustal magnetic fields (Lillis et al. 2013). Our understanding of present- day Martian atmospheric escape has improved greatly thanks to observations undertaken by, e.g., the Mars Atmosphere and Volatile EvolutioN mission (MAVEN) (Jakosky et al. 2015) in conjunction with detailed the- oretical modeling (Lee et al. 2015a; Fang et al. 2017; Bougher et al. 2017; Ma et al. 2017; Dong et al. 2018b). In this Letter, we will make use of the one-way coupled framework developed by Dong et al. (2015) and Lee et al. (2015b), known to accurately reproduce MAVEN observations, for studying the ion and pho- tochemical escape rates over the history of Mars while 2 self-consistently accounting for increased EUV and so- lar wind. The outline of the paper is as follows. In Sec. 2, we will describe the models and our numerical setup. We follow this up by describing and analyzing our results in Sec. 3. We conclude by summarizing the salient points in Sec. 4. 2. MODEL DESCRIPTIONS AND SETUP Here, we briefly outline the three sophisticated 3D global models for the Martian (i) ionosphere- thermosphere, (ii) exosphere, and (iii) magnetosphere. We simulate the ionosphere and thermosphere by employing the Mars Global Ionosphere Thermosphere Model (M-GITM) (Bougher et al. 2015). M-GITM is a 3D "whole atmosphere" (ground to exobase) non- hydrostatic model that includes all of the important ion- neutral chemistry and key radiative processes. M-GITM currently solves for neutral and ion densities, as well as neutral temperatures and winds around the globe. In this study, we initialize the Martian atmosphere by us- ing current parameters since it has been shown that both surface pressure (Dong et al. 2017b) and atmospheric composition (Brain et al. 2016) do not have a significant impact on atmospheric escape rates. Above certain altitudes (i.e., beyond the exobase), the fluid assumption is generally not valid anymore, thus a kinetic model has to be used to model the nearly col- lisionless exosphere. The dissociative recombination of O+ 2 (that not only splits the recombined molecular O2 into atomic O but also gives the resultant atomic O ad- ditional kinetic energy) is the most important reaction, primarily responsible for producing the dayside atomic oxygen exosphere. In order to simulate the 3D hot oxygen corona and the associated photochemical escape (i.e., loss of energetic atomic oxygen to space), we use the 3D Mars Adaptive Mesh Particle Simulator (AMPS) that solves the Boltzmann equation in the test-particle mode using the Direct Simulation Monte Carlo (DSMC) method. In Mars AMPS, the ionosphere-thermosphere inputs are taken from M-GITM (Lee et al. 2015b). Both M-GITM and Mars AMPS operate in the Geographic (GEO) coordinate system. 2 , CO+ Lastly, the 3-D BATS-R-US Mars multifluid MHD (MF-MHD) model starts from 100 km above the Mar- tian surface unlike its Earth counterpart that starts from ∼ 2-3 Earth radii. MF-MHD solves separate continuity, momentum and energy equations for four ion fluids H+, O+, O+ 2 (Najib et al. 2011; Dong et al. 2014). It includes a self-consistent ionosphere and the concomi- tant photochemistry such as photoionization, charge ex- change, electron impact ionization and ion-electron re- combination. In order to capture variations in the lower ionosphere, the radial resolution near the inner bound- ary is set to 5 km. The crustal magnetic fields are im- plemented based on the 60 degree harmonic expansion model adopted in Dong et al. (2014), the strongest of which we set to face nightside in this study for simplicity. For modeling the ionosphere, magnetosphere and the as- sociated atmospheric ion loss, we take the advantage of the existent one-way coupled framework, i.e., the M- GITM thermosphere and the AMPS oxygen exosphere are used as inputs for the MF-MHD model. MF-MHD runs in the Mars-centered Solar Orbital (MSO) coordi- nate system, where the +x-axis points from Mars to the Sun, the +z-axis is perpendicular to the Martian orbital plane and points northward, and the y-axis completes the right-hand system. We study four cases and the corresponding parameters are listed in Table 1. They include the extreme ultra- violet (EUV) strength, the time before present (tBP ), nominal solar wind density (nsw) and velocity (vsw), interplanetary magnetic field (BIM F ) and the angle associated with an away sector Parker spiral (φIM F ) (Ribas et al. 2005; Boesswetter et al. 2010). The rea- son we halt our analysis at ∼ 4 Ga is because little is known of the pre-Noachian period (Carr & Head 2010) and the solar wind parameters are very uncertain during this epoch (Lundin et al. 2007). 3. RESULTS AND DISCUSSION Figure 1 depicts the temperature and winds of the Martian thermosphere at ∼ 200 km for equinox condi- tions. An inspection of Figure 1 reveals that a high EUV flux is correlated with a hotter thermosphere. Therefore, the EUV heating of the thermosphere is self-consistently computed, which is very important for deriving the at- mospheric ion and photochemical losses. Figure 2 shows the atomic hot oxygen density dis- tribution in the meridian plane from AMPS based on the M-GITM input. The presented asymmetry in the hot oxygen density distribution is a result of higher O+ 2 abundance on the dayside than nightside. Compared to the current epoch with relatively low EUV flux, ancient Mars had a more intensive and extensive oxygen corona resulting from the enhanced O+ 2 density at higher EUV flux. In Figure 3, we present the MF-MHD calculation of O+ ion escaping from the planet. One of the features of the MF-MHD model is that it captures the asym- metric ion escape plume, resulting from the Lorentz force term in the individual ion momentum equations (Najib et al. 2011; Dong et al. 2014), as observed by MAVEN (Dong et al. 2017c). The asymmetric ion es- cape plume becomes less evident at earlier epochs be- cause of both the extended corona and the smaller ion gyroradius (∝ vsw/By) in this period. As seen from Fig- ure 3, more O+ ions escape from the planet at higher EUV and stronger solar wind. We list the atmospheric ion and photochemical escape rates in Table 1 and plot them in Figure 4; see also Fig. 4 of Luhmann et al. (1992) and Chassefi`ere et al. (2007) where similar calculations were undertaken based on less comprehensive methods. The calculated atmospheric es- cape rates in Table 1 are consistent with the density con- Input parameters (first six columns) used for the different cases (Ribas et al. 2005; Boesswetter et al. 2010), ion Table 1. escape rates (7th to 9th columns) and photochemical escape rate (last column). Note that 1 EUV (below) refers to the EUV flux received at Mars during the moderate phase of the solar cycle at the current epoch. 3 EUV tBP (Ga) nsw (cm−3) 1 3 6 10 0.0 2.77 3.57 3.93 2.51 10.26 24.75 46.99 vsw (km/s) BIM F (nT) φIM F (degree) O+ (s−1) O+ 2 (s−1) CO+ 401 578 726 858 3.01 7.06 12.17 18.16 58.0 64.6 68.2 70.5 1.8×1024 2.4×1025 2.4×1026 1.1×1027 2.6×1024 6.6×1024 9.3×1024 1.2×1025 2 (s−1) Ohot (s−1) 2.7×1025 8.5×1025 9.9×1025 1.0×1026 3.6×1023 1.4×1024 2.7×1024 4.1×1024 Solar Local Time 10 16 12 14 500.0 m/s 18 20 22 500 1 0 . 5 8 4 3EUV Solar Local Time 10 12 14 16 250.0 m/s 20 22 18 1EUV 5 9 . 8 0 3 300 2 4 6 8 e d u t i t a L 250 200 t e r u a r e p m e T 150 2 4 6 8 e d u t i t a L 2 4 6 8 e d u t i t a L 400 300 t e r u a r e p m e T 200 800 700 600 500 400 300 e r u t a r e p m e T Solar Local Time 10 12 14 16 500.0 m/s 20 22 18 6EUV 6 6 . 1 3 1 6 4 . 8 3 6 600 500 400 300 e r u t a r e p m e T 3 7 . 7 9 1 200 e d u t i t a L 2 4 6 8 Solar Local Time 10 16 12 14 500.0 m/s 22 20 18 10EUV 8 4 . 7 5 1 0 8 . 4 9 7 9 1 . 5 4 2 Figure 1. Color contours of temperature (in K) at ∼ 200 km for 1, 3, 6 and 10 EUV under equinox conditions (that approximately represents an average over one Mars orbit). The arrows in each panel indicate the relative magnitude (reference is given in top right corner) and the direction of the horizontal winds. The vertical axes (i.e. latitude) range between −90◦ and 90◦. Note that the colorbar varies in different panels. tour plots illustrated in Figures 2 and 3. It is noteworthy that the difference in photochemical loss between the 6 EUV and 10 EUV cases is nearly indistinguishable. The underlying reason is that the enhanced collision proba- bility between hot oxygen and thermal species in the ex- tended thermosphere can deflect hot/energetic particles more efficiently and thus decreases the escape probabil- ity of hot O (Zhao & Tian 2015). Interestingly, the pho- tochemical escape rate of atomic hot oxygen dominates over ion losses at the current epoch whilst the atmo- spheric ion escape rate becomes an order of magnitude larger than photochemical loss at ancient times, indi- cating that atmospheric losses are primarily controlled by ion escape for early Mars. In addition, compared to molecular ion species (O+ nant ion depleted at early epochs. 2 and CO+ 2 ), O+ is the domi- An analytic estimate of the total atmospheric escape rate N from weakly magnetized planets due to the ero- sion by the solar wind is via N ∝ (Rp/a)2 M⋆, where Rp and a are the planet's radius and semi-major axis respectively, while M⋆ denotes the solar mass-loss rate (Zendejas et al. 2010; Dong et al. 2018a); also see, e.g., Cravens et al. (2017a) for related analyses. We end up with N ∝ t−2.33, where t is the age of the star, because M⋆ exhibits this time dependence for solar-type stars (Wood et al. 2005). The photochemical atmospheric loss will be primarily driven by the EUV flux. Hence, if the photochemical escape rate NO is proportional to 4 Z Z Z Z n(Ohot) in cm-3 (1EUV) n(Ohot) in cm-3 (1EUV) 4 4 2 2 0 0 -2 -2 -4 -4 n(Ohot) in cm-3 (6EUV) n(Ohot) in cm-3 (6EUV) 4 4 2 2 0 0 -2 -2 -4 -4 3.5 3.0 2.5 2.0 1.5 1.0 0.5 3.5 3.0 2.5 2.0 1.5 1.0 0.5 n(Ohot) in cm-3 (3EUV) n(Ohot) in cm-3 (3EUV) n(Ohot) in cm-3 (10EUV) n(Ohot) in cm-3 (10EUV) 3.5 3.0 2.5 2.0 1.5 1.0 0.5 3.5 3.0 2.5 2.0 1.5 1.0 0.5 -4 -4 -2 -2 0 0 X X 2 2 4 4 -4 -4 -2 -2 2 2 4 4 0 0 X X Figure 2. Comparison of the hot oxygen density (in cm−3) distribution in the x-z meridian plane for different EUV cases using a logarithmic scale. n(O+) in cm-3 (1EUV) n(O+) in cm-3 (1EUV) n(O+) in cm-3 (3EUV) n(O+) in cm-3 (3EUV) Z Z Z Z 10 10 5 5 0 0 -5 -5 -10 -10 10 10 5 5 0 0 -5 -5 -10 -10 n(O+) in cm-3 (6EUV) n(O+) in cm-3 (6EUV) 4 3 2 1 0 -1 -2 -3 4 3 2 1 0 -1 -2 -3 n(O+) in cm-3 (10EUV) n(O+) in cm-3 (10EUV) 4 3 2 1 0 -1 -2 -3 4 3 2 1 0 -1 -2 -3 -15 -10 -15 -10 -5 -5 X X 0 0 5 5 -15 -10 -15 -10 0 0 5 5 -5 -5 X X Figure 3. Logarithmic scale contour plots of the O+ ion density (in cm−3) in the x-z meridian plane for different EUV and solar wind cases. the EUV flux ΦEU V (Cravens et al. 2017b), we obtain NO ∝ t−1.19 because ΦEU V displays this time depen- dence (Ribas et al. 2005). We find that the analytical trends are in good agreement with the numerical simu- lations at later epochs (t > 1 Gyr) but are less accurate for ancient Mars. From Figure 4, it can be seen that the total ion es- cape rate was & 100 times higher than today at ∼ 4 Ga. O + O + 2 CO + 2 Total ion hot O 1027 ) 1 − s ( e t a R e p a c s E 1026 1025 1024 −4.0 −3.5 −3.0 −2.5 −2.0 −1.5 −1.0 −0.5 0.0 Time (Ga) Figure 4. Calculated ion and photochemical escape rates over the Martian history. Hence, our results are consistent with Mars having lost much of its atmosphere early in its history, leading to the Martian climate changing from a warm and wet en- vironment in the past to the desiccated, frigid and thus inhospitable one documented today. Our simulations in- dicate that the total photochemical and ion atmospheric losses over the span ∼ 0-4 Ga are approximately equal to each other, and their sum amounts to ∼ 0.1 bar being lost over this duration. If we assume that the oxygen lost through a combination of ion and photochemical escape mechanisms was originally derived from surface water, we find ∼ 3.8 × 1017 kg of water has been lost from Mars between 0 to ∼ 4 Ga; this mass corresponds to a global surface depth of ∼ 2.6 m (the depth will be greater if the water bodies were more localized). The calculations do not include other potentially important loss processes such as sputtering; therefore, it provides a lower limit on the escape rates. Our result is more conservative com- pared to earlier studies (Zhang et al. 1993; Luhmann 1997; Chassefi`ere et al. 2007; Valeille et al. 2010) that predicted O(10) m of water was depleted over Martian history. Before proceeding further, recall that our analytic es- timates were expressible as N ∝ t−α and NO ∝ t−β with α ≈ 2.33 and β ≈ 1.19. Our choice of these values was motivated by the fact that the solar wind parameters used in the numerical simulations (Boesswetter et al. 2010) were consistent with Wood et al. (2005), and the ancient EUV fluxes were based on Ribas et al. (2005). In actuality, the exponent α is not tightly constrained and recent studies favor α . 1 (Cranmer & Saar 2011; Johnstone et al. 2015). Similarly, other methods for cal- culating the EUV flux over time - based, for instance, on X-ray (Chadney et al. 2015) and Ly-α (Linsky et al. 2014) emission - lead to different values of β. Since the integrated mass lost over the duration 0 to ∼ 4 Ga (due to ion loss mechanisms) is proportional to (α − 1)−1 (cid:2)(23/3)α−1 − 1(cid:3), we find that the overall es- timated value changes at most by one order of magni- tude for α ∈ (0.5, 2.5); the factor of 23/3 occurs because the current age of 4.6 Gyr (0 Ga) is divided by 0.6 Gyr (∼ 4 Ga). The same conclusion also holds true for pho- tochemical escape since the corresponding mass lost is found via the substitution α → β. Hence, it seems rea- sonable to conclude that the basic conclusions in this Section are not greatly altered if the values of α or β are changed. 4. CONCLUSION In our solar system, Mars represents a classic example of a planet where planetary habitability has been unam- biguously affected by atmospheric losses. In this Letter, we have studied the atmospheric ion and photochemical escape rates from Mars over time. We found that the atmospheric ion escape rates vary significantly over the planet's history, ranging from O (cid:0)1027(cid:1) s−1 at ∼ 4 Ga to O (cid:0)1024(cid:1) s−1 in the present epoch. The correspond- ing photochemical escape rate lies between O (cid:0)1026(cid:1) s−1 at ∼ 4 Ga and O (cid:0)1025(cid:1) s−1 today. Therefore, our sim- ulations are consistent with the idea that Mars could have transitioned from having a thick atmosphere and global water bodies to its current state with a tenuous atmosphere and arid conditions quite early in its history. The total atmospheric loss over time predicted by simu- lations may, perhaps, be tested against observations (to some degree) by using isotope ratios, since the lighter ions are picked up preferentially compared to the heavier ions, akin to the method used by Jakosky et al. (2017). However, we caution the reader that the uncertainties involved with the solar wind and EUV flux increase as we move towards more ancient epochs, implying that our conclusions concerning atmospheric losses over time will also be subject to a certain degree of variability. Our results also have implications for the rapidly ex- panding domain of exoplanets if one views Mars as a prototype for small rocky exoplanets. Water can be lost from the atmospheres of exoplanets in the habitable zone (HZ) of M-dwarfs over relatively fast timescales (Bolmont et al. 2017), compared to heavier molecules (e.g. CO2). Since the total number of rocky exoplan- 5 ets in the HZ of M-dwarfs is expected to be ∼ 1010 (Dressing & Charbonneau 2015), the possibility of exo- planets with atmospheric compositions similar to Venus and Mars cannot be ruled out. Our work demonstrates that such exoplanets, as well as those around young solar-type stars, could be subjected to high atmospheric escape rates early in their history. For exoplanets or- biting M-dwarfs, the situation could be even worse due to the more intense particle and radiation environments that exoplanets experience in their close-in HZs. If the escape rates scale as 1/a2, it is possible for ∼ 100 bars to be lost from a Mars-like exoplanet in the HZ of an M-dwarf of mass ∼ 0.1 M⊙ over a span of ∼ 4.0 Gyr. Equivalently, this corresponds to a global water depth of ∼ 2.6 km being depleted if the source of atmospheric oxygen was surface water. In turn, if their atmospheres and oceans end up being altogether depleted over sub- Gyr timescales, this could lead to detrimental effects in- sofar their habitability is concerned (Dong et al. 2017b, 2018a; Lingam & Loeb 2018). Thus, from a broader perspective, our work demon- strates that atmospheric loss is not static but dynamical in nature and that high escape rates will typically occur early in the host star's history. It is therefore necessary to take this time-dependence into account when model- ing atmospheric loss from early Mars, and Mars-like exo- planets in the future. We also expect to incorporate the impact of extreme space weather events that are highly frequent on young and/or low-mass stars in our future study (Dong et al. 2017a) to model the atmospheric loss and evolution of early Mars/Mars-like planets. This research was supported by NASA grant NNH10CC04C through MAVEN Project and NASA grant 80NSSC18K0288. Resources supporting this work were provided by the NASA High-End Computing (HEC) Program through the NASA Advanced Supercomputing (NAS) Divi- sion at Ames Research Center. The Space Weather Modeling Framework that contains the BATS-R-US code used in this study is publicly available from http://csem.engin.umich.edu/tools/swmf. For distri- bution of the model results used in this study, please contact the corresponding author. REFERENCES Boesswetter, A., Lammer, H., Kulikov, Y., Motschmann, Brain, D. A., Bagenal, F., Ma, Y.-J., Nilsson, H., & U., & Simon, S. 2010, Planet. Space Sci., 58, 2031 Stenberg Wieser, G. 2016, J. Geophys. Res. E, 121, 2364 Bolmont, E., Selsis, F., Owen, J. E., et al. 2017, Mon. Not. R. Astron. Soc., 464, 3728 Bougher, S. W., Pawlowski, D., Bell, J. M., et al. 2015, J. Geophys. Res. E, 120, 311 Carr, M. H., & Head, J. W. 2010, Earth and Planetary Science Letters, 294, 185 Chadney, J. M., Galand, M., Unruh, Y. C., Koskinen, T. T., & Sanz-Forcada, J. 2015, Icarus, 250, 357 Chassefi`ere, E., Leblanc, F., & Langlais, B. 2007, Planet. Bougher, S. W., Roeten, K. J., Olsen, K., et al. 2017, J. Space Sci., 55, 343 Geophys. Res. A, 122, 1296 Cranmer, S. R., & Saar, S. H. 2011, Astrophys. J., 741, 54 6 Cravens, T. E., Hamil, O., Houston, S., et al. 2017a, J. Johnstone, C. P., Gudel, M., Brott, I., & Luftinger, T. Geophys. Res. A, 122, 10 Cravens, T. E., Rahmati, A., Fox, J. L., et al. 2017b, J. Geophys. Res. A, 122, 1102 de Pater, I., & Lissauer, J. J. 2001, Planetary Sciences (Cambridge Univ. Press) di Achille, G., & Hynek, B. M. 2010, Nat. Geosci., 3, 459 Dong, C., Bougher, S. W., Ma, Y., et al. 2014, Geophys. Res. Lett., 41, 2708 2015, Astron. Astrophys., 577, A28 Lammer, H. 2013, Origin and Evolution of Planetary Atmospheres: Implications for Habitability, Springer Briefs in Astronomy (Springer), doi:10.1007/978-3-642-32087-3 Lee, Y., Combi, M. R., Tenishev, V., et al. 2015a, Geophys. Res. Lett., 42, 9015 Lee, Y., Combi, M. R., Tenishev, V., Bougher, S. W., & Lillis, R. J. 2015b, J. Geophys. Res. E, 120, 1880 Dong, C., Huang, Z., Lingam, M., et al. 2017a, Astrophys. Lillis, R. J., Robbins, S., Manga, M., Halekas, J. S., & Frey, J. Lett., 847, L4 H. V. 2013, J. Geophys. Res. E, 118, 1488 Dong, C., Jin, M., Lingam, M., et al. 2018a, Proc. Natl. Lingam, M., Dong, C., Fang, X., Jakosky, B. M., & Loeb, Acad. Sci. USA, 115, 260 Dong, C., Lingam, M., Ma, Y., & Cohen, O. 2017b, Astrophys. J. Lett., 837, L26 Dong, C., Bougher, S. W., Ma, Y., et al. 2015, J. Geophys. Res. A, 120, 7857 -. 2018b, J. Geophys. Res. A, arXiv:1804.00937 Dong, Y., Fang, X., Brain, D. A., et al. 2017c, J. Geophys. Res. A, 122, 4009 A. 2018, Astrophys. J., 853, 10 Lingam, M., & Loeb, A. 2018, Int. J. Astrobiol., 17, 116 Linsky, J. L., Fontenla, J., & France, K. 2014, Astrophys. J., 780, 61 Luhmann, J. G. 1997, J. Geophys. Res., 102, 1637 Luhmann, J. G., Johnson, R. E., & Zhang, M. H. G. 1992, Geophys. Res. Lett., 19, 2151 Lundin, R., Lammer, H., & Ribas, I. 2007, SSRv, 129, 245 Ma, Y. J., Russell, C. T., Fang, X., et al. 2017, J. Geophys. Dressing, C. D., & Charbonneau, D. 2015, Astrophys. J., Res. A, 122, 1714 807, 45 McKay, C. P. 2010, Cold Spring Harb. Perspect. Biol., 2, Ehlmann, B. L., & Edwards, C. S. 2014, Annu. Rev. Earth a003509 Planet. Sci., 42, 291 Ehlmann, B. L., Anderson, F. S., Andrews-Hanna, J., et al. 2016, J. Geophys. Res. E, 121, 1927 Fair´en, A. G., Dohm, J. M., Rodr´ıguez, J. A. P., et al. 2016, Astrobiology, 16, 143 Fang, X., Ma, Y., Masunaga, K., et al. 2017, J. Geophys. Res. A, 122, 4117 Hurowitz, J. A., Grotzinger, J. P., Fischer, W. W., et al. Najib, D., Nagy, A. F., T´oth, G., & Ma, Y. 2011, J. Geophys. Res. A, 116, A05204 Ribas, I., Guinan, E. F., Gudel, M., & Audard, M. 2005, Astrophys. J., 622, 680 Valeille, A., Bougher, S. W., Tenishev, V., Combi, M. R., & Nagy, A. F. 2010, Icarus, 206, 28 Wood, B. E., Muller, H.-R., Zank, G. P., Linsky, J. L., & Redfield, S. 2005, Astrophys. J. Lett., 628, L143 Wordsworth, R. D. 2016, Annu. Rev. Earth Planet. Sci., 44, 2017, Science, 356, aah6849 381 Jakosky, B. M., & Phillips, R. J. 2001, Nature, 412, 237 Zendejas, J., Segura, A., & Raga, A. C. 2010, Icarus, 210, Jakosky, B. M., Slipski, M., Benna, M., et al. 2017, Science, 539 355, 1408 Jakosky, B. M., Lin, R. P., Grebowsky, J. M., et al. 2015, Space Sci. Rev., 195, 3 Zhang, M. H. G., Luhmann, J. G., Bougher, S. W., & Nagy, A. F. 1993, J. Geophys. Res., 98, 10915 Zhao, J., & Tian, F. 2015, Icarus, 250, 477
1612.04166
1
1612
2016-12-13T13:44:11
Peculiar architectures for the WASP-53 and WASP-81 planet-hosting systems
[ "astro-ph.EP", "astro-ph.SR" ]
We report the detection of two new systems containing transiting planets. Both were identified by WASP as worthy transiting planet candidates. Radial-velocity observations quickly verified that the photometric signals were indeed produced by two transiting hot Jupiters. Our observations also show the presence of additional Doppler signals. In addition to short-period hot Jupiters, we find that the WASP-53 and WASP-81 systems also host brown dwarfs, on fairly eccentric orbits with semi-major axes of a few astronomical units. WASP-53c is over 16 $M_{\rm Jup} \sin i_{\rm c}$ and WASP-81c is 57 $M_{\rm Jup} \sin i_{\rm c}$. The presence of these tight, massive companions restricts theories of how the inner planets were assembled. We propose two alternative interpretations: a formation of the hot Jupiters within the snow line, or the late dynamical arrival of the brown dwarfs after disc-dispersal. We also attempted to measure the Rossiter--McLaughlin effect for both hot Jupiters. In the case of WASP-81b we fail to detect a signal. For WASP-53b we find that the planet is aligned with respect to the stellar spin axis. In addition we explore the prospect of transit timing variations, and of using Gaia's astrometry to measure the true masses of both brown dwarfs and also their relative inclination with respect to the inner transiting hot Jupiters.
astro-ph.EP
astro-ph
Mon. Not. R. Astron. Soc. 000, 1–23 (2014) Printed 26 June 2018 (MN LATEX style file v2.2) Peculiar architectures for the WASP-53 and WASP-81 planet-hosting systems(cid:63) Amaury H. M. J. Triaud1,2,3,4†, Marion Neveu-VanMalle1,5, Monika Lendl6,1, David R. Anderson7, Andrew Collier Cameron8, Laetitia Delrez9, Amanda Doyle10, Michael Gillon9, Coel Hellier7, Emmanuel Jehin9, Pierre F. L. Maxted7, Damien S´egransan1, Barry Smalley7, Didier Queloz5,1, Don Pollacco10, John Southworth7, Jeremy Tregloan-Reed11,7, St´ephane Udry1, Richard West10 1Observatoire Astronomique de l'Universit´e de Gen`eve, Chemin des Maillettes 51, CH-1290 Sauverny, Switzerland 2Institute of Astronomy, University of Cambridge, Madingley Road, CB3 0HA, Cambridge, United Kingdom 3Centre for Planetary Sciences, University of Toronto at Scarborough, 1265 Military Trail, Toronto, ON, M1C 1A4, Canada 4Department of Astronomy & Astrophysics, University of Toronto, Toronto, ON, M5S 3H4, Canada 5Cavendish Laboratory, J J Thomson Avenue, Cambridge CB3 0HE, UK 6Space Research Institute, Austrian Academy of Sciences, Schmiedlstr. 6, 8042 Graz, Austria 7Astrophysics Group, Keele University, Staffordshire, ST5 5BG, UK 8SUPA, School of Physics & Astronomy, University of St. Andrews, North Haugh, KY16 9SS, St. Andrews, Fife, Scotland, UK 9Institut d'Astrophysique et de G´eophysique, Universit´e de Li`ege, All´ee du 6 Aout 17, Sart Tilman, 4000 Li`ege 1, Belgium 10Department of Physics, University of Warwick, Coventry CV4 7AL, UK 11SETI Institute, Mountain View, CA, 94043, USA Accepted ?. Received ?; in original form ? ABSTRACT We report the detection of two new systems containing transiting planets. Both were identified by WASP as worthy transiting planet candidates. Radial-velocity observa- tions quickly verified that the photometric signals were indeed produced by two transit- ing hot Jupiters. Our observations also show the presence of additional Doppler signals. In addition to short-period hot Jupiters, we find that the WASP-53 and WASP-81 sys- tems also host brown dwarfs, on fairly eccentric orbits with semi-major axes of a few astronomical units. WASP-53c is over 16 MJup sin ic and WASP-81c is 57 MJup sin ic. The presence of these tight, massive companions restricts theories of how the inner planets were assembled. We propose two alternative interpretations: a formation of the hot Jupiters within the snow line, or the late dynamical arrival of the brown dwarfs after disc-dispersal. We also attempted to measure the Rossiter–McLaughlin effect for both hot Jupiters. In the case of WASP-81b we fail to detect a signal. For WASP-53b we find that the planet is aligned with respect to the stellar spin axis. In addition we explore the prospect of transit timing variations, and of using Gaia's astrometry to measure the true masses of both brown dwarfs and also their relative inclination with respect to the inner transiting hot Jupiters. Key words: planetary systems – planets and satellites: individual: WASP-81, WASP- 53 – binaries: eclipsing – brown dwarfs 6 1 0 2 c e D 3 1 . ] P E h p - o r t s a [ 1 v 6 6 1 4 0 . 2 1 6 1 : v i X r a (cid:63) using data collected at ESO's La Silla Observatory, Chile: HARPS on the ESO 3.6m (Prog IDs 087.C-0649, 089.C-0151, 090.C-0540, 091.C-0184 & 093.C-0474), the ESO NTT (Prog ID 088.C-0204), the Swiss Euler telescope, and TRAPPIST. The data is publicly available at the CDS Strasbourg and on demand to the main author. c(cid:13) 2014 RAS 1 FOREWORDS The discovery of 51 Peg b (Mayor & Queloz 1995) initi- ated a debate about the origin (formation and evolution) of † E-mail: [email protected] 2 Amaury H. M. J. Triaud et al. (a) WASP-53 (b) WASP-81 Figure 1. Variations in the magnitude of WASP-53 and WASP-81 leading to their identification as transiting planet candidates. The grey dots show individual WASP measurements, whereas the black dots show the median magnitude within each of 200 phase bins. hot Jupiters that continues to rage to this day. The ques- tion is whether they formed in situ, within the so-called snow line (Bodenheimer, Hubickyj & Lissauer 2000; Baty- gin, Bodenheimer & Laughlin 2015; Lee & Chiang 2015), or beyond, followed by inward migration (Pollack et al. 1996; Alibert et al. 2005; Helled et al. 2014). The second hypothesis needs an explanation of how hot Jupiters have migrated, either via angular momentum transfer with their protoplanetary disc (Lin, Bodenheimer & Richardson 1996; Ward 1997; Baruteau et al. 2014) or thanks to dynamical in- teractions followed by tidal circularisation (high-eccentricity migration) (Rasio & Ford 1996; Wu, Murray & Ramsahai 2007; Naoz et al. 2011; Petrovich 2015a). Any framework has to explain why gas giants are found both close and far from their host star (not least Jupiter and Saturn), and also that they frequently orbit on planes that are inclined, sometimes retrograde, with respect to the equatorial plane of their host star (H´ebrard et al. 2008; Winn et al. 2009; Schlaufman 2010; Anderson et al. 2010; Triaud et al. 2010; Albrecht et al. 2012a; Lendl et al. 2014; Winn & Fabrycky 2015). Observations of the spin–orbit angle, thanks to the Rossiter–McLaughlin effect (Queloz et al. 2000), were ini- tially thought to provide a clean test between disc-driven migration and dynamical and tidal migration (Gaudi & Winn 2007). While some studies indicate that observations are compatible with planets undergoing orbital realignment (Triaud 2011; Albrecht et al. 2012b; Dawson 2014), addi- tional theoretical arguments imply that misalignments can arise via a variety of processes. They can be primordial, with the planets forming on inclined planes (Thies et al. 2011; Lai, Foucart & Lin 2011; Batygin 2012; Lai 2014; Spalding & Batygin 2015), or they can arise later (C´ebron et al. 2011; Rogers et al. 2013). Thus inclined hot Jupiters are compat- ible with both dynamical interactions and with disc-driven migration. Obviously, disc-driven migration and high-eccentricity migration could both be correct and each produce a frac- tion of the hot Jupiters, as recent observation appear to im- ply (Anderson et al. 2015). Finally, results from Guillochon, Ramirez-Ruiz & Lin (2011) suggest that high-eccentricity migration requires in most cases some amount of disc-driven migration. In this paper, we spend the first sections to describe the discovery of two new planetary systems. Both present a tight hierarchical architecture, composed of an inner, transiting, hot Jupiter, and an outer brown-dwarf companion. WASP- 53 & WASP-81 are reminiscent of a challenge proposed by Hatzes & Wuchterl (2005) to test disc-driven migration. If both brown dwarfs were on their current orbits during the protoplanetary phase, they would have truncated the disc within the snow line thus preventing disc-driven migration and only allowing in-situ formation. In the final sections we describe a number of formation and evolution scenarios, some involving disc-driven migra- tion and others not. We conclude that a number of scenarios will become testable soon, with the arrival of precise astro- metric measurements produced by ESA's Gaia satellite. We also compute the equilibrium eccentricities of the inner gas giants (Mardling 2007) and find it is unlikely that their k2 Love number (Batygin, Bodenheimer & Laughlin 2009) will be measured soon. 2 WASP IDENTIFICATION The WASP survey1 (Pollacco et al. 2006) consists of two sets of eight 11-cm refractive telescope mounted together. One set is located at the Observatorio del Roque de Los Muchachos, La Palma (Spain), while the other is installed in Sutherland, hosted by the South African Astronomical Observatory. WASP has observed in excess of 30 million stars since 2004, thousands of times each. The photometric data reduction and the candidate selection are described in Collier Cameron et al. (2007). WASP-53 (2MASS J02073820–2039426; K3, J = 10.959) and WASP-81 (2MASS J20164989+0317385; F9, J = 11.263) have been observed 21 120 and 13 292 times by WASP. They were two unremarkable and anonymous stars before a short, box-like photometric signals were iden- tified at P = 3.309866 d and P = 2.716554 d, respectively. 1 https://wasp-planets.net c(cid:13) 2014 RAS, MNRAS 000, 1–23 WASP-53b was sent for radial-velocity verification in 2010- 08-10, and WASP-81b in 2011-05-09, with the first spectra acquired on 2010-12-05 and 2011-09-29. The WASP data are shown in Fig. 1. 3 PHOTOMETRIC OBSERVATIONS All follow-up photometric timeseries, their dates, filters, number of data points and the detrending functions that were used in the analysis, are detailed in Table C1. Figure 3 shows the corrected photometry for WASP-53, and Figure 4 shows WASP-81. Raw fluxes and residuals are shown in Fig- ure C1 & C2, respectively. Below, we provide some details on the observations and reduction, although we encourage readers to refer to cited papers for fuller information. 3.1 EulerCam EulerCam is mounted at the 1.2m Euler Swiss telescope lo- cated at ESO La Silla Observatory (Chile) It has a pixel scale of 0.215" for a field of view of 14.7×14.5'. The telescope is an alt-azimuthal design. EulerCam is mounted behind a field de-rotator. To ensure the best photometric precision, each star is kept on the same pixels, using a digital feed- back scheme that compares the newly acquired frame with a composite of earlier frames and their offset from a recorded position. We obtained four transits of WASP-53b, all using an r'-Gunn filter, and with a slightly defocused telescope to improve the observing efficiency and PSF sampling. Two of the transit, on 2011-09-22 and 2012-12-02 were scheduled to coincide with the radial-velocity time-series obtained using HARPS (see Sect 4.2). WASP-81 was observed with Euler- Cam throughout three transits, one of which, on 2013-08-05, was obtained while HARPS collected a radial-velocity time- series. Our 2012-06-07 observations were obtained using an I-Cousins filter, and a focused telescope. The telescope was defocused slightly for latter two observations, for which a z'- Gunn (2012-09-24) and an r'-Gunn (2013-08-05) were used. All EulerCam images were reduced using standard im- age correction methods and lightcurves were obtained us- ing differential aperture photometry, with a careful selec- tion of aperture and reference stars. For further details on the EulerCam instrument and the associated data reduction procedures, please refer to Lendl et al. (2012). The times of observations are provided in JD(UTC), changed later to BJD(TDB) during the global analysis. 3.1.1 A nearby stellar source During spectroscopic observations an additional light source was noticed near WASP-81A on CORALIE's guiding cam- era images. On 2014-04-29 we obtained focused images of the WASP-81 system with EulerCam through Geneva B (6 images), Geneva V (3 images) and r(cid:48)-Gunn (3 images) fil- ters (Fig. 2). We used astronomy.net (Lang et al. 2010) to calculate a precise astrometric solution and performed PSF fitting on the images using daophot (Stetson 1987). The results are in Table 1. c(cid:13) 2014 RAS, MNRAS 000, 1–23 The WASP-53 and 81 systems 3 Figure 2. A 1.5'x1.5' field of view centred on WASP-81, obtained with EulerCam, clearly showing a visual companion to the North. The pair is most likely unrelated. Table 1. Observation parameters for the companion source to WASP-81A Filter ∆ Mag Mag 5.64 ± 0.03 5.16 ± 0.01 4.87 ± 0.01 BG V G r(cid:48) weighted mean: 18.78 ± 0.07 17.61 ± 0.02 17.14 ± 0.04 separation (arcsec) 4.34 ± 0.02 4.33 ± 0.01 4.32 ± 0.01 4.33 ± 0.01 PA (degrees) 3.5 ± 0.2 3.5 ± 0.3 3.5 ± 0.3 3.5 ± 0.3 We extracted magnitudes for the visual companion rela- tive to the primary target. Combining with apparent magni- tudes for WASP-81A (Table 2), and using a E(B − V )=0.05 (Sect. 4), we obtained apparent magnitudes for the visual companion. We find that the companion has colours consis- tent with a K3–K4 spectral type. If it is a dwarf, this im- plies a distance modulus of order 10.2 (1.1 kpc), compared to 8.0 (400 pc) for WASP-81A; if it were a giant it would be further away still. If we place the companion on a main- sequence isochrone (Marigo et al. 2008), and if WASP-81A were at the same distance, WASP-81A would have to be a 2-Gyr old, 1.4–1.6 M(cid:12) star, contradicting our spectroscopic analysis as well as the mean stellar density obtained from the tranait. If, instead, WASP-81A is on the main sequence, then the companion has to be below the main sequence to be at the same distance. The two objects are most therefore likely unrelated. We placed an HR diagram of the pair into the appendices. For future reference we provide here the position of the visual companion. Gaia will soon produce parallaxes and proper motions that should confirm our analysis. If instead they are found at the same distance, implying they are grav- itationally bound, then the companion must be an M dwarf with a much redder B − V than the one we measured. 4 Amaury H. M. J. Triaud et al. Figure 3. Photometry on WASP-53 at the time of transit, using EulerCam, TRAPPIST and the NTT. Left, we have the detrended data with, in red, the most likely model. Right, we show the residuals. Raw photometry and full models are available in Fig. C1. 3.2 TRAPPIST Both WASP systems were observed with the 0.6-m TRAP- PIST robotic telescope (TRAnsiting Planets and PlanetesI- mals Small Telescope), also located at La Silla. Three tran- sits were obtained on WASP-53, and seven on WASP-81. TRAPPIST is equipped with a thermoelectrically-cooled 2K×2K CCD, which has a pixel scale of 0.6" that trans- lates into a 22'×22' field of view. For details of TRAPPIST, see Gillon et al. (2011) and Jehin et al. (2011). Two filters were used: a blue-blocking filter that has a transmittance of > 90% from 500 nm to beyond 1000 nm, and an"I + z(cid:48)" filter that has a transmittance of > 90% from 750 nm to beyond 1100 nm. During the runs the positions of the stars on the chip were maintained to within a few pixels thanks to a"software guiding" system that regularly derives an astro- metric solution for the most recently acquired image and sends pointing corrections to the mount if needed. After a standard pre-reduction (bias, dark, and flatfield correc- tion), the stellar fluxes were extracted from the images using the iraf/daophot2 aperture photometry software (Stetson 1987). For each light curve we tested several sets of reduc- tion parameters and kept the one giving the most precise photometry for the stars of similar brightness as the target. After a careful selection of reference stars the transit light curves were finally obtained using differential photometry. Some light curves were affected by a meridian flip; that is, the 180◦ rotation that the German equatorial mount tele- scope has to undergo when the meridian is reached. This movement results in different positions of the stellar images on the detector before and after the flip, and thus in a pos- sible jump of the differential photometry at the time of the flip. We have accounted for this in our light curve analysis by including a normalization offset in our model at the time of the flip (see Fig. C1 & C2 as well as Table C1). More details on data acquisition and data reduction can be found 2 IRAF is distributed by the National Optical Astronomy Obser- vatory, which is operated by the Association of Universities for Research in Astronomy, Inc., under cooperative agreement with the National Science Foundation. c(cid:13) 2014 RAS, MNRAS 000, 1–23 −2−1 0 1 20.850.900.951.00Hours, from mid−transitFlux . . . .EulerCam r'2011−07−22EulerCam r'2011−09−03TRAPPIST BB2011−09−13EulerCam r'2011−09−13EulerCam r'2011−09−23EFOSC2 r'2011−10−26TRAPPIST BB2012−07−30TRAPPIST I+z'2012−11−03−2−1 0 1 20.850.900.951.00Hours, from mid−transitFlux . . . .EulerCam r'2011−07−22EulerCam r'2011−09−03TRAPPIST BB2011−09−13EulerCam r'2011−09−13EulerCam r'2011−09−23EFOSC2 r'2011−10−26TRAPPIST BB2012−07−30TRAPPIST I+z'2012−11−03 The WASP-53 and 81 systems 5 Figure 4. Photometry on WASP-81 at the time of transit, using EulerCam and TRAPPIST. Left, we show the detrended data with, in red, the most likely model. Right, we show the residuals. Raw photometry and full models are available in Fig. C2. in Gillon et al. (2013) and Delrez et al. (2014), whose proce- dures were followed here as well. The times of observations are provided in JD(UTC), changed later to BJD(TDB) dur- ing the global analysis. 3.3 New Technology Telescope One transit of WASP-53 was observed using the EFOSC2 instrument on the NTT at ESO's observation of La Silla (ProgID 088.C-0204, PI Tregloan-Reed; see Tregloan-Reed & Southworth (2013) for further details of this observ- ing run). The instrument has a 2K×2K CCD covering a 4.1'×4.1' field of view, and a pixel scale of 0.12". Observa- tions were curtailed shortly after the end of the transit due to the onset of daytime. We observed through a Gunn r fil- ter (ESO filter #784) with heavy defocussing and exposure times of 150s. The data were reduced using the defot pipeline (Southworth et al. 2009), which utilises an aperture- photometry routine aper.pro ported from daophot (Stet- son 1987). The radius of the inner aperture was 45 pixels c(cid:13) 2014 RAS, MNRAS 000, 1–23 and the sky annulus extended from 60 to 100 pixels. Debi- assing and flat-fielding the data did not make a significant difference to the results, so we neglected these calibrations. A differential-photometry light curve was obtained for WASP-53 versus an ensemble comparison star. Due to the small field of view of NTT/EFOSC2, we were able to use only three comparison stars, all of which were at least two magnitudes fainter than WASP-53 in the r-band. The weights of the comparison stars, used in summing their fluxes to create the ensemble comparison star, were op- timised to minimise the scatter in the data outside tran- sit. Finally, the timestamps were moved to the BJD(TDB) timescale using routines from Eastman, Siverd & Gaudi (2010). 4 SPECTROSCOPIC OBSERVATIONS We collected 98 CORALIE spectra on WASP-53 between the dates of 2010-12-04 and 2016-11-21, as well as 83 HARPS spectra between the date of 2011-08-28 and 2014-09-28. On WASP-81, we acquired 67 spectra with −2 0 20.750.800.850.900.951.00Hours, from mid−transitFlux . . . .TRAPPIST I+z'2011−09−26TRAPPIST I+z'2012−05−20TRAPPIST I+z'2012−05−31TRAPPIST BB2012−06−19EulerCam r'2012−07−08TRAPPIST I+z'2012−07−19EulerCam r'2012−09−24TRAPPIST BB2013−07−07EulerCam r'2013−08−06TRAPPIST BB2013−08−06−2 0 20.750.800.850.900.951.00Hours, from mid−transitFlux . . . .TRAPPIST I+z'2011−09−26TRAPPIST I+z'2012−05−20TRAPPIST I+z'2012−05−31TRAPPIST BB2012−06−19EulerCam r'2012−07−08TRAPPIST I+z'2012−07−19EulerCam r'2012−09−24TRAPPIST BB2013−07−07EulerCam r'2013−08−06TRAPPIST BB2013−08−06 6 Amaury H. M. J. Triaud et al. CORALIE from 2011-09-28 to 2015-07-08, and 32 using HARPS between 2013-04-20 and 2016-10-21. Table 2. Stellar parameters of WASP-53A and WASP-81A 4.1 Spectral analysis The analysis was performed on the standard pipeline re- duction products, using the methods given in Doyle et al. (2013). The Hα line was used to give an initial estimate of the effective temperature (Teff ). The surface gravity (log g(cid:63)) was determined from the Ca i line at 6439A, along with the Na i D lines. Additional Teff and log g(cid:63) diagnotics were per- formed using the Fe lines. An ionisation balance between Fe i and Fe ii was required, along with a null dependence of the abundance on either equivalent width or excitation potential. This null dependence was also required to de- termine the microturbulence (ξt). The parameters obtained from the analysis are listed in Table 2. Some of those pa- rameters will later be employed as priors to compute the most likely physical parameters of each system. The ele- mental abundances were determined from equivalent width measurements of several clean and unblended lines, and ad- ditional least squares fitting of lines was performed when required. The quoted error estimates include that given by the uncertainties in Teff , log g(cid:63), and ξt, as well as the scatter due to measurement and atomic data uncertainties. The individual HARPS spectra of WASP-53 were co- added to produce a single spectrum with an average S/N in excess of 100:1. The macroturbulence was assumed to be zero, since for mid-K stars it is expected to be lower than that of thermal broadening (Gray 2008). The projected stel- lar rotation velocity (v sin i(cid:63)) was determined by fitting the profiles of several unblended lines, yielding an upper limit of 2.7 ± 0.3 km s−1 for WASP-53. . There is no significant detection of lithium in WASP-53. The equivalent width up- per limit of 13mA, correspond to an abundance upper limit of log A(Li) < 0.32 ± 0.16. This implies an age of at least several hundreds Myr (Sestito & Randich 2005). Similarly, the combination of the HARPS spectra ob- tained on WASP-81 produced a combined spectrum with an average S/N of 60:1. Here, we used a macroturbulent value of 3.84±0.73 km s−1 after a relation from Doyle et al. (2014). v sin i(cid:63) was found to be 1.20 ± 0.69 km s−1. 4.2 Radial velocities The spectra were reduced using the standard CORALIE and HARPS reduction software. They have been shown to reach remarkable precision and accuracy, reaching below 1 m s−1 (e.g. Lovis et al. (2006); Marmier et al. (2013); L´opez-Morales et al. (2014)). We extracted the radial veloc- ities for both stars by cross-correlating each spectrum with a binary mask. For WASP-53A we used a mask correspond- ing to a K5 spectral type, and for WASP-81A we employed a G2 mask. We fit the corresponding cross-correlation func- tion with a Gaussian, whose mean provides us with the ra- dial velocity (Baranne et al. 1996). The corresponding values are displayed in the Journal of Observations (Appendix B), along with observational data such as the individual expo- sure times. The observations are graphically presented in Fig. 5a for WASP-53 and in Fig. 5b for WASP-81. Parameter α δ mB mV mR mr(cid:48) mI mJ mH mK Teff (K) log g(cid:63) (km s−1) ξt(km s−1) v sin i(cid:63) (km s−1) [Fe/H] [Ca/H] [Sc/H] [Ti/H] [V/H] [Cr/H] [Mn/H] [Co/H] [Ni/H] log A(Li) Mass (M(cid:12)) Radius (R(cid:12)) Spectral Type Distance (pc) WASP-53A 02h07(cid:48)38.22(cid:48)(cid:48) −20◦39(cid:48)43.0(cid:48)(cid:48) 17.46 ± 0.30 a 12.19 ± 0.30 b 11.85 ± 0.30 a 12.29 ± 0.30 c 11.653 ± 0.020 d 10.959 ± 0.026 e 10.474 ± 0.022 e 10.390 ± 0.023 e 4950 ± 60 4.40 ± 0.20 0.60 ± 0.25 < 2.7 ± 0.3 0.22 ± 0.11 0.16 ± 0.15 0.19 ± 0.11 0.26 ± 0.15 0.44 ± 0.20 0.22 ± 0.11 0.29 ± 0.20 0.24 ± 0.11 0.20 ± 0.12 < 0.32 ± 0.16 0.87 ± 0.08 0.96 ± 0.24 K3 235 ± 55 WASP-81A 20h16(cid:48)49.89(cid:48)(cid:48) +03◦17(cid:48)38.7(cid:48)(cid:48) 13.14 ± 0.30 a 12.29 ± 0.10 b 12.67 ± 0.30 a 12.36 ± 0.30 c 11.326 ± 0.053 b 11.263 ± 0.027 e 10.913 ± 0.024 e 10.892 ± 0.026 e 5890 ± 120 4.27 ± 0.09 0.94 ± 0.15 1.20 ± 0.73 −0.36 ± 0.14 −0.25 ± 0.09 −0.18 ± 0.18 −0.14 ± 0.11 −0.29 ± 0.12 −0.40 ± 0.11 −0.55 ± 0.09 −0.36 ± 0.14 −0.34 ± 0.15 1.21 ± 0.10 1.04 ± 0.09 1.24 ± 0.15 G1 410 ± 70 Note: Mass and Radius estimate using the Torres, Andersen & Gim´enez (2010) calibration. Spectral Type estimated from Teff using the table in Gray (2008). Abundances are relative to the solar values obtained by Asplund et al. (2009). references a) NOMAD; (Zacharias et al. 2004) b) TASS; (Droege et al. 2006) c) CMC14; ViZier I/304/out d) DENIS; (DENIS Con- sortium 2005) e) 2MASS; Skrutskie et al. (2006) 4.2.1 WASP-53 We first started to monitor WASP-53A with CORALIE, on 2010-12-05 targeting the orbital phase 0.75. The second spectrum was obtained 28 nights later, close to phase 0.25, and revealed a blue-shifted movement of nearly 400 m s−1, compatible with a planetary object. The star was immedi- ately flagged for intense follow-up, with the third spectrum acquired two nights after the second. The star was moving rapidly, and there was no sign of a change in the line width (FWHM), nor of its shape (span of the bisector slope), as can be visually inspected in Figs. 5 and 6. However, despite being obtained at phase 0.81, the star's velocity was puz- zling, being red-shifted by 300 m s−1 compared to the first epoch. Our strategy has always been to follow any radial- velocity movement and identify its origin, planetary or not. Observations were continued. Before the observing season was over, we had confirmed a radial-velocity oscillation at a period matching the WASP signal (caused by WASP-53b), plus a rapid rise in the ra- dial velocity. The following season we observed with both HARPS and CORALIE. WASP-53b's motion was quickly c(cid:13) 2014 RAS, MNRAS 000, 1–23 (a) WASP-53 (b) WASP-81 The WASP-53 and 81 systems 7 Figure 5. Radial velocities and models for WASP-53 (left) and WASP-81 (right). HARPS points are represented by discs, CORALIE is shown as triangles, pointed upwards prior to the upgrade and downwards for data acquired since. top: Radial velocity timeseries with the preferred two-planet model adjusted to the data. middle: same as on top, minus the inner planet. Ten alternative models, sampled randomly from the posterior are displayed in pink. Rossiter–McLaughlin sequences were removed from these plots; they can be found in Fig. 7. Residuals are shown below the main plots. Further down, in order, we have the variation in the slope of the bisector span, and the variation in the FWHM of the cross-correlation function. recovered and found to be in phase. Our monitoring con- tinued and we observed the velocity of the star rise until it plateaued, indicating that an additional, massive, highly eccentric object had just finished passing through perias- tron. Our observations are on-going and, to this day, the velocity of the star has yet to decrease to the level ob- served at our first spectrum. In total we present 98 spec- tra with CORALIE, including 25 since an upgrade that saw the installation of new, octagonal, fibres (Nov 2014), and of a Fabry-P´erot (Apr 2015) for the wavelength cal- ibration throughout the night. We also gathered 83 spec- tra with HARPS, which include three timeseries obtained c(cid:13) 2014 RAS, MNRAS 000, 1–23 during transit, in order to capture the Rossiter–McLaughlin effect. The CORALIE data were divided into two independent datasets in order to account for any offset between before and after the upgrade. The HARPS set was divided into four sets, one for each Rossiter–McLaughlin effect (plus the mea- surement obtained the night before and after transit), and one set containing all the rest of the data. We obtained sim- ilar results by analysing the HARPS data as one set, how- ever the division mitigates against any activity effect, such as done in Triaud et al. (2009). The Journal of Observations (Appendix B) is separated in several tables according to the various sub-samples of radial-velocities. 5500 6000 6500 7000 7500 8000−1000−500 0 500radial velocities (m/s)BJD − 2 450 000 . . . . 6000 6500 7000 7500−500 0 500 1000 1500radial velocities (m/s)BJD − 2 450 000 . . . .−800−600−400−200 0−50 0 50−200−100 0 100 200 5500 6000 6500 7000 7500 8000−200−100 0 100 200radial velocities (m/s)O − C (m/s)∆ bis (m/s)∆ fwhm (m/s)BJD − 2 450 000 . . . .−500 0 500 1000 1500−50 0 50−100 0 100 6000 6500 7000 7500−100 0 100radial velocities (m/s)O − C (m/s)∆ bis (m/s)∆ fwhm (m/s)BJD − 2 450 000 . . . . 8 Amaury H. M. J. Triaud et al. 4.2.2 WASP-81 CORALIE acquired our first spectrum on the WASP-81 system on 2011-09-29. The following night another was ob- served at the opposite phase. Its radial velocity revealed a variation and the target was placed in high priority. Within a month we had confirmed a variation at the WASP photo- metric period, and the star had set. The following season, we intended to monitor just enough to confirm that the os- cillation was still in phase and to start routine long-term monitoring, as in the case of WASP-47 (Neveu-VanMalle et al. 2016). The first measurement had a value clearly be- low expectations so we resumed an intense follow-up. As with WASP-53 we requested observing time on HARPS and monitored the system in parallel with CORALIE. On the third season, as we had predicted, the velocity reached a minimum and started rising. HARPS was the first instru- ment on sky for the fourth season. We had anticipated that the system would have returned to a similar velocity as in the first points. We were surprised to find that the star was nearly 2 km s−1 higher than in the previous season. Shortly after, velocities started to drop and the outer orbit closed earlier this year. In total 67 spectra were collected with CORALIE, 14 of which were after the upgrade. With HARPS we gathered 32 measurements, of which 19 were obtained during a single night as an attempt to detect the Rossiter–McLaughlin effect. As with WASP-53 we separate the CORALIE data into two sets, before and after upgrade. Since we do not detect the R–M effect, all of the HARPS points are analysed as part of the same set. Refer to the Journal of Observations (Appendix B) for further details. 5 GLOBAL ANALYSIS We analysed all of the photometric data and all of the radial velocity data together. We estimated the observed and phys- ical parameters of the system using an MCMC algorithm, as detailed in Gillon et al. (2012). Our modus operandi is similar with the difference described below. The photometric lightcurves were modelled using the formalism of Mandel & Agol (2002) for a transiting planet. The TRAPPIST and EulerCam lightcurves timestamps were transformed into BJD(TDB) from JD(UTC). The limb darkening was included in the form of a quadratic law, and its parameters were allowed to float, within the constraints of priors. The priors were computed by interpolating the data tabulated by Claret (2004), consistent with the stel- lar parameters of Table 2. In addition, on each photomet- ric timeseries, we allowed for a quadratic polynomial as a function of time in order to adjust for differential extinc- tion relative to the ensemble of comparison stars. After this treatment some lightcurves still contained a significant resid- ual scatter. For these we tried a number of other detrend- ing functions, selecting them on the basis of a reduction in the global Bayesian Information Criterion (BIC thereafter; Schwarz (1978)). Those comparisons were performed by sys- tematically using the same starting seed for the random ele- ments. Details about which functions were selected and ap- plied can be found in Appendix C. For those lightcurves where any new degree of complexity led to a worsened BIC, but where nevertheless the overall χ2 implied a poor fit, we scaled3. our error bars so as to approach a general re- duced χ2 r = 1. The uncertainty increase ensures that our confidence intervals are not under-estimated. Those correc- tion factors are available in Table C1. A similar approach is used for the radial-velocities, by quadratically adding a jitter term to some sequences. Two of the nine sequences required a jitter of order 1.5-2 m s−1. Individual detrended lightcurves and their transit model can be visually inspected in Fig. 3 & 4. Residuals and fully modelled lightcurves are presented in Appendix C. The radial velocities around the orbit were adjusted with two eccentric Keplerian functions (as in Hilditch (2001)), neglecting Newtonian effects. Since we cover less than an orbital period for WASP-53c and barely one for WASP-81c, we did not include a long term trend. Its inclu- sion leads to higher BIC values and its slope is mostly un- constrained. The Rossiter–McLaughlin effect was computed using the code written by Gim´enez (2006), following the formalism of Kopal (1942). For WASP-81b the effect was not detected and the final fit does not include it. For both systems we find the inner planets' orbits to be consistent with circular. Although allowing for eccentricity adds two parameters and increases the BIC, we nevertheless let these parameters float so as to include their uncertainties when marginalising the other parameters. √ v sin I(cid:63) cos β, The MCMC's jump parameters are mostly set to match observables, which are then converted to physical parame- ters to compute the relevant models. D is the transit depth, b the impact parameter, W the transit width. T0 is the mid- transit time and P the orbital period. We combine, the √ √ eccentricity e and the angle of periastron ω into the pair e sin ω which helps when exploring small eccen- e cos ω, tricities (Triaud et al. 2011). Similarly we also construct the pair v sin I(cid:63) sin β to model the Rossiter– McLaughlin effect. β is the projected spin–orbit angle, and v sin I(cid:63) the measure of the projected rotation velocity of the star, which in principle should match v sin i(cid:63) from the spec- tral analysis. Instead of the semi-amplitude K we use the 1 − e2P 1/3. This jump parameter K2 such that K2 = K helps reduce some correlation between the parameters (Ford 2006) helping with the exploration of parameter space. For similar reasons we combine the limb-darkening coefficients into c1 = 2u1 + u2, and c2 = u1 − 2u2 following the recom- mendation of Holman et al. (2006). In the case of two Euler lightcurves for WASP-53 we also fit a sine function through the data, with a period P and a T0. Additional subscript indicate which lightcurves those are for. √ √ All jump parameters are well constrained within one scale. Except for Pc and K2,c, we sample our parame- ters' posteriors using Gaussian priors, whose variance is set with an initial Gibbs sampler. As an improvement over Gillon et al. (2012) Pc and K2,c were sampled using non- informative priors in log space, otherwise known as Jeffrey priors. However their values remained largely within one er- ror bar, and the application of Gaussian steps does not lead to qualitatively different results. The metallicity [Fe/H], and 3 we realise that this approach is not self-consistent, but our method is sufficient for small re-adjustements of the uncertainties and unambiguous detections. c(cid:13) 2014 RAS, MNRAS 000, 1–23 (a) WASP-53b (b) WASP-81b The WASP-53 and 81 systems 9 Figure 6. Radial velocity measurements phased with the orbit of the inner planet, after subtracting the variation due to the outer object. The preferred model is drawn in thick red, with another ten alternate models randomly picked from the posterior shown in pink. Residuals are shown below the main plots. Further down we have the variation in the slope of the bisector span, and the variation in the FWHM of the cross-correlation function. The symbols are as in Fig. 5. (a) WASP-53b (b) WASP-81b Figure 7. As for Fig. 6 but zoomed around transit time to show the Rossiter–McLaughlin effect. c(cid:13) 2014 RAS, MNRAS 000, 1–23 −200 0 200−50 0 50−200−100 0 100 200−0.4−0.20.00.20.4−200−100 0 100 200vrad (m/s)O − C (m/s)∆ bis (m/s)∆ fwhm (m/s)φ . . . .−150−100−50 0 50 100 150−50 0 50−100 0 100−0.4−0.20.00.20.4−100 0 100vrad (m/s)O − C (m/s)∆ bis (m/s)∆ fwhm (m/s)φ . . . .−50 0 50−20 0 20−50 0 50−0.03−0.02−0.010.000.010.020.03−50 0 50vrad (m/s)O − C (m/s)∆ bis (m/s)∆ fwhm (m/s)φ . . . .−50 0 50−40−20 0 20 40−50 0 50−0.04−0.020.000.020.04−50 0 50vrad (m/s)O − C (m/s)∆ bis (m/s)∆ fwhm (m/s)φ . . . . 10 Amaury H. M. J. Triaud et al. Table 3. Median and 1σ confidence regions for the jump param- eters that evolve in our MCMC chains. Errors on the last two digits of each parameters are given in brackets. Parameters not dependent on information contained within the adjusted data are highlighted with asterisks Parameters (Units) WASP-53 WASP-81 the star [Fe/H] (dex) Teff (K) c1,JR c2,JR c1,BB c2,BB c1,I+z c2,I+z c1,r(cid:48) c2,r(cid:48) inner planet Db bb (R(cid:12)) Wb (d) T0,b (BJD -2 450 000) eb cos ωb Pb (d) √ √ eb sin ωb K2,b (ms−1d1/3) √ v sin I(cid:63) cos βb √ v sin I(cid:63) sin βb outer planet T0,c (BJD -2 450 000) ec cos ωc Pc (day) √ √ ec sin ωc K2,c (ms−1d1/3) 0.22(+0.11) (−0.11) 4953(+59) (−60) – – 1.069(+37) (−38) 0.031(+31) (−30) 0.995(+62) (−61) −0.020(+37) (−37) 1.316(+29) (−29) 0.272(+28) (−28) −0.36(+0.14) (−0.14) 5870(+120) (−120) 0.918(+33) (−33) −0.280(+23) (−23) 0.772(+57) (−57) −0.321(+37) (−38) 0.807(+55) (−57) −0.297(+30) (−32) – – * * * * * * * * * * 0.01831(+33) (−34) 0.562(+20) (−22) 0.09469(+65) (−64) 5943.56695(+11) (−12) 3.3098443(+20) (−20) −0.070(+24) (−21) −0.085(+46) (−33) 485.9(+2.7) (−2.7) 0.91(+0.11) (−0.13) −0.07(+0.19) (−0.19) 0.01254(+27) (−26) 0.15(+0.10) (−0.10) 0.14501(+73) (−64) 6195.57462(+20) (−20) 2.7164762(+23) (−23) −0.026(+83) (−76) −0.00(+0.12) (−0.12) 140.7(+4.7) (−4.6) – – 5456(+11) (−13) 2840(+170) (−130) −0.867(+12) (−12) −0.292(+30) (−30) 3685(+110) (−89) 6936.5(+2.6) (−2.5) 1297.2(+8.1) (−7.8) 0.5871(+40) (−42) −0.4609(+61) (−61) 10 557(+51) (−50) jump parameters: 23 21 effective temperature Teff complete this list of jump param- eters. They are controlled by priors obtained from Table 2 and used to compute at every MCMC step a stellar mass and a stellar radius produced in a fashion similar to Torres, Andersen & Gim´enez (2010). For our final analysis we set 10 chains of 100 000 steps, starting from different seeds. All converged to similar BIC values. The first 20 000 steps were systematically removed (to allow for burn-in), and the remainder were analysed leading to our results. For each of our 10 chains we extract the median value for each parameter and compare them to one another. They are usually of order 0.1% different from one another, except for the jump parameters responsible for modelling WASP-53c, which can vary as much as 60%. This is because the orbit is not closed. We discuss this further in the next section. Table 4. Final estimates for the median and 1σ confidence re- gions, for various interesting parameters of the WASP-53 and WASP-81 systems. They were estimated from the posteriors of the jump parameters outlined in Table 3. Errors on the last two digits of each parameters are given in brackets. Upper limits are for 3σ confidence. Parameters not dependent on information con- tained within the adjusted data are highlighted with asterisks Physical Parameters WASP-53 WASP-81 the star M(cid:63) (M(cid:12)) R(cid:63) (R(cid:12)) ρ(cid:63) (ρ(cid:12)) L(cid:63) (L(cid:12)) Teff (K) log g(cid:63) (cgs) [Fe/H] (dex) v sin I(cid:63) (km s−1) inner planet Pb (day) T0,b (BJD -2 450 000) Kb (m s−1) 0.839(+54) (−54) 0.798(+23) (−23) 1.648(+91) (−85) 0.344(+23) (−23) 4953(+60) (−60) 4.553(+19) (−20) 0.22(+0.11) (−0.11) 0.86(+0.21) (−0.21) 1.080(+59) (−58) 1.283(+40) (−37) 0.513(+30) (−37) 1.76(+0.20) (−0.18) 5870(+120) (−120) 4.258(+22) (−27) −0.36(+0.14) (−0.14) – * * * * 3.3098443(+20) (−20) 5943.56695(+11) (−12) 326.1(+1.8) (−1.8) 2.7164762(+23) (−23) 6195.57462(+23) (−20) 100.8(+3.4) (−3.3) Mb (MJup) Rb (RJup) ρb (ρJup) log gb (cgs) ab/R(cid:63) Tb,eq (K) ab (AU) ib (deg) βb (deg) eb outer planet Pc (day) T0,c (BJD -2 450 000) Kc (m s−1) Mc sin ic (MJup) ac (AU) ec ωc (deg) 2.132(+92) (−94) 1.074(+37) (−37) 1.72(+0.15) (−0.13) 3.680(+22) (−22) 11.05(+0.20) (−0.19) 1053(+16) (−16) 0.04101(+83) (−91) 87.08(+0.16) (−0.15) −4(+12) (−12) < 0.030 > 2840(+170) (−130) 5456(+11) (−13) > 475.6(+8.2) (−8.0) > 16.35(+0.85) (−0.82) > 3.73(+0.16) (−0.14) 0.8369(+69) (−70) 198.6(+2.0) (−2.0) 0.729(+36) (−35) 1.429(+51) (−46) 0.250(+23) (−23) 2.967(+27) (−30) 6.56(+0.13) (−0.16) 1623(+38) (−37) 0.03908(+70) (−72) 88.69(+0.88) (−0.92) – < 0.066 1297.2(+8.1) (−7.8) 6936.5(+2.6) (−2.5) 1169.3(+6.9) (−6.6) 56.6(+2.0) (−2.0) 2.426(+44) (−45) 0.5570(+44) (−44) 321.86(+0.52) (−0.51) The posterior probability distributions have been stored and can be requested by email to the lead author. We present the median values and 1σ region of our posteriors in Table 3 for the jump parameters, and in Table 4 for the physical pa- rameters. The results are discussed in the following section. c(cid:13) 2014 RAS, MNRAS 000, 1–23 The WASP-53 and 81 systems 11 Figure 8. Planetary apastron versus binary periastron (or separation if eccentricity is unknown) in astronomical units for known S-type planetary systems. The colour of the dots reflects the logarithm of the ratio of the planet-hosting star mass to the mass of its stellar companion(s) (white = 0.2, black = 18). WASP-53 and WASP-81 are highlighted as a blue and green diamond, respectively. The small black dots represent four systems containing a gas-giant and a brown dwarf. The dotted line is a 1:1 line, and the plain line is a 3:1 contour. Above that line systems are usually unstable (Dvorak 1986; Holman & Wiegert 1999). Data collected from openexoplanetcatalogue.com, from www.univie.ac.at/adg/schwarz/multiple.html, and from exoplanet.eu. 6 RESULTS 6.1 WASP-53 WASP-53 is a system composed of a star and two orbiting objects. WASP-53b is a hot Jupiter with a mass Mb = 2.1± 0.1MJup, with a radius Rb = 1.07 ± 0.04RJup. We find the inner orbit to be consistent with zero eccentricity, placing a 99% confidence limit of 0.03. The Rossiter–McLaughlin effect is weakly detected. We find a lower amplitude than we anticipated, likely because v sin i(cid:63) is over-estimated owing to an under-estimation of macroturbulence, as already noted for a number of late-type dwarfs (e.g. Triaud et al. (2011, 2015)). We find the spin–orbit angle β = −4◦ ± 12. The planet appears to be coplanar. The orbit of WASP-53c does not close within the span of our observations, which affects the various chains we launched. The median values on individual jump parame- ters values can vary by as much as 50%, which is reflected in the large errors in Table 3 and 4. This means that while we provide median values and their 1σ confidence ranges, those are in fact more akin to lower limits on Mc, ec, Pc etc. WASP-53c is at least 16MJup, with a period likely longer than 2500 days. The eccentricity of its orbit is high with our data being most consistent with 0.84 ± 0.01. We were lucky to observe the system during the final phases of WASP-53c's periastron passage (but unlucky to miss the first half). If removing the first season of CORALIE data, we only detect a quadratic drift with a weak curvature and would never have guessed the presence of such a massive companion within the system. 6.2 WASP-81 WASP-81 is a system composed of a star and two orbiting objects. WASP-81b is a hot Jupiter whose mass is Mb = c(cid:13) 2014 RAS, MNRAS 000, 1–23 Table 5. Dates on which WASP-53c and WASP-81c may transit. Numbers are calendar dates and Barycentric Julian Dates (BJD) – 2 450 000. WASP-53c WASP-81c passed dates – – 2616.0+170−130 (2002-12-07) 5455.5+11−13 (2010-09-16) future dates 8295.1+170−130 (2018-06-25) – 3044.9+24−24 (2004-02-09) 4342.1+16−16 (2007-08-29) 5639.3+8.5−8.2 (2011-03-18) 6936.5+2.6−2.5 (2014-10-06) 8233.8+8.5−8.2 (2018-04-25) 9531.0+16−16 (2021-11-12) 0.73±0.04MJup and radius Rb = 1.43±0.05RJup. Its orbit is consistent with being circular and we place a 99%-confidence upper limit at 0.07. The Rossiter–McLaughlin effect is not detected. Its amplitude is projected to be less than 10 m s−1. The low impact parameter means that the spin–orbit angle will be degenerate with v sin I(cid:63) as in Triaud et al. (2011). WASP-81c has a minimum mass Mc = 57 ± 2MJup, a period Pc = 1297± 8 days and an eccentricity of order 0.56. 7 DISCUSSION We have discovered two hot Jupiters orbiting the primary star of two tight binary systems. WASP-53b is super-Jupiter in mass, while WASP-81b is sub-Jupiter. Both occupy or- bits that are typical for hot Jupiters (eg. Santerne et al. 2016). Those two planets are both accompanied by brown- dwarf-mass objects, on highly eccentric orbits of a few AU. In Fig. 8, we plot other known planetary systems orbiting 10+010+110+210+310+40.010.11.010.0100.0binary periastron [AU]planet apastron [AU] . . . . 12 Amaury H. M. J. Triaud et al. (a) WASP-53 (b) WASP-81 Figure 9. MEGNO maps showing regions of stability in green, and chaos in red, showing where planets with the masses of WASP-53b and WASP-81b could exist. Any value in excess of 4 has the same colour. All regions on the left-hand side of the graphs are stable, and all regions on the right-hand side are unstable (within the outer orbit). Maps computed by integrating for 5 × 104 years (a) WASP-53 (b) WASP-81 Figure 10. MEGNO maps showing regions of stability in green, and chaos in red, for mass-less particles between the positions of the inner and outer objects in the WASP-53 and WASP-81 systems. Any value in excess of 4 has the same colour. Maps computed by integrating for 5 × 104 years one star of a multiple stellar system, showing how atypical WASP-53 and WASP-81 are within the current exoplanet population. Only four other systems have a stellar compan- ion with periastra closer than 10 AU: Kepler-444 (Dupuy et al. 2016), KOI-1257 (Santerne et al. 2014), HD 59686 (Or- tiz et al. 2016) and the astonishing, maybe retrograde, ν Oct (Ramm 2015) (the leftmost red dot, above the plain line). In addition, four gas giants have outer brown-dwarf com- panions: HAT-P-13 (Knutson et al. 2014), HIP 5158 (Feroz, Balan & Hobson 2011), HD 168443 (Sahlmann et al. 2011), and HD 38529 (Benedict et al. 2010), in architectures similar to WASP-53 and WASP-81. We now review the elements that make those systems stand out. We speculate about their origin, and propose some observational tests to verify some of our scenarios. We use the stability criterion numerically determined by Holman & Wiegert (1999) to compute the widest orbital separation that each of the hot Jupiters could have occupied. In the case of WASP-53 we obtain a critical semi-major axis acrit = 0.16± 0.15 AU, and for WASP-81, acrit = 0.38± 0.06 AU. This criterion was numerically determined for a mass ratio µ = m2/(m1 + m2) > 0.1, which is not satisfied for either of our systems, and likely explains the large uncer- tainty for WASP-53. We therefore proceed with a stability criterion devised by Petrovich (2015b) for hierarchical plan- etary system. We find that unstable orbit start emerging for 0.14 < ab < 0.17 for WASP-53, and 0.30 < ab < 0.38 for WASP-81, in good agreement with the previous approach. We further verify this with the method proposed by Cincotta, Giordano & Sim´o (2003), which uses a marker called the MEGNO (the Mean Exponential Growth fac- tor of Nearby Orbits), as implemented in rebound by Rein & Tamayo (2015). The MEGNO is a good tracker of or- bital chaos, meaning that infinitesimal changes in initial pa- c(cid:13) 2014 RAS, MNRAS 000, 1–23 rameters lead to diverging solutions. For quasi-periodic or- bits, thus those showing no chaotic behaviour, the MEGNO reaches a value of 2 (Hinse et al. 2010). Larger values of the MEGNO, typically > 4, indicate significant changes in the orbital parameters, a sign of chaos. Although this does not necessarily translate by unstable orbits (e.g. Deck et al. 2012), it often tracks them as we saw above. For our case the MEGNO outlines where nearly closed orbits exist and therefore informs us on where any disc material may have been stable, or where additional planets may exist. We used the parameters provided in Tab. 4 for the outer companions, assumed coplanarity between the inner and outer orbits, and computed the MEGNO for a particle with a mass and separation for WASP-53b and WASP-81b. We integrated each system for 5 × 106 years and obtained values of 2.0062 and 1.9999 respectively, indicating stability. We expanded those simulation and explore the (a, e) param- eter space. To compute the maps presented in Fig. 9 we in- tegrated each of the pixels over 5 × 104 years. We observe from those results that wide regions of chaos exist, which is consistent with the work of Holman & Wiegert (1999) and Petrovich (2015b). Both WASP-53 and WASP-81 appear stable over long periods of time. In the case of WASP-53 we observe that only regions closer in than 0.15 AU retain stable orbits. For WASP-81 there is slightly more space with orbits within 0.3 AU being generally stable. We also investigated whether other planets could exist between objects b and c by adding a third massless par- ticle. The results are displayed in Fig. 10, which show a range of stable orbits (in green). While this suggests that other planets could still be identified within our systems, we think it unlikely since hot Jupiters are usually found to be isolated. Only one hot Jupiter is known to have other plane- tary companions within an astronomical unit (Becker et al. 2015; Neveu-VanMalle et al. 2016), and a recent analysis of the Kepler data shows that a lack of nearby companions is an important aspect that sets hot Jupiters apart from other gas giants (Huang, Wu & Triaud 2016). 7.1 Determination of the mass and orbital inclination of WASP-53c and WASP-81c Gaia (Perryman et al. 2001) is an ESA mission launched in 2013 currently scanning the sky and measuring stellar posi- tions with a precision of order 30 micro-arcseconds for stars brighter than optical magnitude 12 (de Bruijne 2012). Sev- eral studies have investigated Gaia's potential for detecting gas giants using astrometry (Casertano et al. 2008; Sozzetti et al. 2014; Perryman et al. 2014; Sahlmann, Triaud & Mar- tin 2015), with Neveu et al. (2012) looking into the combi- nation of radial-velocities with astrometric measurements. From Perryman et al. (2014), WASP-53c and WASP- 81c will produce an astrometric displacement of their host stars, α, defined as: (cid:16) Mp (cid:17)(cid:16) a (cid:17)(cid:18) d (cid:19)−1 α = M(cid:63) 1AU 1pc arcsec (1) The WASP-53 and 81 systems 13 √ Nobs = 70 astrometric measurements4 with typical uncer- tainties σ = 40µas (de Bruijne 2012) to be collected on our two targets. This translates to an astronometric signal-to- noise, where S/N = α Nobs/σ, in excess of 50 for both systems. If instead ic = 10◦, we obtain α = 1500, and 1700µas respectively5. The amplitude of the orbital motion alone should inform us of the mutual inclination between the inner and outer planet. For astrometric signal-to-noise values of > 20 the orbital inclination will typically be es- timated with a precision of < 10◦ (Sahlmann, Triaud & Martin 2015). 7.2 Transit timing variations Using rebound (Rein & Spiegel 2015) we integrated the system over a few orbital periods of the outer companion and recorded when transits of the inner planet happened. For WASP-53b we expect total transit-timing variations, caused by the perturbing effect of the outer companion, to be of order 35s. However, most of the variation happens dur- ing periastron, which means that during nearly 100 transit epochs we covered for WASP-53b, we expect no variation to be measurable). We expect a detectable signal to appear within the next two years if our solution for the outer com- panion is correct. After WASP-53c swings again via peri- astron, the ephemeris for WASP-53b will become offset by approximately 35s. This offset remains constant until the following periastron when it will offset again by the same amount. We repeated the procedure for WASP-81 and find a similar TTV behaviour for WASP-81b with offset of ap- proximately 30s compared to the value we produce here. 7.3 Estimating k2 Secular interactions between pairs of orbiting planets usually excites their orbital eccentricities. In the case of WASP-53 and WASP-81, the innermost planetary orbit will be affected by tidal forces which tend to damp eccentricity, while the outer, massive companion occupies a highly eccentric orbit which will excite the inner planet's eccentricity. Mardling (2007) investigated this secular problem and found that the inner planet will reach an equilibrium eccentricity, called a fixed-point, with a value dependent on the planet's internal density profile, and parametrised by k2, the tidal Love num- ber (Sterne 1939). The more mass that gets included into the core of the planet, the larger will be the fixed-point eccen- tricity (Batygin, Bodenheimer & Laughlin 2009). As such a measure of the eccentricity of the tidally damped, inner or- bit can yield the core-mass fraction of exoplanets. Recently, Buhler et al. (2016) investigated the case of the HAT-P-13 system that presents an architecture similar to WASP-53 and WASP-81, and managed to constrain the core mass of the transiting hot Jupiter to 11 M⊕ by measuring an eccen- tricity of 0.007 ± 0.001. We see here how our two systems compare. We estimated the inner eccentricity, first assuming WASP-53 and WASP-81 will move on the sky by α = 260, and 300 µas respectively, caused by their outer com- panions, assuming an orbital inclination ic = 90◦, which is the poorest scenario possible. We can expect of order 4 78 expected measurements for WASP-53, and 60 for WASP-81 according to the following tool: http://gaia.esac.esa.int/gost/ 5 we do not detect a secondary set of lines in either of our spectra. This is equivalent to a limit of ic > 2◦, and ic > 5◦ respectively. c(cid:13) 2014 RAS, MNRAS 000, 1–23 14 Amaury H. M. J. Triaud et al. no tidal damping, using eq. 36 in Mardling (2007), and found fixed-point eccentricities of 0.00053 for WASP-53b, and 0.0027 for WASP-81b (HAT-P-13's parameters yield 0.0087). Any tidal dissipation will reduce these values by an amount dependent on the internal composition. Similarly, any mutual inclination between the outer and inner orbits will reduce these values. Sadly, because the eccentricities we expect of WASP-53b and WASP-81b are so small, we find it unlikely that their internal composition will be determined soon. If however, the eccentricity is one day measured, we should expect apsidal alignment or anti-alignment between the inner and outer orbits. The Love number can also be extracted from the way a transit lightcurve gets affected by apsidal precession (Ragozzine & Wolf 2009). 7.4 Possible scenarios for the planets' formation and orbital evolution We here speculate about the sequence of events leading to systems like WASP-53 and WASP-81, and also to HAT-P- 13b (Bakos et al. 2009), HD 59686Ab (Ortiz et al. 2016) and others. The presence of both a planet and a brown-dwarf mass object (one on the planetary side, the other on the stellar side of the brown-dwarf desert; Grether & Lineweaver 2006; Sahlmann et al. 2011) within the same system possi- bly suggests that core accretion and gravitational collapse can both operate within the same disc environment. They may also be smaller fragments from the nebula that created WASP-53A and WASP-81A. Both WASP-53b and WASP-81b could initially have formed on circumbinary orbits, to be captured by the pri- mary following a dynamical instability. This sort of scenario has been investigated by Sutherland & Fabrycky (2016), and shown to be an unlikely outcome, which, when it does happen, favours capture of the planet by the secondary in- stead of the primary. Therefore, WASP-53b and WASP-81b most likely formed within a disc surrounding WASP-53A and WASP-81A. We can think of two alternate scenarios and make an ap- peal for theorists to investigate them since they could teach us much about how gas giants form. WASP-53c and WASP-81c are too massive compared to the protoplanetary disc to have followed a type-II mi- gration (Duffell et al. 2014; Durmann & Kley 2015). If no or little orbital evolution has happened within the two sys- tems that we studied, then WASP-53b and WASP-81b must have formed well within the snow line, on orbits shorter than 0.15 and 0.3 AU respectively. Theoretical work by Batygin, Bodenheimer & Laughlin (2015) and Lee & Chiang (2015) suggests that the formation of gas giants within one astro- nomical unit is feasible. According to Fung, Shi & Chiang (2014) gas can flow through the gap carved by a planet, par- ticularly so if it reaches masses near that of a brown dwarf. The more massive the object, the more perturbations are produced at the outer gap's edge, launching streams of gas that replenish the inner disc to provide enough mass to al- low the formation of a gas giant. The large eccentricities of WASP-53c and WASP-81c might have enhanced this ef- fect. However, Lambrechts, Johansen & Morbidelli (2014) and Rosotti et al. (2016b) find that planets more massive than 20–30 Earth masses prevent the flow of dust grains across the same gap. Accordingly, if WASP-53b and WASP- 81b formed in-situ via core-accretion, they could only have used solids within about 0.2 AU, before being able to ac- crete gas. If this scenario is correct, systems like WASP-53 and WASP-81 can inform us about the efficiency of core ac- cretion, as well the minimum core mass necessary to accrete significant gas envelopes. This would also leave the planets poorer in metals than otherwise, something which is pos- sible to determine via transmission spectroscopy (Seager & Deming 2010; Madhusudhan et al. 2014) WASP-53c and WASP-81c are both eccentric. It has been argued that disc-planet interactions can excite the ec- centricity of gap-opening planets, but not to the values that we observe for WASP-53c and WASP-81c (e.g. Goldreich & Sari 2003; D'Angelo, Lubow & Bate 2006; Rosotti et al. 2016a; Teyssandier & Ogilvie 2016). This might imply that WASP-53c and WASP-81c reached their current orbital pa- rameters after disc dispersal, possibly due to dynamical in- teractions with third, yet unseen companions to WASP-53 and WASP-81. If this is the case then either WASP-53b and WASP-81b disc-migrated well before WASP-53c and WASP- 81c reached their current orbits, or, WASP-53b and WASP- 81b reached their current orbit following a high-eccentricity migration produced by the same instability that left their outer companions on eccentric orbital paths. In either of those cases, we expect a significant mutual inclination be- tween the inner and outer orbits, which Gaia should in prin- ciple be able to measure. Coplanarity would favour the sce- nario outlined in the previous paragraph. 8 CONCLUDING WORDS WASP-53 & WASP-81 are peculiar systems composed of both a planet and a brown dwarf. This orbital set-up is a relic of its past formation. Investigating them in further studies, notably with the help of Gaia, will prove invalu- able for understanding planet formation and the subsequent orbital evolution, but also the relation between planet for- mation, brown dwarf formation and stellar formation. NOTA BENE Dates are given in the BJD-TDB standard. The radii we used for Jupiter and the Sun are the volumetric mean radii. For clarity, we used the subscripts (cid:63) for the star, b for the inner planet, and c for the outer object, all throughout. ACKNOWLEDGMENTS AT would like to acknowledge inspiring discussions with the following people: Rosemary Mardling, Hanno Rein, Simon Hodgkin, Daniel Tamayo, Richard Booth, Cristobal Petro- vich, Yanqin Wu, Kaitlin Kratter, Jean Teyssandier, Ari Sil- burt and David Martin. TRAPPIST is a project funded by the Belgian Fund for Scientific Research (Fond National de la Recherche Scien- tifique, F.R.S-FNRS) under grant FRFC 2.5.594.09.F, with the participation of the Swiss National Science Fundation (SNF). The Euler Swiss telescope is supported by the Swiss National Science Foundation. We are all very grateful to ESO and its La Silla staff for their continuous support. c(cid:13) 2014 RAS, MNRAS 000, 1–23 WASP-South is hosted by the South African Astronomical Observatory while SuperWASP-North is hosted by the Issac Newton Group at the Observatorio del Roque de los Mucha- chos on La Palma. We are grateful for the ongoing support and assistance of these observatories. M. Gillon and E. Jehin are Research Associates at the F.R.S-FNRS; L. Delrez received the support of the F.R.I.A. fund of the FNRS. This publication makes use of data products from 2MASS, whose data was obtained through Simbad and VizieR services hosted at the CDS-Strasbourg. The Two Micron All Sky Survey is a joint project of the University of Massachusetts and the Infrared Processing and Analy- sis Center/California Institute of Technology, funded by the National Aeronautics and Space Administration and the Na- tional Science Foundation. References to exoplanetary systems were obtained through the use of the paper repositories, ADS and arXiv, but also through frequent visits to the exoplanet.eu (Schnei- der et al. 2011) and exoplanets.org (Wright et al. 2011) web- sites. REFERENCES Albrecht S., Winn J. N., Butler R. P., Crane J. D., Shect- man S. A., Thompson I. B., Hirano T., Wittenmyer R. A., 2012a, ApJ, 744, 189 Albrecht S. et al., 2012b, ApJ, 757, 18 Alibert Y., Mousis O., Mordasini C., Benz W., 2005, ApJL, 626, L57 Anderson D. R. et al., 2010, ApJ, 709, 159 -, 2015, ApJL, 800, L9 Asplund M., Grevesse N., Sauval A. J., Scott P., 2009, ARA&A, 47, 481 Bakos G. ´A. et al., 2009, ApJ, 707, 446 Baranne A. et al., 1996, A&AS, 119, 373 Baruteau C. et al., 2014, Protostars and Planets VI, 667 Batygin K., 2012, Nature, 491, 418 Batygin K., Bodenheimer P., Laughlin G., 2009, ApJL, 704, L49 Batygin K., Bodenheimer P. H., Laughlin G. P., 2015, ArXiv e-prints The WASP-53 and 81 systems 15 D'Angelo G., Lubow S. H., Bate M. R., 2006, ApJ, 652, 1698 Dawson R. I., 2014, ApJL, 790, L31 de Bruijne J. H. J., 2012, Ap&SS, 341, 31 Deck K. M., Holman M. J., Agol E., Carter J. A., Lissauer J. J., Ragozzine D., Winn J. N., 2012, ApJL, 755, L21 Delrez L. et al., 2014, A&A, 563, A143 DENIS Consortium, 2005, VizieR Online Data Catalog, 2263, 0 Doyle A. P., Davies G. R., Smalley B., Chaplin W. J., Elsworth Y., 2014, MNRAS, 444, 3592 Doyle A. P. et al., 2013, MNRAS, 428, 3164 Droege T. F., Richmond M. W., Sallman M. P., Creager R. P., 2006, PASP, 118, 1666 Duffell P. C., Haiman Z., MacFadyen A. I., D'Orazio D. J., Farris B. D., 2014, ApJL, 792, L10 Dupuy T. J., Kratter K. M., Kraus A. L., Isaacson H., Mann A. W., Ireland M. J., Howard A. W., Huber D., 2016, ApJ, 817, 80 Durmann C., Kley W., 2015, A&A, 574, A52 Dvorak R., 1986, A&A, 167, 379 Eastman J., Siverd R., Gaudi B. S., 2010, PASP, 122, 935 Feroz F., Balan S. T., Hobson M. P., 2011, MNRAS, 416, L104 Ford E. B., 2006, ApJ, 642, 505 Fung J., Shi J.-M., Chiang E., 2014, ApJ, 782, 88 Gaudi B. S., Winn J. N., 2007, ApJ, 655, 550 Gillon M. et al., 2013, A&A, 552, A82 Gillon M., Jehin E., Magain P., Chantry V., Hutsem´ekers D., Manfroid J., Queloz D., Udry S., 2011, in European Physical Journal Web of Conferences, Vol. 11, European Physical Journal Web of Conferences, p. 6002 Gillon M. et al., 2012, A&A, 542, A4 Gim´enez A., 2006, ApJ, 650, 408 Goldreich P., Sari R., 2003, ApJ, 585, 1024 Gray D. F., 2008, The Observation and Analysis of Stellar Photospheres, Gray, D. F., ed. Grether D., Lineweaver C. H., 2006, ApJ, 640, 1051 Guillochon J., Ramirez-Ruiz E., Lin D., 2011, ApJ, 732, 74 Hatzes A. P., Wuchterl G., 2005, Nature, 436, 182 H´ebrard G. et al., 2008, A&A, 488, 763 Helled R. et al., 2014, Protostars and Planets VI, 643 Hilditch R. W., 2001, An Introduction to Close Binary Stars, Hilditch, R. W., ed. Becker J. C., Vanderburg A., Adams F. C., Rappaport Hinse T. C., Christou A. A., Alvarellos J. L. A., S. A., Schwengeler H. M., 2015, ApJL, 812, L18 Go´zdziewski K., 2010, MNRAS, 404, 837 Benedict G. F., McArthur B. E., Bean J. L., Barnes R., Harrison T. E., Hatzes A., Martioli E., Nelan E. P., 2010, AJ, 139, 1844 Bodenheimer P., Hubickyj O., Lissauer J. J., 2000, Icarus, 143, 2 Buhler P. B., Knutson H. A., Batygin K., Fulton B. J., Holman M. J., Wiegert P. A., 1999, AJ, 117, 621 Holman M. J. et al., 2006, ApJ, 652, 1715 Huang C., Wu Y., Triaud A. H. M. J., 2016, ApJ, 825, 98 Jehin E. et al., 2011, The Messenger, 145, 2 Knutson H. A. et al., 2014, ApJ, 785, 126 Kopal Z., 1942, Proceedings of the National Academy of Fortney J. J., Burrows A., Wong I., 2016, ApJ, 821, 26 Science, 28, 133 Casertano S. et al., 2008, A&A, 482, 699 C´ebron D., Moutou C., Le Bars M., Le Gal P., Far`es R., 2011, in European Physical Journal Web of Conferences, Vol. 11, European Physical Journal Web of Conferences, p. 3003 Lai D., 2014, MNRAS, 440, 3532 Lai D., Foucart F., Lin D. N. C., 2011, MNRAS, 412, 2790 Lambrechts M., Johansen A., Morbidelli A., 2014, A&A, 572, A35 Lang D., Hogg D. W., Mierle K., Blanton M., Roweis S., Cincotta P. M., Giordano C. M., Sim´o C., 2003, Physica D 2010, AJ, 139, 1782 Nonlinear Phenomena, 182, 151 Claret A., 2004, A&A, 428, 1001 Collier Cameron A. et al., 2007, MNRAS, 380, 1230 Lee E. J., Chiang E., 2015, ApJ, 811, 41 Lendl M. et al., 2012, A&A, 544, A72 -, 2014, A&A, 568, A81 c(cid:13) 2014 RAS, MNRAS 000, 1–23 16 Amaury H. M. J. Triaud et al. Lin D. N. C., Bodenheimer P., Richardson D. C., 1996, Nature, 380, 606 L´opez-Morales M. et al., 2014, ApJL, 792, L31 Lovis C. et al., 2006, Nature, 441, 305 Madhusudhan N., Crouzet N., McCullough P. R., Deming D., Hedges C., 2014, ApJL, 791, L9 Mandel K., Agol E., 2002, ApJL, 580, L171 Mardling R. A., 2007, MNRAS, 382, 1768 Marigo P., Girardi L., Bressan A., Groenewegen M. A. T., Silva L., Granato G. L., 2008, A&A, 482, 883 Marmier M. et al., 2013, A&A, 551, A90 Mayor M., Queloz D., 1995, Nature, 378, 355 Naoz S., Farr W. M., Lithwick Y., Rasio F. A., Teyssandier J., 2011, Nature, 473, 187 Neveu M., Sahlmann J., Queloz D., S´egransan D., 2012, in Orbital Couples: Pas de Deux in the Solar System and the Milky Way, Arenou F., Hestroffer D., eds., pp. 81–85 Neveu-VanMalle M. et al., 2016, A&A, 586, A93 Ortiz M. et al., 2016, ArXiv e-prints Perryman M., Hartman J., Bakos G. ´A., Lindegren L., 2014, ApJ, 797, 14 Perryman M. A. C. et al., 2001, A&A, 369, 339 Petrovich C., 2015a, ApJ, 799, 27 -, 2015b, ApJ, 808, 120 Pollacco D. L. et al., 2006, PASP, 118, 1407 Pollack J. B., Hubickyj O., Bodenheimer P., Lissauer J. J., Podolak M., Greenzweig Y., 1996, Icarus, 124, 62 Queloz D., Eggenberger A., Mayor M., Perrier C., Beuzit J. L., Naef D., Sivan J. P., Udry S., 2000, A&A, 359, L13 Ragozzine D., Wolf A. S., 2009, ApJ, 698, 1778 Ramm D. J., 2015, MNRAS, 449, 4428 Rasio F. A., Ford E. B., 1996, Science, 274, 954 Rein H., Spiegel D. S., 2015, MNRAS, 446, 1424 Rein H., Tamayo D., 2015, MNRAS, 452, 376 Rogers T. M., Lin D. N. C., McElwaine J. N., Lau H. H. B., 2013, ApJ, 772, 21 Rosotti G. P., Booth R. A., Clarke C. J., Teyssandier J., Facchini S., Mustill A. J., 2016a, ArXiv e-prints Rosotti G. P., Juhasz A., Booth R. A., Clarke C. J., 2016b, MNRAS, 459, 2790 Sahlmann J. et al., 2011, A&A, 525, A95 Sahlmann J., Triaud A. H. M. J., Martin D. V., 2015, MN- RAS, 447, 287 Santerne A. et al., 2014, A&A, 571, A37 -, 2016, A&A, 587, A64 Schlaufman K. C., 2010, ApJ, 719, 602 Schneider J., Dedieu C., Le Sidaner P., Savalle R., Zolo- tukhin I., 2011, A&A, 532, A79 Schwarz G., 1978, Annals of Statistics, 6, 461 Seager S., Deming D., 2010, ARA&A, 48, 631 Sestito P., Randich S., 2005, A&A, 442, 615 Skrutskie M. F. et al., 2006, AJ, 131, 1163 Southworth J. et al., 2009, MNRAS, 396, 1023 Sozzetti A., Giacobbe P., Lattanzi M. G., Micela G., Mor- bidelli R., Tinetti G., 2014, MNRAS, 437, 497 Spalding C., Batygin K., 2015, ArXiv e-prints Sterne T. E., 1939, MNRAS, 99, 451 Stetson P. B., 1987, PASP, 99, 191 Sutherland A. P., Fabrycky D. C., 2016, ApJ, 818, 6 Teyssandier J., Ogilvie G. I., 2016, MNRAS, 458, 3221 Thies I., Kroupa P., Goodwin S. P., Stamatellos D., Whit- worth A. P., 2011, MNRAS, 417, 1817 Torres G., Andersen J., Gim´enez A., 2010, A&ApR, 18, 67 Tregloan-Reed J., Southworth J., 2013, MNRAS, 431, 966 Triaud A. H. M. J., 2011, A&A, 534, L6 Triaud A. H. M. J. et al., 2010, A&A, 524, A25 -, 2015, MNRAS, 450, 2279 -, 2009, A&A, 506, 377 -, 2011, A&A, 531, A24 Ward W. R., 1997, Icarus, 126, 261 Winn J. N., Fabrycky D. C., 2015, ARA&A, 53, 409 Winn J. N., Johnson J. A., Albrecht S., Howard A. W., Marcy G. W., Crossfield I. J., Holman M. J., 2009, ApJL, 703, L99 Wright J. T. et al., 2011, PASP, 123, 412 Wu Y., Murray N. W., Ramsahai J. M., 2007, ApJ, 670, 820 Zacharias N., Monet D. G., Levine S. E., Urban S. E., Gaume R., Wycoff G. L., 2004, in Bulletin of the Ameri- can Astronomical Society, Vol. 36, American Astronomi- cal Society Meeting Abstracts, p. 1418 APPENDIX A: JOURNAL OF OBSERVATIONS c(cid:13) 2014 RAS, MNRAS 000, 1–23 Table A1: Radial velocities of WASP-53 obtained with CORALIE before its recent upgrade. BJD is the barycentric Julian date – 2 450 000 days. Vrad is the radial velocity obtained by fitting a cross-correlation function with a Gaussian, σRV is the error on Vrad. FWHM is the full width at half maximum of the cross-correlation function, and contrast is the amplitude. The WASP-53 and 81 systems 17 Vrad BJD (days) (km s−1) -4.38964 -4.75981 -4.10113 -4.53580 -4.45769 -3.81514 -3.85825 -4.43614 -3.86707 -4.37790 -4.19644 -4.32590 -4.35247 -3.80772 -4.11077 -3.81662 -3.69384 -3.50787 -3.88919 -3.68608 -3.89564 -3.81609 -3.49432 -3.62453 -3.61455 -4.11173 -3.93644 -3.49648 -4.07174 -3.71694 -3.46744 -4.01495 -3.46497 -3.84547 -3.44630 -4.02707 -3.49841 -3.68569 -3.54201 -3.87751 -3.99424 -3.81564 -4.11729 -3.50543 -3.81243 -3.46799 -3.71479 -4.09012 -4.06430 -3.84938 -3.61756 -3.93777 -4.08678 -4.05183 55535.622498 55563.633734 55565.620787 55583.571236 55586.577359 55588.564415 55591.537016 55593.534932 55595.533798 55596.534633 55599.531272 55600.531357 55613.530286 55614.520174 55619.515487 55624.511532 55769.931969 55770.811670 55772.921899 55777.849432 55782.883847 55802.854702 55806.876427 55809.812708 55810.864649 55811.834443 55825.779642 55826.823506 55834.845196 55852.655157 55869.776795 55887.725010 55889.585568 55910.593618 55952.575538 55974.534361 56101.929818 56108.902853 56130.929240 56158.920126 56165.794126 56166.840915 56182.742965 56184.735109 56186.764241 56190.675191 56196.864653 56235.613045 56245.697375 56264.722841 56309.616625 56335.551640 56460.921265 56490.864375 Table continues next page... σRV (km s−1) 0.01414 0.01675 0.02222 0.01833 0.01936 0.02958 0.01594 0.02059 0.02407 0.01454 0.02246 0.02605 0.03126 0.03751 0.03829 0.04668 0.03211 0.01869 0.02320 0.01545 0.01539 0.01659 0.01352 0.01453 0.01691 0.01961 0.01342 0.02008 0.01455 0.01847 0.01511 0.01502 0.01504 0.01446 0.01294 0.02632 0.02543 0.03173 0.02272 0.02491 0.02180 0.01755 0.01762 0.01419 0.01735 0.02295 0.01309 0.02545 0.01518 0.01631 0.01619 0.02599 0.01861 0.02101 (%) FWHM contrast (km s−1) 7.76355 7.83813 7.85708 7.76400 7.88629 7.88939 7.85784 7.88560 7.86921 7.89625 7.82083 7.81524 7.75947 7.98353 7.86610 7.99027 7.83406 7.84814 7.76734 7.81802 7.75827 7.79265 7.77050 7.85935 7.82177 7.82148 7.83048 7.81726 7.82559 7.89498 7.84842 7.76766 7.78976 7.80091 7.90378 7.83726 7.87117 7.87649 7.82397 7.83489 7.84562 7.86863 7.83280 7.81771 7.89344 7.92760 7.83567 7.83788 7.90888 7.88340 7.91453 8.01659 7.87168 7.86144 37.811 38.764 29.332 37.735 36.602 36.377 37.219 36.521 36.778 37.250 35.628 34.950 35.915 12.241 12.606 31.484 40.671 40.222 20.208 39.166 40.337 40.844 40.390 40.282 40.031 39.929 39.001 39.820 40.420 39.010 39.370 39.004 38.829 39.026 38.024 38.734 40.333 39.586 39.617 39.943 35.852 39.564 38.786 39.385 39.090 37.779 38.860 37.690 37.534 39.189 40.096 39.469 43.357 40.429 slope bisector span (km s−1) 0.00572 -0.01915 0.06569 -0.01908 0.03945 0.10109 0.01104 0.01531 0.01591 0.02008 0.04899 0.02073 0.02255 0.08784 0.07253 0.16010 -0.01480 0.05379 0.03801 -0.01838 0.05484 0.00480 -0.00408 -0.03700 0.00585 -0.00983 -0.03320 -0.10400 0.02738 -0.02472 -0.03676 0.02644 -0.01253 -0.05197 0.02529 -0.02393 -0.02903 0.02577 -0.10018 -0.04714 -0.00917 -0.03721 0.04594 -0.00971 0.05377 0.04893 -0.02100 0.01474 0.00491 0.03949 -0.03900 0.01497 -0.00289 -0.05535 exposure (sec) 1800.679 1800.736 1800.743 1800.754 1800.755 1800.756 1800.759 1800.758 1800.859 1800.739 1800.775 1800.736 1800.830 1800.768 1800.726 1800.767 1800.816 1800.797 1800.748 1800.821 1800.740 1800.739 1800.777 1800.739 1800.798 1800.777 1800.718 1800.818 1800.758 1800.759 1800.736 1800.754 1800.755 1800.778 1800.688 1800.748 1800.799 1800.940 1800.817 1800.781 1800.759 1800.777 1800.798 1800.758 1800.778 1800.820 2700.667 1800.777 1800.756 1800.749 1800.739 1800.758 1800.756 1800.755 c(cid:13) 2014 RAS, MNRAS 000, 1–23 18 Amaury H. M. J. Triaud et al. JDB (days) 56505.803479 56516.868634 56538.754792 56543.888666 56544.693786 56545.715662 56562.781196 56563.802018 56565.713332 56567.702293 56591.740189 56603.739663 56644.683453 56682.557464 56817.930262 56877.826276 56878.777852 56920.870295 56961.797598 Vrad (km s−1) -3.53226 -4.11013 -3.46848 -3.98696 -3.50813 -3.63146 -3.94366 -3.97705 -3.68586 -3.59801 -3.50574 -3.82508 -3.51463 -4.10162 -4.05370 -4.07700 -3.63053 -4.12036 -3.54257 σRV (km s−1) 0.01677 0.02259 0.02937 0.01538 0.03240 0.01620 0.01417 0.01494 0.01541 0.02820 0.01941 0.01907 0.01493 0.01957 0.02181 0.01910 0.03452 0.01832 0.01781 (%) FWHM contrast (km s−1) 7.88179 7.86987 7.87498 7.87143 7.85548 7.94485 7.82567 7.91152 7.84101 7.86619 7.79994 7.84847 7.76010 7.85501 7.86074 7.80224 7.88373 7.83603 7.78155 41.428 41.506 40.897 39.824 41.040 37.812 40.715 40.932 40.339 41.893 39.462 40.983 38.600 40.062 41.914 38.824 34.977 39.391 38.574 slope bisector span (km s−1) 0.04581 -0.04888 0.01746 0.02301 0.06695 -0.00368 -0.00693 0.00453 0.00271 0.03071 0.00814 0.01256 -0.01970 -0.00216 0.01356 0.01261 0.04216 -0.03675 0.05712 exposure (sec) 1800.799 1800.857 1800.779 1800.763 1800.802 1800.819 1800.786 1800.836 1800.777 1800.775 1800.719 1800.779 1800.778 1800.770 1602.786 1800.058 1800.685 1800.677 1800.601 Table A2: Radial velocities of WASP-53 obtained with CORALIE after its recent upgrade. BJD is the barycentric Julian date – 2 450 000 days. Vrad is the radial velocity obtained by fitting a cross-correlation function with a Gaussian, σRV is the error on Vrad. FWHM is the full width at half maximum of the cross-correlation function, and contrast is the amplitude. BJD (days) 56989.740666 57001.640882 57003.640065 57004.682396 57012.592977 57027.631091 57063.528866 57065.534295 57192.938473 57205.925822 57261.823294 57341.740033 57362.678177 57367.607494 57381.547386 57389.610441 57413.581920 57587.836041 57616.796911 57681.574518 57689.706629 57691.675531 57711.743027 57713.676572 57713.698411 Vrad (km s−1) -3.92092 -3.53647 -4.11709 -3.59763 -3.75798 -3.72146 -4.02456 -3.74442 -3.84251 -3.96730 -4.14072 -3.88956 -3.49435 -4.16776 -3.92186 -3.61683 -4.09479 -3.52915 -3.78390 -4.02549 -3.71986 -4.13251 -4.18272 -3.56151 -3.54916 σRV (km s−1) 0.02509 0.02300 0.01937 0.02263 0.02566 0.03574 0.05181 0.03302 0.03767 0.04649 0.04184 0.05660 0.01850 0.02672 0.02486 0.02988 0.03288 0.02814 0.02024 0.03158 0.02108 0.01765 0.03598 0.02072 0.02373 (%) FWHM contrast (km s−1) 7.82426 7.73609 7.71262 7.74968 7.80047 7.68990 7.77355 7.76903 7.74347 7.82306 7.70847 7.82722 7.75263 7.72596 7.66936 7.76154 7.77853 7.79372 7.67173 7.69350 7.72289 7.72530 7.68311 7.80813 7.77659 42.096 42.361 41.800 41.276 42.030 38.087 41.187 41.060 45.120 42.941 43.479 43.813 43.012 43.081 41.940 43.413 41.771 41.294 42.253 43.049 42.465 42.658 42.736 42.301 42.830 slope bisector span (km s−1) -0.11250 0.00145 -0.01624 -0.00541 -0.03335 0.04302 -0.02410 -0.00757 0.01185 0.03207 0.00495 -0.08024 -0.00904 0.00961 -0.00042 -0.02558 0.04996 -0.06745 -0.03829 0.03337 0.02337 -0.01465 -0.03452 -0.01114 -0.04693 exposure (sec) 1800.683 1800.780 1800.075 1800.932 1800.233 1800.770 1800.102 1800.943 1800.861 1800.925 1800.924 1800.895 2700.715 2700.804 2700.686 1800.083 2700.626 1800.382 1800.431 1800.382 1800.354 1800.394 1800.352 1800.353 1800.372 c(cid:13) 2014 RAS, MNRAS 000, 1–23 Table A3: Radial velocities of WASP-53 obtained with HARPS. BJD is the barycentric Julian date – 2 450 000 days. Vrad is the radial velocity obtained by fitting a cross-correlation function with a Gaussian, σRV is the error on Vrad. FWHM is the full width at half maximum of the cross-correlation function, and contrast is the amplitude. The WASP-53 and 81 systems 19 Vrad BJD (days) (km s−1) -3.87536 -3.77124 -3.46797 -3.49415 -4.00995 -3.90801 -3.47586 -3.44345 -3.74232 -3.74253 -3.75174 -3.73814 -3.74724 -3.77343 -3.78034 -3.78879 -3.79623 -3.81542 -3.80869 -3.80569 -3.81191 -3.82316 -3.83545 -3.84303 -3.84995 -3.84996 -4.09963 -4.04309 -4.07244 -4.09707 -3.67055 -4.08058 -3.82192 -4.04984 -3.45497 -3.70159 -3.70009 -3.69652 -3.72787 -3.72226 -3.72504 -3.72870 -3.74305 -3.74230 -3.74590 -3.76612 -3.77051 -3.79008 -3.78008 -3.78198 -3.79791 -3.80101 -3.80062 -3.82351 55802.804248 55802.902737 55803.735877 55803.876858 55825.652716 55825.842017 55826.656742 55826.837326 55827.646391 55827.656368 55827.669238 55827.680986 55827.692515 55827.703846 55827.715374 55827.726810 55827.738546 55827.749681 55827.761637 55827.773177 55827.784612 55827.796036 55827.807576 55827.819324 55827.830760 55827.842091 55828.639175 55828.880860 55831.640517 55831.862305 56108.917970 56109.931976 56158.807248 56159.819284 56190.816064 56191.702180 56191.713245 56191.724577 56191.736348 56191.747992 56191.759624 56191.771060 56191.782391 56191.794012 56191.805748 56191.817172 56191.828724 56191.840043 56191.851583 56191.863100 56191.874628 56191.886168 56191.897592 56191.909027 Table continues next page... σRV (km s−1) 0.01428 0.00615 0.00592 0.01413 0.00918 0.00695 0.00682 0.00407 0.01085 0.00678 0.00631 0.00634 0.00573 0.00561 0.00546 0.00536 0.00512 0.00512 0.00518 0.00438 0.00451 0.00414 0.00448 0.00454 0.00553 0.00523 0.00613 0.00432 0.00490 0.00367 0.01289 0.00352 0.00781 0.01179 0.00746 0.00717 0.00584 0.00596 0.00617 0.00600 0.00565 0.00555 0.00534 0.00596 0.00535 0.00464 0.00455 0.00529 0.00543 0.00526 0.00667 0.00761 0.00746 0.00717 (%) FWHM contrast (km s−1) 6.35185 6.44231 6.43227 6.44454 6.43519 6.40550 6.42903 6.45292 6.42916 6.46451 6.49718 6.46326 6.44915 6.42803 6.47191 6.45148 6.43309 6.43640 6.44554 6.43072 6.43123 6.44297 6.44773 6.43978 6.47049 6.47341 6.47747 6.47502 6.49495 6.44317 6.48511 6.42686 6.45609 6.43642 6.40310 6.40858 6.40354 6.39529 6.41546 6.41604 6.37832 6.39147 6.39205 6.39070 6.39200 6.40073 6.39737 6.40980 6.41797 6.40679 6.40276 6.47274 6.45821 6.48428 48.819 49.068 49.138 48.445 50.675 50.365 50.419 49.300 48.557 48.782 48.731 48.738 49.065 49.040 48.934 49.091 49.178 49.235 49.108 49.362 49.319 49.391 49.167 49.196 48.899 48.922 48.749 49.127 48.903 49.351 48.048 49.455 48.928 49.040 49.253 49.055 49.321 49.155 49.184 49.269 49.470 49.478 49.326 49.423 49.398 49.412 49.586 49.416 49.150 49.445 49.074 48.383 48.542 47.642 slope bisector span (km s−1) 0.01421 0.01648 0.02630 0.03957 -0.00062 0.03042 0.03065 0.02401 0.03541 0.01240 0.01620 0.05819 0.00441 0.01460 0.01938 0.00973 0.00681 0.01215 0.02622 0.02610 0.01551 0.03097 0.01898 0.03451 0.00391 0.00081 0.01814 0.01328 0.00429 0.03159 0.01989 0.01035 0.02814 0.01204 0.02240 0.05467 0.02801 0.05629 0.01366 0.01401 0.04169 0.02959 0.01126 0.03308 0.01416 0.00858 0.01365 0.01943 0.03092 0.02901 0.02995 0.02179 0.03297 -0.00093 exposure (sec) 900.000 900.000 900.000 900.000 900.006 900.006 900.006 900.006 600.000 900.006 900.006 900.001 900.006 900.006 900.006 900.001 900.006 900.006 900.006 900.006 900.006 900.006 900.006 900.006 900.006 900.006 900.006 900.006 900.006 900.006 900.000 1800.000 600.000 600.006 899.999 799.999 900.000 900.000 899.999 900.000 900.000 900.000 900.000 900.000 900.006 899.999 899.999 900.000 900.006 900.006 900.006 900.006 900.006 900.000 c(cid:13) 2014 RAS, MNRAS 000, 1–23 20 Amaury H. M. J. Triaud et al. JDB (days) 56192.821879 56193.794605 56215.823377 56221.537217 56239.592460 56256.602476 56257.580498 56264.522496 56264.535284 56264.547494 56264.559264 56264.571220 56264.583638 56264.595350 56264.607201 56264.619145 56264.630926 56264.643541 56264.655415 56264.667243 56264.679106 56264.700945 56307.533598 56323.548244 56460.933662 56567.891937 56608.727539 56927.722485 56928.789883 Vrad (km s−1) -4.07292 -3.56742 -4.10629 -3.74752 -3.87591 -3.60375 -3.53847 -3.70152 -3.70567 -3.71370 -3.72208 -3.72100 -3.72408 -3.73021 -3.73698 -3.74812 -3.76424 -3.77312 -3.78187 -3.77665 -3.78417 -3.80013 -3.69537 -3.45869 -4.06790 -3.50754 -3.69785 -4.07286 -3.51368 σRV (km s−1) 0.00787 0.01278 0.00762 0.01691 0.00851 0.01122 0.00792 0.00630 0.00475 0.00443 0.00418 0.00415 0.00410 0.00391 0.00406 0.00392 0.00387 0.00419 0.00454 0.00507 0.00526 0.00563 0.00686 0.00931 0.01142 0.00974 0.01233 0.00838 0.00559 (%) FWHM contrast (km s−1) 6.41754 6.43435 6.50823 6.51008 6.45703 6.44653 6.44921 6.42133 6.44658 6.45767 6.43897 6.45931 6.43960 6.43981 6.45517 6.43708 6.43620 6.45208 6.44031 6.47638 6.45092 6.44897 6.45089 6.50816 6.50197 6.50537 6.50649 6.44451 6.43032 48.966 48.521 48.183 49.285 48.675 47.974 48.513 48.725 49.066 49.036 49.154 49.191 49.306 49.248 49.099 49.233 49.161 49.142 48.969 48.785 48.850 48.770 48.804 48.367 48.107 48.057 49.948 50.116 49.834 slope bisector span (km s−1) 0.05738 -0.07894 0.00481 0.02560 -0.00723 0.02296 0.02276 0.02209 0.03476 0.00891 0.01169 0.02381 0.02411 0.00984 0.04027 0.01867 0.01727 0.02687 -0.00318 0.02202 0.02116 -0.00597 0.01119 0.03162 0.02895 0.03808 -0.01752 0.04495 0.02100 exposure (sec) 600.006 599.999 900.000 899.999 600.004 600.000 600.008 700.000 900.000 900.001 900.001 900.000 900.001 900.001 900.001 900.000 900.001 900.000 900.000 900.001 900.000 900.000 600.001 600.001 600.000 900.000 900.001 900.001 900.001 Table A4: Radial velocities of WASP-81 obtained with CORALIE before its recent upgrade. BJD is the barycentric Julian date - 2 450 000 days. Vrad is the radial velocity obtained by fitting a cross-correlation function with a Gaussian, σRV is the error on Vrad. FWHM is the full width at half maximum of the cross-correlation function, and contrast is the amplitude. Vrad BJD (days) (km s−1) -60.15011 -60.31782 -60.23827 -60.35013 -60.17890 -60.33501 -60.34664 -60.34739 -60.48137 -60.58673 -60.68491 -60.54821 -60.51078 -60.73022 -60.65253 -60.71255 -60.72716 -60.75905 -60.64218 -60.73914 -60.56578 -60.64301 55833.506244 55834.511338 55835.513064 55851.553175 55852.509304 55856.541346 55859.509870 55862.507592 56050.916807 56067.826597 56068.860596 56069.893252 56075.788615 56076.818326 56101.786777 56103.782366 56108.832891 56133.765892 56135.751516 56147.744152 56151.718366 56154.664701 Table continues next page... σRV (km s−1) 0.02609 0.01861 0.02606 0.02264 0.02674 0.02613 0.02357 0.02726 0.02097 0.02138 0.02516 0.03319 0.02761 0.03529 0.02870 0.01969 0.03353 0.02643 0.03190 0.04144 0.02640 0.05536 (%) FWHM contrast (km s−1) 7.87708 8.06275 7.95214 7.95814 7.99858 7.94386 7.88504 8.04740 8.07575 8.03943 8.05512 8.04313 8.02011 8.07559 7.97893 8.04937 8.05096 8.07510 8.09815 7.95476 7.99750 8.11592 28.619 28.959 29.089 26.734 27.682 25.704 28.508 28.210 29.178 30.584 30.442 30.912 30.192 30.163 30.248 30.321 30.385 30.056 28.651 26.863 29.524 29.581 slope bisector span (km s−1) -0.02494 0.00051 -0.10707 -0.00195 0.01749 0.08807 -0.00745 0.02566 -0.02474 0.00355 0.03245 -0.07498 -0.01121 -0.04915 -0.05570 0.05327 -0.06728 -0.04689 -0.14222 -0.04738 0.08671 -0.05746 exposure (sec) 1800.760 1800.778 1800.742 1800.720 1800.738 1800.760 1800.780 1800.736 1800.735 1800.755 1800.776 1800.735 1800.778 1800.798 1800.841 1800.796 1800.796 1800.777 1800.753 2700.746 2700.607 1800.799 c(cid:13) 2014 RAS, MNRAS 000, 1–23 JDB (days) 56158.595576 56181.617460 56182.616959 56204.525708 56230.518367 56431.902050 56455.806063 56487.786778 56509.710623 56511.692477 56518.621811 56531.555446 56547.652096 56548.506904 56556.531108 56560.494763 56565.533770 56573.536534 56585.548076 56595.517658 56602.513782 56610.522915 56764.896812 56776.887472 56804.815200 56811.785740 56835.751223 56856.750436 56886.668064 56920.621957 56954.527818 Vrad (km s−1) -60.75601 -60.64235 -60.80774 -60.82325 -60.64910 -60.84674 -60.72687 -60.74249 -60.66513 -60.85952 -60.71971 -60.63273 -60.58724 -60.61018 -60.54362 -60.71382 -60.70908 -60.65052 -60.46305 -60.51555 -60.25658 -60.19533 -58.50304 -58.68466 -58.96099 -58.86942 -59.01187 -59.28657 -59.46175 -59.54393 -59.78127 σRV (km s−1) 0.02555 0.01367 0.01615 0.02655 0.02851 0.02090 0.02870 0.02033 0.01397 0.02225 0.01770 0.03202 0.01827 0.01534 0.05433 0.03762 0.01949 0.01500 0.01694 0.01739 0.02280 0.02343 0.01351 0.02343 0.02049 0.01887 0.03321 0.01680 0.01926 0.02171 0.02459 (%) FWHM contrast (km s−1) 7.92280 7.97978 7.95422 8.06633 7.95769 7.98365 7.95465 7.98515 7.95360 7.94794 7.97615 7.97035 8.03302 7.99134 8.01616 7.96320 7.86719 7.97570 7.98314 7.94433 8.11640 7.99369 8.03570 7.97764 7.95361 8.04819 8.10950 8.01407 7.92853 8.03739 7.99078 30.002 29.506 29.259 27.984 22.694 30.384 29.802 30.749 30.834 30.979 30.608 29.748 30.090 28.749 29.137 29.254 29.094 29.996 29.017 29.102 27.697 28.294 30.083 30.796 30.061 30.119 28.842 29.501 30.839 27.815 24.160 The WASP-53 and 81 systems 21 slope bisector span (km s−1) -0.02869 0.01627 0.01257 -0.01402 -0.00547 -0.00998 -0.02273 0.00328 -0.02276 0.01657 0.00907 0.00804 0.02345 -0.03196 -0.01045 0.01284 -0.02711 -0.00324 0.05210 0.00994 0.02679 0.07335 0.00910 0.02733 -0.00476 0.03930 -0.03052 -0.03712 -0.03399 -0.01603 0.00379 exposure (sec) 1800.841 2700.606 2700.707 2700.603 2700.696 2700.679 1800.797 2700.713 2700.732 1800.741 1620.506 2700.637 2700.599 2700.680 2700.631 1800.774 2700.712 2700.713 2700.649 2700.571 2700.797 2700.613 2700.766 2700.763 2700.910 2700.909 2700.800 2700.187 2700.767 2700.428 2699.955 Table A5: Radial velocities of WASP-81 obtained with CORALIE before its recent upgrade. BJD is the barycentric Julian date - 2 450 000 days. Vrad is the radial velocity obtained by fitting a cross-correlation function with a Gaussian, σRV is the error on Vrad. FWHM is the full width at half maximum of the cross-correlation function, and contrast is the amplitude. (%) FWHM contrast (km s−1) 7.95261 7.89406 7.91874 7.97087 7.93436 7.90803 7.96576 8.04004 7.87866 7.81752 7.90775 7.98951 7.85147 7.86095 34.569 34.000 33.627 32.551 32.440 32.325 32.327 33.592 32.934 32.403 32.227 32.002 31.815 32.269 slope bisector span (km s−1) 0.01499 0.06709 -0.06934 -0.07702 -0.11635 0.03218 0.01866 0.00022 -0.16166 -0.05037 -0.03345 -0.09351 0.02024 -0.03440 exposure (sec) 2700.846 2700.943 2700.764 2700.014 2452.822 2700.157 2700.768 600.528 1800.381 2700.072 2700.069 2700.037 1800.422 1800.000 BJD (days) 57186.791639 57194.736148 57211.722549 57256.542816 57271.550777 57294.561176 57324.515825 57584.816334 57595.700710 57650.537435 57652.595356 57661.566328 57680.545006 57682.504365 Vrad (km s−1) -60.24486 -60.23109 -60.44937 -60.39378 -60.56686 -60.47338 -60.41128 -60.83346 -60.85481 -60.75600 -60.87248 -60.75291 -60.81445 -60.89638 σRV (km s−1) 0.03042 0.04093 0.06100 0.04425 0.04207 0.03198 0.05112 0.09427 0.05770 0.02171 0.01754 0.01948 0.04449 0.02700 c(cid:13) 2014 RAS, MNRAS 000, 1–23 22 Amaury H. M. J. Triaud et al. Table A6: Radial velocities of WASP-81 obtained with HARPS. BJD is the barycentric Julian date – 2 450 000 days. Vrad is the radial velocity obtained by fitting a cross-correlation function with a Gaussian, σRV is the error on Vrad. FWHM is the full width at half maximum of the cross-correlation function, and contrast is the amplitude. One datum, which was not used in the analysis, is highlighted with an asterisk. BJD (days) 6403.902377 6407.886489 6411.902064* 6438.878243 6454.847018 6457.848776 6459.922794 6510.582235 6510.591263 6510.601795 6510.613242 6510.624769 6510.636309 6510.647628 6510.659260 6510.670810 6510.682546 6510.693877 6510.705324 6510.717072 6510.728495 6510.740046 6510.751689 6510.763124 6510.774455 6510.786527 6511.701655 6564.490867 6565.479782 6736.904747 6761.912737 6801.852737 6927.561062 Vrad (km s−1) -60.75558 -60.90138 -73.15969 -60.76828 -60.87117 -60.77664 -60.93969 -60.74526 -60.70229 -60.72294 -60.73820 -60.72645 -60.73310 -60.74584 -60.74602 -60.75529 -60.75335 -60.74670 -60.74468 -60.75111 -60.75617 -60.78206 -60.77077 -60.75516 -60.75574 -60.78525 -60.81984 -60.51083 -60.69072 -58.61384 -58.54571 -58.92484 -59.65603 σRV (km s−1) 0.01422 0.01261 0.09454 0.01674 0.01534 0.02008 0.01386 0.01857 0.01341 0.00980 0.00957 0.00900 0.00930 0.00988 0.00985 0.01062 0.01157 0.01072 0.01284 0.01365 0.01286 0.01203 0.01090 0.01073 0.01039 0.01380 0.01049 0.01246 0.01538 0.01996 0.01394 0.03752 0.01215 (%) FWHM contrast (km s−1) 6.77096 6.79512 1.71111 6.81762 6.72971 6.77591 6.78047 6.81107 6.82417 6.75473 6.85763 6.77831 6.79062 6.77485 6.78777 6.81317 6.77487 6.78601 6.76216 6.77729 6.80405 6.75481 6.79623 6.76794 6.79067 6.74017 6.77291 6.78860 6.75616 6.74955 6.80035 6.73650 6.81727 36.778 36.264 2.286 37.631 36.706 36.474 36.319 36.977 36.599 36.746 36.762 36.900 36.931 36.935 37.029 37.005 37.079 36.671 36.814 36.867 36.945 37.010 36.943 37.020 36.959 36.801 36.841 36.623 35.156 37.023 37.818 39.960 37.578 slope bisector span (km s−1) 0.00282 0.02080 789.3231 -0.00578 -0.03163 0.00183 -0.02629 0.06192 0.00540 -0.00742 -0.01984 -0.03973 -0.04372 0.00680 0.00791 0.01163 -0.04139 -0.02780 0.02584 0.01253 -0.03819 -0.00396 -0.00076 -0.01169 0.00137 -0.01100 -0.03984 -0.02331 -0.01447 -0.06442 -0.03933 0.02316 -0.01936 exposure (sec) 600.000 600.001 600.000 900.000 600.001 600.001 600.001 600.001 600.001 900.002 900.001 900.000 900.001 900.001 900.001 900.001 900.000 900.002 900.001 900.001 900.001 900.000 900.001 900.001 900.001 900.000 900.001 600.001 600.000 900.001 900.001 900.001 900.002 c(cid:13) 2014 RAS, MNRAS 000, 1–23 The WASP-53 and 81 systems 23 This paper has been typeset from a TEX/ LATEX file prepared by the author. Figure B1. Hertzsprung–Russell diagram showing the relative positions of WASP-81A and WASP-81B assuming a similar dis- tance. The pair is unlikely to be related. Models are from Marigo et al. (2008). APPENDIX B: THE VISUAL COMPANION TO WASP-81 APPENDIX C: MODELS APPLIED TO THE PHOTOMETRIC DATA c(cid:13) 2014 RAS, MNRAS 000, 1–23 24 Amaury H. M. J. Triaud et al. System Date Instrument Filter Texp Np WASP-53 WASP-53 WASP-53 WASP-53 WASP-53 WASP-53 WASP-53 WASP-53 WASP-81 WASP-81 WASP-81 WASP-81 WASP-81 WASP-81 WASP-81 WASP-81 WASP-81 WASP-81 2011-07-22 2011-09-03 2011-09-13 2011-09-13 2011-09-23 2011-10-26 2012-07-30 2012-11-03 2011-09-26 2012-05-20 2012-05-31 2012-06-19 2012-07-08 2012-07-19 2012-09-24 2013-07-07 2013-08-06 2013-08-06 EulerCAM EulerCAM TRAPPIST EulerCAM EulerCAM EFOSC2 TRAPPIST TRAPPIST TRAPPIST TRAPPIST TRAPPIST TRAPPIST EulerCAM TRAPPIST EulerCAM TRAPPIST EulerCAM TRAPPIST Gunn r' Gunn r' blue blocking blue blocking Gunn r' Gunn r' Gunn r' I + z(cid:48) I + z(cid:48) I + z(cid:48) I + z(cid:48) blueblocking Gunn r' I + z(cid:48) Gunn r' blue blocking Gunn r' blue blocking 90s 120s 12s 80s 180s 150s 12s 15s 12s 25s 12s 8s 120s 20s 120s 10s 80s 10s 110 84 744 127 77 54 603 464 287 372 812 644 116 501 132 917 248 1082 Baseline function p(t2 + xy) p(t2 + xy) p(t2) + o p(t2 + xy + sin(P3t + T0,3)) p(t2 + xy + sin(P4t + T0,4)) p(t2) p(t2) p(t2) p(t2) + o p(t2) + o p(t2) + o p(t2) + o p(t2) p(t2) + o p(t2) p(t2) + o p(t2) p(t2) + o CF 1.3 1.8 1.5 1.7 2.3 1.0 1.5 1.0 0.9 1.1 1.4 1.5 1.9 1.6 1.4 1.9 1.4 1.3 Table C1. Photometric time-series used in this work. For each light curve this table shows the date of acquisition, the instrument and filter used, the exposure time Texp, the number of data points, the baseline function selected for our global analysis (see Sec. 5), and the error correction factor CF used in our global analysis. For the baseline function, p(N ) denotes, respectively, a N -order polynomial function of time ( = t), x and y positions ( = xy); o denotes an offset at the time of a meridian flip of TRAPPIST (see Gillon et al. 2012). On two instances we also fit a sinusoidal baseline of the form sin(P t + T0) where P is the period, and T0 is the phase. c(cid:13) 2014 RAS, MNRAS 000, 1–23 (a) EulerCam r(cid:48) 2011-07-22 (b) EulerCam r(cid:48) 2011-09-03 (c) TRAPPIST BB 2011-09-13 The WASP-53 and 81 systems 25 (d) EulerCam r(cid:48) 2011-09-13 (e) EulerCam r(cid:48) 2011-09-23 (f) TRAPPIST BB 2011-09-23 (g) EFOSC2 r(cid:48) 2011-10-26 (h) TRAPPIST I + z(cid:48) 2012-07-30 Figure C1. Flux as a function of time, centred around mid-transit time of WASP-53b. The red line shows the full model, including the detrending. c(cid:13) 2014 RAS, MNRAS 000, 1–23 0.980.991.00−1 0 1 2−0.0050.0000.005fluxO − Ctime (hours) . . . .0.980.991.00−2−1 0 1 20.000fluxO − Ctime (hours) . . . .0.981.00−2−1 0 1 2−0.010.000.01fluxO − Ctime (hours) . . . .0.980.991.00−1 0 1 2−0.0050.0000.005fluxO − Ctime (hours) . . . .0.980.991.00−1 0 1 2−0.0050.0000.005fluxO − Ctime (hours) . . . .0.981.00−2−1 0 1−0.010.000.01fluxO − Ctime (hours) . . . .0.980.991.00−1.5−1.0−0.50.00.51.0−0.0020.0000.002fluxO − Ctime (hours) . . . .0.981.00−1.5−1.0−0.50.00.51.0−0.020.000.02fluxO − Ctime (hours) . . . . . . . . 26 Amaury H. M. J. Triaud et al. (a) TRAPPIST I + z(cid:48) 2011-09-26 (b) TRAPPIST I + z(cid:48) 2012-05-20 (c) TRAPPIST I + z(cid:48) 2012-05-31 (d) TRAPPIST I + z(cid:48) 2012-06-19 (e) EulerCam r(cid:48) 2012-07-08 (f) TRAPPIST I + z(cid:48) 2012-07-19 (g) EulerCam r(cid:48) 2012-09-24 (h) TRAPPIST BB 2013-07-07 (i) EulerCam r(cid:48) 2013-08-06 (j) TRAPPIST BB 2013-08-06 Figure C2. Flux as a function of time, centred around mid-transit time of WASP-81b. The red line shows the full model, including the detrending. c(cid:13) 2014 RAS, MNRAS 000, 1–23 0.991.001.011.02−1.5−1.0−0.50.0−0.010.000.01fluxO − Ctime (hours) . . . .0.981.00−2−1 0 1−0.010.000.01fluxO − Ctime (hours) . . . .0.981.00−1 0 1 2 30.000.02fluxO − Ctime (hours) . . . .0.981.00−0.50.00.51.01.52.02.53.0−0.010.000.01fluxO − Ctime (hours) . . . .0.991.00−3−2−1 0 10.000fluxO − Ctime (hours) . . . .0.981.00−2−1 0 1 2−0.010.000.01fluxO − Ctime (hours) . . . .0.991.00−2−1 0 1 20.000fluxO − Ctime (hours) . . . .0.981.001.02−2−1 0 1 2 3−0.020.000.02fluxO − Ctime (hours) . . . .0.991.00−3−2−1 0 1 2 30.000fluxO − Ctime (hours) . . . .0.981.00−2−1 0 1 2 30.00fluxO − Ctime (hours) . . . . . . . . . . . .
1510.02094
1
1510
2015-10-07T20:07:34
Growing the gas-giant planets by the gradual accumulation of pebbles
[ "astro-ph.EP" ]
It is widely held that the first step in forming the gas giant planets, such as Jupiter and Saturn, is to form solid `cores' of roughly 10 M$_\oplus$. Getting the cores to form before the solar nebula dissipates ($\sim\!1-10\,$Myr) has been a major challenge for planet formation models. Recently models have emerged in which `pebbles' (centimeter- to meter-size objects) are first concentrated by aerodynamic drag and then gravitationally collapse to form 100 --- 1000 km objects. These `planetesimals' can then efficiently accrete leftover pebbles and directly form the cores of giant planets. This model known as `pebble accretion', theoretically, can produce 10 M$_\oplus$ cores in only a few thousand years. Unfortunately, full simulations of this process show that, rather than creating a few 10 M$_\oplus$ cores, it produces a population of hundreds of Earth-mass objects that are inconsistent with the structure of the Solar System. Here we report that this difficulty can be overcome if pebbles form slowly enough to allow the planetesimals to gravitationally interact with one another. In this situation the largest planetesimals have time to scatter their smaller siblings out of the disk of pebbles, thereby stifling their growth. Our models show that, for a large, and physically reasonable region of parameter space, this typically leads to the formation of one to four gas giants between 5 and 15 AU in agreement with the observed structure of the Solar System.
astro-ph.EP
astro-ph
Growing the Gas Giant Planets by the Gradual Accumulation of Pebbles Harold F. Levison1, Katherine A. Kretke1, and Martin J. Duncan2 1 Southwest Research Institute and NASA Solar System Exploration Research Virtual Insti- tute, 1050 Walnut St, Suite 300, Boulder, Colorado 80302, USA 2 Department of Physics, Engineering, and Astronomy, Queen's University, Kingston, On- tario K7L 3N6, Canada Prepared for Nature July 5, 2018 5 1 0 2 t c O 7 . ] P E h p - o r t s a [ 1 v 4 9 0 2 0 . 0 1 5 1 : v i X r a 1 It is widely held that the first step in forming the gas giant planets, such as Jupiter and Saturn, is to form solid 'cores' of roughly 10 M⊕ [1, 2]. Getting the cores to form before the solar nebula dissipates (∼ 1−10 Myr [3]) has been a major challenge for planet formation models [4, 5]. Recently models have emerged in which 'pebbles' (centimeter- to meter-size objects) are first concentrated by aerodynamic drag and then gravitationally collapse to form 100 - 1000 km objects [6, 7, 8, 9]. These 'planetesimals' can then efficiently accrete leftover pebbles [10] and directly form the cores of giant planets [11, 12]. This model known as 'pebble accretion', theoretically, can produce 10 M⊕ cores in only a few thousand years [11, 13]. Unfortunately, full simulations of this process [13] show that, rather than creating a few 10 M⊕ cores, it produces a population of hundreds of Earth-mass objects that are inconsistent with the structure of the Solar System. Here we report that this difficulty can be overcome if pebbles form slowly enough to allow the planetesimals to gravitationally interact with one another. In this situation the largest planetesimals have time to scatter their smaller siblings out of the disk of pebbles, thereby stifling their growth. Our models show that, for a large, and physically reasonable region of parameter space, this typically leads to the formation of one to four gas giants between 5 and 15 AU in agreement with the observed structure of the Solar System. Our models consist of a series of computer simulations that follow the evolution of a population of objects in a disk around the Sun. The solar composition disk has surface density distribution Σ ∝ r−1 AU, where rAU is the distance from the Sun, and consists of both gas and solids. We assume that an initial population of planetesimals, which follow the surface density of the disk, form quickly and thus exist at the beginning of our simulations. These planetesimals contain only a small fraction of the mass of solids available (a free 2 parameter of our models). Pebbles either also existed at the beginning of the simulation, or are allowed to form over some period of period of time (again a free parameter) starting at the beginning of the calculation. The evolution of this system is followed numerically and includes the effects of gravitational interactions, interactions between bodies in the disk and the gas (although nebular tidal migration is neglected), accretion (including enhancements due to the aerodynamic drag on pebbles; e.g. Ref. [11]), and collisional fragmentation. Details of our setup and numerical techniques are described in the Methods Section. In Figure 1 we present the results of two different simulations that employ the same parameters except for the method of pebble formation. In the first (Figure 1a), we assume that the pebbles are leftovers of planetesimal formation and thus they exist at the beginning of the simulation. Note the fast accretion times - embryos (defined to be objects that grow significantly) evolved from roughly the mass of Pluto to that of the Earth in only 1000 years! However, even though Earth-mass embryos form quickly, this simulation does not reproduce the Solar System because rather than forming a few cores, it creates ∼ 100 Earth-mass objects. These objects subsequently scatter one another to high-eccentricity, high-inclination orbits, stalling growth. This result is clearly inconsistent with observations. Ref. [13] finds that this is a generic outcome for this type of pebble accretion simulations and thus it could not have occurred in the Solar System. It is indicative of the results one would achieve if the dynamical interactions between planetesimals is neglected (which we refer to as the standard model). In Figure 1b, we employ a substantial modification to the pebble accretion theory where we couple an extended timescale for pebbles formation (a conjecture supported both by the observed presence of cm to mm-sized grains, which are expected to drift rapidly out of protoplanetary disks, in disks of a wide variety of ages; e.g. Ref.[14] and theoretical models of pebble formation [15]) to the dynamical stirring due to planetesimal self-gravity (so-called 3 viscous stirring [16]). The behavior of this simulation is radically different from that in Figure 1a in that, rather than producing a large population of Earth-mass objects, a few giant planets form. Only 5 objects grow to 1 M⊕ or larger in this simulation, and there are two gas giants at 10 Myr. Also, note that the timescale for growth is very different in the two calculations. The first Earth-mass object grows in 1000 years in the standard pebble accretion run, as opposed to taking slightly over 400,000 years in this simulation. In the latter case, the growth time is determined by the rate at which pebbles were created. The outcomes of the two simulations are remarkably different because, in the Figure 1b simulation only the most massive embryos are able to accrete a significant amount of pebbles. This is due to the fact that the embryos grow slowly enough in this model that they can gravitationally interact with one another as they accrete the pebbles. This viscous stirring has two effects on the embryo's accretion rates. First, it increases the relative velocities and thus decreases the capture cross-section (c.f. Methods Section, Eq. 4). More importantly, encounters between embryos lead to increases in the inclinations, i, of the embryos and thus the distances they travel above and below the disk midplane. Once the inclinations of the embryos become larger than that of the aerodynamically damped pebbles, the embryos spend much of their time above or below the location where the bulk of the pebbles lie. This effectively starves them and stifles their growth. Due to the role of viscous stirring in determining the outcome in these simulations, we refer to this process as viscously stirred pebble accretion (VSPA). Figure 2 illustrates this effect. In the standard pebble accretion model, the system inclinations remain small and thus all the planetesimals can grow. This because the timescale for viscous stirring is longer than that of the growth. In this case it can be shown (c.f. Methods Section) that dMe/dt ∝ M β e and β ≪ 1. The small value of β implies that smaller embryos can catch up with larger ones, leading to a population of like-sized objects. 4 However, the dynamical evolution of the system is very different in the VSPA simulation. Here viscous stirring can act thereby leading to an increase of inclinations. The magnitude of this increase is set by the mass of the largest embryo - note the increase with time of both the inclination of the smaller embryos and the largest embryo mass in Figure 2b. Initially the inclinations of the embryos are smaller than those of the pebbles, but by 3000 years, most of the smaller embryos begin to spend a significant amount of time above or below the pebbles. The time at which embryos are excited out of the swarm of pebbles can be estimated with a simple calculation. The inclinations of the pebbles are set by the balance between turbulent excitation and aerodynamic drag in the disk. For our example, the average pebble inclination is 0.0016 radians. Given enough time, a population of embryos will stir one another to the point where their relative velocities are comparable to their surface escape velocities. Since i ∼ vrel/vc, even if we assume objects do not grow, our population of planetesimals should reach inclinations of ∼ 0.06 - significantly larger than that of the pebbles. As a result, this effect should be important during much of our simulation. Indeed, Figure 2 shows that most of the embryos in the system are in this state after only ∼ 3000 years. The exception is the most massive embryos (Figure 2b), which tend to have low incli- nations as a result of a combination of gravitational interactions with smaller planetesimals (so-called dynamical friction [16, 17]) and with the gas (so-called Type I inclination damping [18]). As a result, the larger embryos grow relatively quickly while the smaller embryos grow very slowly, if at all. Recall that the standard model of pebble accretion formed a large num- ber of planets because β < 1, which allowed smaller embryos to catch up with larger ones. Figure 3 shows the relationship between dMe/dt and Me in our fiducial VSPA run as the system evolves. At early times, β is slightly larger than 1, leading to a few embryos becoming dominate in the population. When the largest objects in the system have masses between 5 ∼ 8 × 10−3 M⊕ and ∼ 0.02 M⊕ their equilibrium inclinations decrease, leading to a spurt of growth where β is large (∼ 4). This allows a small number of embryos to become separated in mass from the rest, and explains why only a small number of embryos become massive enough to become giant planet cores. Indeed, for the small embryos, dMe/dt decreases with time because of their increasing inclinations (see Figure 2). However for Me > ∼ 0.02 M⊕, β becomes small again as the strongly damped, large embryos accrete in the same manner as in the standard pebble accretion scenario. This last phase is important because it forces these proto-cores to have similar masses as they grow. Thus, they all reach the mass where they can directly accrete gas at roughly the same time. This last phase might also explain why the four cores of the giant planets likely originally had similar masses. Two gas giants grow in our fiducial simulation (Jupiter and Saturn?). By varying parameters in the models, our systems produced between 0 and 4 gas giants. Results of the simulations can be found in Extended Data Tables 1 and 2. There is another constraint that any model of giant planet formation in the Solar System must satisfy in order to be considered a success. The distribution of small bodies in the outer Solar System indicates that the orbits of the giant planets moved substantially after they formed [19, 20, 21]. In particular, Uranus and Neptune likely formed within 15 or 20 AU of the Sun and were delivered to their current orbits by either a smooth migration [20] or a mild gravitational instability [21]. Both processes require that a population of planetesimals existed on low-eccentricity orbits beyond the giant planets after the planets finished forming. This population must have: a) not formed planets, and b) survived the planet formation process relatively unscathed. To evaluate this constraint in our simulations, we placed 5 planetesimals with radii similar to Pluto (s = 1350 km) on circular, co-planar orbits with semi-major axes between 20 and 30 AU. None of these objects grew in our fiducial simulation and they all survived on orbits with eccentricities less than 0.07 (see the Methods Section 6 for how this varies with disk parameters). Therefore, the process of viscously stirred pebble accretion reproduces the observed structure of the outer Solar System: two gas giants, a few icy planets, and a disk of planetesimals into which the ice giants can migrate. References [1] Mizuno, H., Nakazawa, K. & Hayashi, C. Instability of a gaseous envelope surrounding a planetary core and formation of giant planets. Progress of Theoretical Physics 60, 699–710 (1978). [2] Pollack, J. B. et al. Formation of the Giant Planets by Concurrent Accretion of Solids and Gas. Icarus 124, 62–85 (1996). [3] Haisch, K. E., Lada, E. A. & Lada, C. J. Disk Frequencies and Lifetimes in Young Clusters. Astrophys. J. Lett. 553, L153–L156 (2001). [4] Goldreich, P., Lithwick, Y. & Sari, R. Final Stages of Planet Formation. Astrophys. J. 614, 497–507 (2004). [5] Levison, H. F., Thommes, E. & Duncan, M. J. Modeling the Formation of Giant Planet Cores. I. Evaluating Key Processes. Astron. J. 139, 1297–1314 (2010). [6] Cuzzi, J. N., Hogan, R. C., Paque, J. M. & Dobrovolskis, A. R. Size-selective Concen- tration of Chondrules and Other Small Particles in Protoplanetary Nebula Turbulence. Astrophys. J. 546, 496–508 (2001). [7] Youdin, A. N. & Goodman, J. Streaming Instabilities in Protoplanetary Disks. Astro- phys. J. 620, 459–469 (2005). 7 [8] Johansen, A. et al. Rapid planetesimal formation in turbulent circumstellar disks. Nature 448, 1022–1025 (2007). [9] Youdin, A. N. On the Formation of Planetesimals Via Secular Gravitational Instabilities with Turbulent Stirring. Astrophys. J. 731, 99 (2011). [10] Ormel, C. W. & Klahr, H. H. The effect of gas drag on the growth of protoplanets. Ana- lytical expressions for the accretion of small bodies in laminar disks. Astron. Astrophys. 520, A43 (2010). [11] Lambrechts, M. & Johansen, A. Rapid growth of gas-giant cores by pebble accretion. Astron. Astrophys. 544, A32 (2012). [12] Lambrechts, M. & Johansen, A. Forming the cores of giant planets from the radial pebble flux in protoplanetary discs. Astron. Astrophys. 572, A107 (2014). [13] Kretke, K. A. & Levison, H. F. Challenges in Forming the Solar System's Giant Planet Cores via Pebble Accretion. Astron. J. 148, 109 (2014). [14] Ricci, L. et al. Dust properties of protoplanetary disks in the Taurus-Auriga star forming region from millimeter wavelengths. Astron. Astrophys. 512, A15+ (2010). [15] Birnstiel, T., Klahr, H. & Ercolano, B. A simple model for the evolution of the dust population in protoplanetary disks. Astron. Astrophys. 539, A148 (2012). [16] Stewart, G. R. & Wetherill, G. W. Evolution of planetesimal velocities. Icarus 74, 542–553 (1988). [17] Ida, S. Stirring and dynamical friction rates of planetesimals in the solar gravitational field. Icarus 88, 129–145 (1990). 8 [18] Ward, W. R. Protoplanet Migration by Nebula Tides. Icarus 126, 261–281 (1997). [19] Fernandez, J. A., Ip, W.-H. Some dynamical aspects of the accretion of Uranus and Neptune - The exchange of orbital angular momentum with planetesimals. Icarus 58, 109-120 (1984). [20] Malhotra, R. The Origin of Pluto's Orbit: Implications for the Solar System Beyond Neptune. Astron. J. 110, 420–+ (1995). [21] Tsiganis, K., Gomes, R., Morbidelli, A. & Levison, H. F. Origin of the orbital architec- ture of the giant planets of the Solar System. Nature 435, 459–61 (2005). Acknowledgements This work was supported by an NSF Astronomy and Astrophysics Research Grant (PI Levi- son). We would like to thank A. Johansen, M. Lambrechts, A. Morbidelli, D. Nesvorny, and C. Ormel for useful discussions. 1 Author Contributions H.F.L. and K.A.K. jointly conceived of the paper and carried out the bulk of the numerical and semi-analytic calculations. M.D. developed a semi-analytic model of viscous stirring and growth rates in a population distribution. All authors contributed to the discussion of the results and to the crafting of the manuscript. 9 Author Information The authors declare no competing financial interests. Correspondence and requests for ma- terials should be addressed to [email protected] Figure Legends Methods Computational Methods: LIPAD For this work we have employed a particle based Lagrangian code (known as LIPAD, Lagrangian Integrator for Planetary Accretion and Dynamics) that can follow the dynam- ical/collisional/accretional evolution of a large number of planetesimals through the entire growth process to become planets. For full details about the code and extensive test suites see Ref. [22]. For details about how pebble accretion is implemented in the code see Ref. [13], but we summarize the most relevant attributes of the code here. LIPAD is built on top of of the N-body integrator SyMBA [23]. In order to handle the very large number of sub-kilometer objects required by many simulations, LIPAD utilizes a concept known as tracer particles. Each tracer represents a large number of small bodies with roughly the same orbit and size, and is characterized by three numbers: the physical radius, the bulk density, and the total mass of the disk particles represented by the tracer. LIPAD employs statistical algorithms that follow the dynamical and collisional interactions between the tracers. When a tracer is determined to have been struck by another tracer, it is assigned a new radius according to the probabilistic outcome of the collision based on a fragmentation law by Ref. [24], using the ice parameters Q0 = 7×107 erg g−1, B = 2.1 erg cm3 g−2, a = −0.45 10 M > N M > N 107 106 105 104 103 102 101 100 106 105 104 103 102 101 100 t (yr) 0 500 1000 104 106 107 t (yr) 0 105 5 ×105 106 107 a) b) 10-6 10-5 10-4 10-3 10-2 10-1 100 101 102 103 M (M ⊕) Figure 1: The cumulative mass distribution of planetesimals and embryos. The growth of planetary embryos in our simulation as illustrated by cumulative mass distributions shown at various times (indicated by color; see the legend). a) An example of standard pebble accretion - all pebbles were in existence at the beginning of the simulation. This is similar to Ref. [13]. b) An example with the same disk parameters as (a), but where the pebbles formed over the lifetime of the disk. This system formed two gas giants and three 11 icy planets with masses between 2 and 4 M⊕. The icy planets are on crossing orbits at the end of the simulation (10 Myr) and thus their number and masses would likely change as the calculation were run to completion. A movie version of this figure is available in the supplementary material. ) n o i t a n i l c n i ( n a t 10-1 10-2 10-3 10-4 10-5 10-6 10-1 10-2 10-3 10-4 10-5 10-6 a) b) t (yr) 0 300 500 1000 1500 t (yr) 0 3000 3 ×105 6 ×105 106 10-6 10-5 10-4 10-2 10-1 100 10-3 M (M ⊕) 12 ) i ( n a t 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0.0 l < s e b b e p f o n o i t c a r F Figure 2: The vertical distribution of pebbles and embryos. A comparison between the vertical distribution, as represented by tan (i), of pebbles and embryos in the simulations shown in Figure 1. The embryos are shown as dots in the figure, where their color indicates the time within the simulation (see legend). Since objects grow at different times within ) ⊕ 100 10-1 10-2 10-3 10-4 10-5 10-6 10-7 10-8 M 1 = M o t d e a c s ( t d / e l M d dM dt ∝M t (yr) 0 3000 3 ×105 4 ×105 6 ×105 106 100 dM dt ∝M4 10-4 10-2 10-3 Embryo Mass (M ⊕) 10-1 Figure 3: Embryo growth rate as a function of mass. The temporal evolution of the relationship between embryo growth rate (dMe/dt) and mass (Me). The dots show our fidu- cial VSPA simulation (Figure 1b), where their color indicates the time within the simulation (see legend). The solid curve is from the standard pebble accretion run in Figure 1a. These values were calculated using procedures described in the Methods Section. The behavior of the system is determined by the slope of the data in the figure, β (note that this is a log-log plot) at each time. For reference, the dotted lines show β = 1 and β = 4. For β < 1, small embryos can catch up with larger ones leading to a population of like-sized objects. For β > 1 the largest embryos run away from their smaller siblings. See Methods section for more details. 13 and b = 1.19. This way, the conglomeration of tracers represents the size distribution of the evolving planetesimal population. In this work, we do not allow our pebbles to collisionally grow or fragment, therefore particles below 1 km in size are not involved in the collisional cascade. LIPAD also includes statistical algorithms for viscous stirring, dynamical friction, and collisional damping among the tracers. The tracers mainly dynamically interact with the larger planetary mass objects via the normal N-body routines, which naturally follow changes in the trajectory of tracers due to the gravitational effects of the planets and vice versa. LIPAD is therefore unique in its ability to accurately handle the mixing and redistribution of material due to gravitational encounters, including gap opening, and resonant trapping, while also following the fragmentation and growth of bodies. Thus, it is well suited to follow the evolution of a population of embryos, planetesimals, and pebbles while they gravitationally interact to form planets. The pebble accretion model follows the prescription in Ref. [10] and is described in detail in Ref. [13]. An overview of the physics follows in the next section. In the simulations presented here we do not allow the growing planets to migrate via type- I migration [18], although we do include type-I eccentricity damping [25]. We also neglect type-II migration. We calculate the aerodynamic drag on all bodies using the formalism of Ref. [26]. Additionally, as we are interested in the gross evolution of a system after the formation of a potential giant planet core, we have added a simple optional prescription allowing cores to accrete gas envelopes. In order to accrete gas the core size must be above a critical value, which depends on the mass accretion rate of solids onto the core. We follow Ref. [27] to determine when core masses are above the critical value given their current mass accretion rate (assuming a grey opacity of 0.02 times that of the ISM [2]). If this criteria is met then 14 we grow the planet of mass Mp on the Kelvin Helmholtz timescale (tKH), so that Following Ref. [28] we approximate this timescale as Mg = Mp tKH . tKH = 109 Mp M⊕!−3 yrs. We limit gas accretion to the Bondi accretion rate, Mg,max = 4πρgG2M 2 p c3 s , (1) (2) (3) where ρg is the gas density and cs is the local sound speed. We note that a major uncertainty in these models is the envelope opacity, which is dominated by the poorly constrained prop- erties of the dust. Varying the opacity dramatically alters both the size of the critical core mass, and perhaps more strikingly, the Kelvin-Helmholtz time. Additionally, we arbitrarily cut off gas accretion when a planet reaches one Jupiter mass instead of including any physics to reduce the accretion of gas after gap opening (e.g. Ref. [29]). Furthermore, we do not include the fact that pebbles will cease accreting onto planetesimal cores once the cores begin perturbing the disk [30]. This effect may prove important in causing some cores to accrete gas sooner, or in allowing cores to grow to larger sizes without accreting gas (creating true Uranus and Neptune analogs). Therefore the end masses of giant planets should be viewed with these caveats in mind. We also note that while this model includes fragmentation, we have found that the production, and subsequent sweep-up of pebbles (as considered by Ref. [31]) is relatively unimportant in these scenarios. This is due to the assumed relatively low initial mass of planetesimals in our simulations (17 M⊕ between 4 and 15 AU in our fiducial simulation). 15 Pebble Accretion Here we present an argument, based on Refs. [10, 11], about why pebbles are effectively accreted by growing planets. If the stopping time (ts) of a pebble is comparable to the time for it to encounter a growing embryo, then it can be deflected out of the gas stream and lose enough orbital energy to become gravitationally bound to the embryo. After capture, the pebble spirals inward due to aerodynamic drag and is accreted. In this case the collisional cross section for accretion is σpeb ≡ π(cid:18)4GMets vrel (cid:19) exp"−2 tsv3 4GMe!γ# , rel (4) where Me is the mass of the embryo, vrel is the relative velocity between the pebble and embryo, G is the gravitational constant, and γ = 0.65 [10]. When the Stokes number of a pebble (τ ≡ tsΩK, where ΩK is the local Keplerian frequency) is near unity, this capture radius can be 107 times larger than the physical cross section alone (πR2 e, where Re is the radius of the embryo) for Earth-sized planets in the region inhabited by the giant planets. This process can cause an isolated 1000 km object to grow into a 10 M⊕ core in only a few thousand years [11, 13]. To zeroth order, the accretion rate of an embryo growing by pebble accretion is dMe/dt = facR, where fac is the fraction of pebbles that drift into an embryos orbit that will be accreted (also called the filtering factor, which is a function of σpeb and the spatial distribution of pebbles), R = 2πrΣpebvrad is the rate at which pebbles are fed to the embryo, Σpeb is the surface density of pebbles at heliocentric distance r, and vrad is the radial velocity of the pebbles due to aerodynamic drag. Pebbles drift at a velocity vrad = −2 τ τ 2 + 1 ηvc, (5) where η is a dimensionless parameter related to the pressure gradient of the gas disk [32] and is on the order of a few times 10−3, and vc is the local circular velocity of the embryo. 16 For τ ∼ 1, a pebble can spiral through the disk is only a few hundred years. These large values of vrad can lead to huge accretion rates and giant planet cores can potentially grow quickly. Pebble Formation Model In this work we utilize a simple prescription to convert dust into pebbles over time. We assume that the gas disk with mass Mg exponentially decays over a timescale tg, so that Mg = − Mg tg . (6) Pebbles (with total mass Mpeb) are formed at a rate proportional to the square of the dust mass (Md) such that Mpeb = kM 2 d , (7) where k determines the rate of pebble formation. The dust in the disk will be lost as the gas disk evolves and as pebbles form, yielding a dust evolution of Md = Mg Md Mg − Mpeb. This leads to a total rate of pebble production of Mpeb = κ tg Md,0 1 (1 + κ)2 exp(t/tg) − κ/(1 + κ)!2 1 , (8) (9) where κ ≡ ktgMd,0. Motivated by observations of disks, we take the time constant for pebble production to be large, with a median production timescale near 1 Myr. We assume that all of the pebbles are produced in 3 Myr. For simplicity, we assume that pebbles are randomly created throughout the disk according to the surface density, and the size of the pebbles is determined by the assumed τ , which is constant in each simulation, but varies between calculations. 17 To test the robustness of our result to the specific assumptions of our pebble formation model, we modified it in two different ways for the runs shown in the Extended Data Tables. In some runs we allow the pebbles to grow as they drift inwards by sweeping up dust. Their growth rate is thus dm dt = ρdπr2 pebvrel, (10) where ρd is the dust mass (which is depleted as pebbles are formed, as dust is swept up, and as the gaseous disk evolves), rpeb is the size of the pebble, and vrel is the relative velocity between the pebble and the dust (which is assumed to be perfectly coupled to the gas). In other runs, we assume a model of pebble formation consistent with the inside out manner suggested by Ref. [15]. We used the implementation as described in Ref. [12] and had a wave of pebble formation which moves outwards as a function of time. In particular, the radius at which pebbles are generated (rg) at time t is rg(t) = (cid:18) 3 16(cid:19)1/3 (GM∗)1/3(ǫdZ0)2/3t2/3, (11) where Z0 is the metallicity of the disk and ǫd encapsulates the efficiency of particle growth. We modify ǫd to adjust the timescale of pebble formation. We found that so long as the coagulation coefficient is small enough that the timescale for pebble formation is long, we can get results that are generally similar to our fiducial model in which pebbles form randomly. Model Parameters and Simulation Statistics Since we are interested in building the gas giant planets, we study the growth of planetesimals spread from 4 to 15 AU. As we integrate the simulation forwards, pebbles are formed between 4 and 30 AU. Each system was evolved 10 Myr using LIPAD. Extended Data Tables 1 and 2 list the 42 simulations that we have completed in our investigation of giant planet formation. The eight important parameters that were varied are: 18 1. The surface density of the gas disk at 1 AU, Σ0. In all our simulations we used a surface density distribution of Σ = Σ0r−1 AU[33], where rAU is the heliocentric distance in AU. For the fiducial simulation, Σ0 = 7200 g/cm2, which is 4 times the surface density of the minimum mass solar nebula at that location [34]. The gas surface density decreases exponentially with a timescale of 2 Myr [3, 35]. 2. We employ a flaring gas disk with a scale height h = 0.047 rq AU AU. We used two values for q: 1.25 from Ref. [34] and 9/7 from Ref. [36]. The later was used in our fiducial run. 3. The size of the largest planetesimal, smax. We draw the initial planetesimals from a distribution of radii, s, of the form dN/ds ∝ s−4.5 such that s is between 100 km [37] and smax. For our fiducial simulation smax = 1350 km, which is slightly larger than Pluto. 4. The fraction of solids in the disk that are initially converted into planetesimals, fpl. For all simulations, we assume a solid-to-gas ratio of 0.01, which includes the contribution of water ice [38], and fpl of the solids are in planetesimals while (1 − fpl) are in pebbles. For the fiducial simulation, fpl = 0.1, which, if extended to 30 AU, is consistent with the mass of planetesimals needed to subsequently deliver the giant planets to their current orbits [21]. 5. The initial Stokes number of a pebble, τ . Note that as pebbles spiral toward the Sun, their Stokes number changes (generally decreases) even when it is assumed that they do not grow. For our fiducial simulation τ = 0.6. 6. The strength of inclination and eccentricity damping due to the gravitational interac- tion with the gas disk, ce. We employ the techniques in Ref. [5], which are based on 19 Ref. [25] and allow for both radial migration and inclination/eccentricity damping. The strength of both processes can be adjusted by varying two dimensionless parameters: ca and ce, respectively. All our simulations have ca = 0, meaning there is no Type I migration. For our fiducial simulation ce = 0.66. 7. LIPAD [22], has the option of allowing pebbles to grow by the accretion of dust particles suspended in the gas. The growth rates are calculated using the particle-in-a-box approximation assuming the distribution of dust follows that of the gas, as described above. Our fiducial simulation has this option disabled. 8. The median time of pebble generation, tgen. For most of our simulations, pebble generation follows the evolution of the gas disk. In this case tgen ∼ 0.7 Myr. However, in two cases we decreased tgen in order to determine whether it affects our results. We found that the value of tgen is not important as long as it is larger than the viscous stirring timescale of the embryos. Times marked with an asterisk indicate that pebbles were generated in an inside-out manner as described above. Additionally, all the disks are turbulent with α = 4 × 10−3 [39]. The last five columns in the table present the basic characteristics of our systems. In particular, N Gas is the total number of gas giant planets that formed during each run. It ranges from 0 to 4. Occasionally, a giant planet or two was lost from the system. Although some of these were ejected due to dynamical instabilities, the majority were pushed off the inner edge of our computational domain by their neighbors as the latter accreted gas. Un- fortunately, the inner edge of the domain needs to be relatively large because it determines the timestep used in the calculation. The N-body integrator requires that the perihelion passage be temporally resolved, and since the location of the domain's inner boundary de- termines the smallest perihelion, it also sets the timestep. In order for the calculations to 20 be practical, we were forced to remove any object with a perihelion distance smaller than 2.7 AU (the location of the snow line). The majority of the lost gas giants would likely have remained in the asteroid belt or migrated into the terrestrial planet region if we had been able to keep them in the calculations. The column labeled N Gas f gives the number of gas giants remaining in the system at 10 Myr. Columns (11) and (12) refer to icy objects with masses greater than 1 M⊕. Again, N ice and N ice f are the total number of planets that formed and the number remaining at 10 Myr, respectively. Unlike their larger siblings, the majority of the lost icy planets were ejected from the planetary system, mainly by encounters with growing gas giants. The number of lost ice planets illustrates the fact that the systems becomes violent after the largest cores begin to accrete gas. The last column shows whether the 5 planetesimals we placed on circular, co-planar orbits between 20 and 30 AU (as proxies for the planetesimal disk needed for later planet migration [21]) survived at 10 Myr. There are a couple of important caveats that must be taken into account to properly interpret the results in the Extended Data Tables. First the parameter space for our calcula- tions is 8-dimensional, and each calculation required many weeks of CPU time. This limited the total number that could be performed. Thus, it was impossible to cover parameter space in a meaningful way. Instead, we surgically approached the problem by choosing a couple of starting locations and varying individual parameters to test their effects. We then moved in a particular direction in parameter space if we believed we would be more likely to pro- duce systems with some desired characteristic. As a result, the distribution of our results seen in the table (for example the number of gas giants) cannot be interpreted as the true distribution that would result if parameter space were uniformly covered. Our second concern is the crude methods we used to model the direct accretion of gas onto the cores to form gas giants. As discussed above, our simple model for calculating for 21 the onset of gas accretion and the final mass of planets likely is missing important physics. These limitations mean that, while our calculations show the pebble accretion produces the correct number of giant planet cores on the correct timescales, the details of the evolution of our systems after the gas giants form should be viewed with skepticism. Generating Figure 3 The data in Figure 3 were generated using the following procedure. We first needed to construct a high resolution distribution of pebbles. This was accomplished by summing the output steps of our integration over all time. Then at each time plotted, we noted the mass and orbit of each embryo. We then generated 1000 clones of each embryo - each with the same semi-major axis, eccentricity, and inclination, but with different orbital angles. For each clone, we calculated its velocity with respect to the pebbles and then σpeb from Eq. (4). From this we estimate the instantaneous accretion rate using the particle-in-a-box Me = ρ σpebvrel where ρ is the local mass density of pebbles. The dots in approximation, the figure show the average of the 1000 clones of the instantaneous accretion rates for each embryo. Code Availability The calculations presented in this paper were performed using LIPAD, a proprietary software product funded by the Southwest Research Institute and is not publicly available. It is based upon the N-Body integrator SyMBA, which is publicly available at http://www.boulder.swri.edu/swifter/ 22 Table 1: Our completed simulations. Please see the Methods Section for a full description of the columns in this table. (1) Σ0 (2) (3) (4) (5) (6) (7) (8) (9) (10) (11) (12) (13) q smax fpl τ ce pebbles tgen N Gas N Gas f N ice N ice f Disk ok? (g/cm2) (km) grow? (Y/N) (Myr) (Y/N) 3600 9/7 1500 3600 9/7 1500 3600 9/7 1500 3600 9/7 1500 3600 1.25 1500 0.2 0.2 0.2 0.2 0.2 0.3 0.3 0.3 0.4 3 4050 9/7 1500 0.18 0.4 4050 9/7 1500 0.18 0.5 4050 9/7 1500 0.18 0.6 4500 9/7 1500 0.16 0.3 4500 9/7 1500 0.16 0.4 4500 9/7 1500 0.16 0.5 1 1 1 1 1 1 1 1 1 1 1 4500 9/7 1500 0.16 0.5 0.33 4500 9/7 1500 0.16 0.5 0.1 4500 9/7 1500 0.16 0.6 4500 9/7 1500 0.16 0.7 4500 9/7 1500 0.16 0.9 5400 1.25 1500 0.15 5400 1.25 1500 0.15 5400 1.25 1500 0.15 5400 1.25 1500 0.15 3 3 3 3 7200 9/7 1350 0.25 0.6 7200 9/7 1350 0.15 0.6 7200 1.25 1500 7200 1.25 1500 7200 1.25 1500 0.1 0.1 0.1 1 1 1 1 1 1 1 1 0.66 0.33 1 1 1 0.25 0 N N N N Y N N N N N N N N N N N Y Y Y Y N N Y Y Y 23 0.7 0.5 0.2 0.7 0.7 0.7 0.7 0.7 0.7 0.7 0.7 0.7 0.7 0.7 0.7 0.7 0.7 0.7 0.7 0.7 0.7 0.7 0.7 0.7 0.7 4 4 4 3 3 4 3 4 4 3 3 4 0 2 0 1 0 0 0 1 4 4 2 3 2 2 3 3 1 2 2 3 4 3 2 3 2 0 2 0 0 0 0 0 1 4 4 2 3 2 12 14 15 9 9 6 7 9 13 7 13 4 7 15 7 1 5 6 5 7 10 8 11 5 1 0 0 1 2 0 1 0 0 0 1 0 0 6 3 7 1 4 6 5 4 2 0 4 1 0 Y Y N Y Y Y Y Y Y Y N Y Y N Y Y Y Y Y Y Y N Y Y Y Table 2: More Simulations. This is a continuation of Extended Data Table 1. Please see the Methods Section for a full description of the columns in this table. (1) Σ0 (2) (3) (4) (5) (6) (7) (8) (9) (10) (11) (12) (13) q smax fpl τ ce pebbles tgen N Gas N Gas f N ice N ice f Disk ok? (g/cm2) (km) (Y/N) (Myr) (Y/N) grow? 900 900 9/7 1500 0.8 0.1 0.66 9/7 1500 0.8 0.1 0.66 1800 9/7 1500 0.4 0.1 0.66 1800 9/7 1500 0.4 0.1 0.66 3600 9/7 1500 0.2 0.1 0.66 3600 9/7 1500 0.2 0.1 0.66 3600 9/7 1500 0.2 0.1 0.66 3600 9/7 1500 0.2 0.6 0.66 3600 9/7 1500 0.2 0.6 0.66 7200 9/7 1500 0.1 0.6 0.66 7200 9/7 1500 0.1 0.6 0.66 7200 9/7 1500 0.1 0.6 0.66 N N N N N N N N N N N N 0 0 5 3 7 6 4 5 2 3 4 5 0 0 4 3 5 4 3 3 2 1 3 3 21 17 8 6 12 9 9 6 11 10 6 5 15 14 0 0 0 0 0 0 4 1 0 1 Y Y N N N N N N Y N N N 0.9* 1.3* 0.9* 1.3* 0.9* 1.3* 0.4* 0.9* 0.4* 0.4* 0.9* 1.3* 24 Extended Data Legends References [22] Levison, H. F., Duncan, M. J. & Thommes, E. A Lagrangian Integrator for Planetary Accretion and Dynamics (LIPAD). Astron. J. 144, 119–138 (2012). [23] Duncan, M. J., Levison, H. F. & Lee, M. H. A Multiple Time Step Symplectic Algorithm for Integrating Close Encounters. Astron. J. 116, 2067–2077 (1998). [24] Benz, W. & Asphaug, E. Catastrophic Disruptions Revisited. Icarus 142, 5–20 (1999). [25] Papaloizou, J. C. B. & Larwood, J. D. On the orbital evolution and growth of proto- planets embedded in a gaseous disc. Astronomy 315, 823–833 (2000). [26] Adachi, I., Hayashi, C. & Nakazawa, K. The gas drag effect on the elliptical motion of a solid body in the primordial solar nebula. Progress of Theoretical Physics 56, 1756–1771 (1976). [27] Rafikov, R. R. Atmospheres of Protoplanetary Cores: Critical Mass for Nucleated Instability. Astrophys. J. 648, 666–682 (2006). [28] Ida, S. & Lin, D. N. C. Toward a Deterministic Model of Planetary Formation. IV. Effects of Type I Migration. Astrophys. J. 673, 487–501 (2008). [29] Dobbs-Dixon, I., Li, S. L. & Lin, D. N. C. Tidal Barrier and the Asymptotic Mass of Proto-Gas Giant Planets. Astrophys. J. 660, 791–806 (2007). [30] Lambrechts, M., Johansen, A. & Morbidelli, A. Separating gas-giant and ice-giant planets by halting pebble accretion. Astron. Astrophys. 572, A35 (2014). 25 [31] Chambers, J. E. Giant planet formation with pebble accretion. Icarus 233, 83–100 (2014). [32] Nakagawa, Y., Sekiya, M. & Hayashi, C. Settling and growth of dust particles in a laminar phase of a low-mass solar nebula. Icarus 67, 375–390 (1986). [33] Andrews, S. M., Wilner, D. J., Hughes, A. M., Qi, C. & Dullemond, C. P. Protoplanetary Disk Structures in Ophiuchus. II. Extension to Fainter Sources. Astrophys. J. 723, 1241–1254 (2010). [34] Hayashi, C. Structure of the Solar Nebula, Growth and Decay of Magnetic Fields and Effects of Magnetic and Turbulent Viscosities on the Nebula. Progress of Theoretical Physics Supplement 70, 35–53 (1981). [35] Hern´andez, J. et al. Spitzer Observations of the λ Orionis Cluster. I. The Frequency of Young Debris Disks at 5 Myr. Astrophys. J. 707, 705–715 (2009). [36] Chiang, E. I. & Goldreich, P. Spectral Energy Distributions of T Tauri Stars with Passive Circumstellar Disks. Astrophys. J. 490, 368–376 (1997). [37] Morbidelli, A., Bottke, W. F., Nesvorn´y, D. & Levison, H. F. Asteroids were born big. Icarus 204, 558–573 (2009). [38] Lodders, K. Solar System Abundances and Condensation Temperatures of the Elements. Astrophys. J. 591, 1220–1247 (2003). [39] Shakura, N. I. & Sunyaev, R. A. Black holes in binary systems. Observational appear- ance. Astron. Astrophys. 24, 337–355 (1973). 26
0912.2076
2
0912
2010-10-10T17:02:41
Detectability of extrasolar moons as gravitational microlenses
[ "astro-ph.EP" ]
We evaluate gravitational lensing as a technique for the detection of extrasolar moons. Since 2004 gravitational microlensing has been successfully applied as a detection method for extrasolar planets. In principle, the method is sensitive to masses as low as an Earth mass or even a fraction of it. Hence it seems natural to investigate the microlensing effects of moons around extrasolar planets. We explore the simplest conceivable triple lens system, containing one star, one planet and one moon. From a microlensing point of view, this system can be modelled as a particular triple with hierarchical mass ratios very different from unity. Since the moon orbits the planet, the planet-moon separation will be small compared to the distance between planet and star. Such a configuration can lead to a complex interference of caustics. We present detectability and detection limits by comparing triple-lens light curves to best-fit binary light curves as caused by a double-lens system consisting of host star and planet -- without moon. We simulate magnification patterns covering a range of mass and separation values using the inverse ray shooting technique. These patterns are processed by analysing a large number of light curves and fitting a binary case to each of them. A chi-squared criterion is used to quantify the detectability of the moon in a number of selected triple-lens scenarios. The results of our simulations indicate that it is feasible to discover extrasolar moons via gravitational microlensing through frequent and highly precise monitoring of anomalous Galactic microlensing events with dwarf source stars.
astro-ph.EP
astro-ph
Astronomy&Astrophysicsmanuscript no. 13844 May 30, 2018 c(cid:13) ESO 2018 0 1 0 2 t c O 0 1 . ] P E h p - o r t s a [ 2 v 6 7 0 2 . 2 1 9 0 : v i X r a Detectability of extrasolar moons as gravitational microlenses Christine Liebig and Joachim Wambsganss Astronomisches Rechen-Institut, Zentrum fur Astronomie der Universitat Heidelberg, Monchhofstrasse 12-14, 69120 Heidelberg, Germany, e-mail: [email protected], [email protected] Received 10 December 2009 / Accepted 25 April 2010 ABSTRACT We evaluate gravitational lensing as a technique for the detection of extrasolar moons. Since 2004 gravitational microlensing has been successfully applied as a detection method for extrasolar planets. In principle, the method is sensitive to masses as low as an Earth mass or even a fraction of it. Hence it seems natural to investigate the microlensing effects of moons around extrasolar planets. We explore the simplest conceivable triple lens system, containing one star, one planet and one moon. From a microlensing point of view, this system can be modelled as a particular triple with hierarchical mass ratios very different from unity. Since the moon orbits the planet, the planet-moon separation will be small compared to the distance between planet and star. Such a configuration can lead to a complex interference of caustics. We present detectability and detection limits by comparing triple-lens light curves to best-fit binary light curves as caused by a double-lens system consisting of host star and planet – without moon. We simulate magnification patterns covering a range of mass and separation values using the inverse ray shooting technique. These patterns are processed by analysing a large number of light curves and fitting a binary case to each of them. A chi-squared criterion is used to quantify the detectability of the moon in a number of selected triple-lens scenarios. The results of our simulations indicate that it is feasible to discover extrasolar moons via gravitational microlensing through frequent and highly precise monitoring of anomalous Galactic microlensing events with dwarf source stars. Key words. gravitational lensing – planetary systems – methods: numerical – methods: statistical 1. Introduction By now hundreds of extrasolar planets have been detected1. For all we know, none of the newly discovered extrasolar planets offers physical conditions permitting any form of life. But the search for planets potentially harbouring life and the search for indicators of habitability is ongoing. One of these indicators might be the presence of a large natural satellite – a moon – which stabilises the rotation axis of the planet and thereby the surface climate (Benn 2001). It has also been suggested that a large moon itself might be a good candidate for offering habit- able conditions (Scharf 2006). In the solar system, most planets harbour moons. In fact, the moons in the solar system outnum- ber the planets by more than an order of magnitude. No moon has yet been detected around an extrasolar planet. The majority of known exoplanets has been discovered through radial velocity measurements, with the first success- ful finding reported by Mayor & Queloz (1995). This method is not sensitive to satellites of those planets, because the stellar "Doppler wobble" is only affected by the orbital movement of the barycentre of a planet and its satellites, though higher-order effects could play a role eventually. Here we consider Galactic microlensing, which has led to the discovery of several relatively low-mass exoplanets since the first report of a successful de- tection by Bond et al. (2004), as a promising technique for the search for exomoons. As early as 1999, it has been suggested that extraso- lar moons might be detectable through transit observations (Sartoretti & Schneider 1999), either through direct observa- tion of lunar occultation or through transit timing variations, as the moon and planet rotate around their common barycen- 1 exoplanet.eu tre, causing time shifts of the transit ingress and egress (cf. also Holman & Murray (2005)). In their simulations of space-based gravitational microlensing Bennett & Rhie (2002) mention the possibility of discovering extrasolar moons similar to our own Moon. Later that year, Han & Han (2002) performed a detailed feasibility study whether microlensing offers the potential to discover an Earth-Moon analogue, but concluded that finite source effects would probably be too severe to allow detections. Williams & Knacke (2004) published the quite original sugges- tion to look for spectral signatures of Earth-sized moons in the absorption spectra of Jupiter-sized planets. Cabrera & Schneider (2007) proposed a sophisticated transit approach using "mu- tual event phenomena", i.e. photometric variation patterns due to different phases of occultation and light reflection of planet and satellite. Han (2008) undertook a new qualitative study of a number of triple-lens microlensing constellations finding "non- negligible" light curve signals to occur in the case of an Earth- mass moon orbiting a 10 Earth-mass planet, "when the planet- moon separation is similar to or greater than the Einstein ra- dius of the planet". Lewis et al. (2008) analysed pulsar time-of- arrival signals for lunar signatures. Kipping (2009a,b) refined and extended the transit timing models of exomoons to include transit duration variations, and Kipping et al. (2009) examined transit detectability of exomoons with Kepler-class photometry and concluded that in optimal cases moon detections down to 0.2 MEarth should be possible. We cover here several new aspects concerning the microlens- ing search for exomoons, extending the work of Han (2008). First, the detectability of lunar light curve perturbations is de- termined with a statistical significance test that does not need to rely on human judgement. Second, all parameters of the two- dimensional three-body geometry, including the position angle 2 C. Liebig & J. Wambsganss: Exomoons as gravitational lenses of the moon with respect to the planet-star axis, are varied. Third, an unbiased extraction of light curves from the selected scenarios enables a tentative prediction of the occurrence rate of detectable lunar light curve signals. A more detailed account of this study is available as Liebig (2009). The paper is structured as follows: In Section 2, we recall the fundamental equations of gravitational microlensing relevant to our work. Our method for quantifying the detection rates for extrasolar moons in selected lensing scenarios is presented in Section 3. In Section 4, we discuss the astrophysical implica- tions of the input parameters of the simulations. Our results are presented in Section 5, together with a first interpretation and a discussion of potential problems of our method. 2. Basics of gravitational microlensing The deflection of light by massive bodies is a consequence of the theory of general relativity (Einstein 1916) and has been experi- mentally verified since 1919 (Dyson et al. 1920), see Paczy´nski (1996) for an introduction to the field or Schneider et al. (2006) for a comprehensive review. The typical scale of angular separations in gravitational lens- ing is the Einstein radius θE, the angular radius of the ring of for- mally infinite image magnification that appears when a source at a distance DS , a lens of mass M at a distance DL, and the ob- server are perfectly aligned: θE = r 4GM c2 DLS DLDS , (1) where G denotes the gravitational constant, c the speed of light. DLS is the distance between lens plane and source plane; in the non-cosmological distance scale of our galaxy DLS = DS − DL holds true. With the Einstein radius also comes the characteristic time scale of transient gravitational lensing events, the Einstein of the lens time tE = DLθE /v relative to the source. with the transverse velocity v ⊥ ⊥ Here we focus on the triple lens case with host star (S ), planet (P) and moon (M), see Figure 1 for illustration. The lens equation can be expressed using complex coordinates, where η shall denote the angular source position and ξ the image po- sitions, cf. choice of notation in Witt (1990) and Gaudi et al. (1998). ξi stands for the angular position of the lensing body i. qi j is the mass ratio between lenses i and j (qi j = Mi ). The lens M j equation gets the following form, if the primary lens, the host star S , has unit mass and is placed in the origin of the lens plane, η = ξ − 1 ξ − qPS ξ − ξP − qMS ξ − ξM . (2) This is a mapping from the lens plane to the source plane, which maps the images of a source star to its actual position in the source plane. As pointed out in Rhie (1997), and explicitly cal- culated in Rhie (2002), the triple-lens equation is a tenth-order polynomial equation in ξ. In microlensing, the images cannot be resolved. Detectable is only the transient change in magnitude of the source star, when lens and source star are in relative motion to each other. The total magnification µ of the base flux is obtained as the inverse of the determinant of the Jacobian of the mapping equation (2), µ = 1 det J(ξ) , with det J(ξ) = 1 − ∂η ∂η ∂ξ ∂ξ . (3) Host Star θPS Moon θMP φ Planet Fig. 1. Geometry of our triple-lens scenario, not to scale. Five parameters have to be fixed: The mass ratios qPS = MPlanet/MStar and qMP = MMoon/MPlanet, the angular separations in the lens plane θPS and θMP and the position angle of the moon φ. These parameters define uniquely the relative projected positions of the three bodies. Gravitational lensing changes the apparent solid angle of a source, not the surface brightness. The magnification µ is the ratio of the total solid angle of the images and the apparent solid angle of the unlensed source. For a status report of the past, present and prospective fu- ture of planet searching via Galactic microlensing the recent white papers of 2008/2009 are a good source of reference (Beaulieu et al. 2008; Dominik et al. 2008; Bennett et al. 2009; Gaudi et al. 2009). 3. Method The simplest gravitational lens system incorporating an extraso- lar moon is a triple-lens system consisting of the lensing star, a planet and a moon in orbit around that planet, as sketched in Figure 1. Most likely, lunar effects will first show up as no- ticeable irregularities in light curves that have been initially ob- served and classified as light curves with planetary signatures. To measure the detectability of a given triple-lens system among binary lenses, we have to determine whether the provided triple- lens light curve differs statistically significantly from binary-lens light curves. To clarify the terminology: We investigate light curves that show a deviation from the single-lens case due to the presence of the planetary caustic, i.e. which would be modelled as a star- plus-planet system in a first approximation. We call lunar de- tectability the fraction of those light curves which display a significant deviation from a star-plus-planet lens model due to the presence of the moon. We measure the difference between a given triple-lens light curve (which is taken to be the "true" underlying light curve of the event) and its best-fit binary-lens counterpart. The best-fit binary-lens light curve is found by a least-square fit. We then employ χ2-statistics to see whether the triple-lens light curve could be explained as a normal fluctuation within the error boundaries of the binary-lens light curve. If this is not the case, the moon is considered detectable. Detection and characterisation are two separate problems in the search for extrasolar planets or moons, though, and we do not make statements about the latter. When characterising an ob- served light curve with clear deviations from the binary model, it is still possible that ambiguous solutions – lunar and non-lunar – arise. This does not reflect on our results. We simply give the fraction of triple-lens light curves significantly deviating from binary-lens light curves, without exploring whether they can be uniquely characterised. A qualitative impression of the lunar influence on the caus- tic structure can be gained from the magnification patterns in Figure 2. Extracted example light curves are presented in Figure 3. Going beyond the qualitative picture, in this paper we quantify the detectability of an extrasolar moon in gravitational microlensing light curves in selected scenarios. C. Liebig & J. Wambsganss: Exomoons as gravitational lenses 3 a d g j b e h k c f i l Fig. 2. Details of the analysed triple-lens magnification maps for an example mass/separation scenario with qPS = 10−3, qMP = 10−2, θPS = 1.3 θE and θMP = 1.0 θP E (the standard scenario as summarised in Table 1), showing the caustic configurations for twelve different lunar positions completing a full circular orbit in steps of 30◦. The relative positions of star, planet and moon are sketched in the lower left of each panel (not to scale). The side lengths of the individual frames are 0.1 θE. A darker shade of grey corresponds to a higher magnification. The straight black lines mark source trajectories, the resulting light curves are displayed in Figure 3. 4 a 2 1.5 g a m ∆ − 1 0.5 d 2 1.5 g a m ∆ − 1 0.5 C. Liebig & J. Wambsganss: Exomoons as gravitational lenses b 2 1.5 1 0.5 c 2 1.5 1 0.5 0.02 0.04 0.06 0.08 0.02 0.04 0.06 0.08 0.02 0.04 0.06 0.08 e 2 1.5 1 0.5 f 2 1.5 1 0.5 0.02 0.04 0.06 0.08 0.02 0.04 0.06 0.08 0.02 0.04 0.06 0.08 2 g 2 h 1.5 g a m ∆ − 1 0.5 1.5 1 0.5 2 i 1.5 1 0.5 0.02 0.04 0.06 0.08 0.02 0.04 0.06 0.08 0.02 0.04 0.06 0.08 2 j 2 k 1.5 g a m ∆ − 1 0.5 1.5 1 0.5 2 l 1.5 1 0.5 0.02 0.04 0.06 0.08 0.02 0.04 0.06 0.08 0.02 0.04 0.06 0.08 t/tE t/tE t/tE Fig. 3. Sample of triple-lens light curves corresponding to the source trajectories depicted in Figure 2. The solid light curves were extracted with an assumed solar source size, RSource = R⊙; the thin, dashed lines show the same light curves with RSource = 3 R⊙. The magnification scale is given in negative magnitude difference (−∆mag), i.e. the unmagnified baseline flux of the source is 0. The time scale in units of the Einstein time tE, assuming a uniform relative motion of source and lens. C. Liebig & J. Wambsganss: Exomoons as gravitational lenses 5 3.1. Rayshooting is analytically solvable, but The triple-lens equation (2) Han & Han (2002) pointed out that numerical noise in the poly- nomial coefficients caused by limited computer precision was too high (∼ 10−15) when solving the polynomial numerically for the very small mass ratios of moon and star (∼ 10−5). To avoid this, we employ the inverse ray-shooting technique, which has the further advantage of being able to account for finite source sizes and non-uniform source brightness profiles more easily. It also gives us the option of incorporating additional lenses (further planets or moons) without increasing the com- plexity of the calculations. This technique was developed by Schneider & Weiss (1986), Kayser et al. (1986) and Wambsganss (1990, 1999) and already applied to planetary microlensing in Wambsganss (1997). Inverse ray-shooting means that light rays are traced from the observer back to the source plane. This is equivalent to trac- ing rays from the background source star to the observer plane. The influence of all masses in the lens plane on the light path is calculated. In the thin lens approximation, the deflection angle is just the sum of the deflection angles of every single lens. After deflection, all light rays are collected in pixel bins of the source plane. Thus a magnification pattern is produced (e.g. Figure 4). Lengths in this map can be translated to angular separations or, assuming a constant relative velocity between source and lens, to time intervals. The number of collected light rays per pixel is proportional to the magnification µ of a background source with pixel size at the respective position in the source plane. The res- olution in magnification of these numerically produced patterns is finite and depends on the total number of rays shot. 3.2. Lightcurvesimulation We obtain simulated microlensing light curves with potential traces of a moon by producing magnification maps (Figure 4) of triple-lens scenarios with mass ratios very different from unity. A light curve is then obtained as a one-dimensional cut through the magnification pattern, convolved with the luminosity profile of a star. Only a few more assumptions are necessary to simulate re- alistic, in principle observable, light curves: angular source size, relative motion of lens and source, and lens mass. As a first ap- proximation to the surface brightness profiles of stars we use a profile with radius RSource and constant surface brightness (ig- noring limb-darkening effects). The light curve is sampled at equidistant intervals. In Section 4.7 we discuss physical impli- cations of the sampling frequency. In order to be able to make robust statistical statements about the detectability of a moon in a chosen setting, i.e. a certain magnification pattern, we anal- yse a grid of source trajectories that delivers about two hundred light curves, see also Figure 5. To have an unbiased sample, the grid is chosen independently of lunar caustic features, but all tra- jectories are required to show pronounced deviations from the single-lens light curve due to the planet. Fig. 4. Magnification pattern displaying the planetary caustic of a triple-lens system with mass ratios qPS = 10−3 and qMP = 10−2, corresponding, e.g., to an M-dwarf, a Saturn-mass planet and an Earth-mass satellite of that planet. Darker shades corre- spond to areas of higher magnification. The third body shows its influence as the small perturbation on the bottom cusp of the planetary caustic. The position angle of the moon is φ = 90◦, i.e. in the orientation of this map the moon is located above the planet. The complete set of parameters is given in Table 1. The side length of the pattern is 0.1 θE in stellar Einstein radii. A source with radius R⊙ (2R⊙, 3R⊙, 5R⊙) is indicated as a black disc (dashed circles) in the lower left. Taking the straight line that cuts the lunar perturbation of the caustic as the source tra- jectory of a solar sized source results in the light curve shown in Figure 6 (solid line). Depending on the assumed source size, different light curves are obtained, see also Figure 9(a) to (c). ditional caustic features, it also slightly moves the location of the double-lens caustics. Since the perturbation introduced by the moon is small, however, we have good starting conditions using the light curve from the binary magnification pattern that corresponds to the triple-lens magnification pattern. Keeping the planet-star separation fixed and changing the planetary mass to the sum of the masses of planet and moon, we are likely to be close to the global best-fit values. In our simplified approach, we only vary the source trajectory to search for the best-fitting binary-lens light curve, which already yields satisfactory results, as can be seen in Figure 6(b). For the fitting procedure we use the simple, straight-forward and robust least-square method. 3.3. Fitting 3.4. Lightcurvecomparison For each triple-lens scenario, we produce an additional magni- fication pattern of a corresponding binary-lens system, where the mass of the moon is added to the planetary mass. As a first approximation, one can compare two light curves with identi- cal source track parameters, see Figure 6(a). Numerically big differences can occur without a significant topological differ- ence, because the additional third body not only introduces ad- We want to use the properties of the χ2-distribution for a test of significance of deviation between the two simulated light curves of the triple lens and the corresponding binary lens. Most commonly, the χ2-distribution is used to test the goodness-of-fit of a model to experimental data. To compare two simulated curves instead, as in our case, one possible approach is to randomly generate artificial data around one of them and then 6 C. Liebig & J. Wambsganss: Exomoons as gravitational lenses Triple Lens Binary Lens Difference 0 0.01 0.02 0.03 0.04 t/tE 0.05 0.06 0.07 0.08 (a) Identical source trajectory. Triple Lens Binary Lens Difference g a m ∆ − g a m ∆ − 1.6 1.4 1.2 1 0.8 0.6 0.4 0.2 0 -0.2 1.6 1.4 1.2 1 0.8 0.6 0.4 0.2 0 -0.2 0 0.01 0.02 0.03 0.04 t/tE 0.05 0.06 0.07 0.08 (b) Best-fit source trajectory. Fig. 6. Triple-lens light curve (solid line) extracted from the magnification pattern in Figure 4, compared with a light curve extracted from the corresponding binary-lens magnification map (dashed line), with no third body and the planet mass increased by the previous moon mass. The difference of the two light curves is plotted as the dotted line. The central peak of the triple- lens light curve is caused by lunar caustic perturbations, that cannot easily be reproduced with a binary lens. In (a), identical source trajectory parameters are used to extract the binary-lens light curve and large residuals remain that can be avoided with a different source trajectory. In (b), the best-fit binary-lens light curve is shown (dashed) in comparison with the triple-lens light curve (solid). Here, the residuals can be attributed to the moon. If the deviation is significant, the moon is detectable in the triple- lens light curve. A source with solar radius at a distance of 8 kpc is assumed. χ2-distributed and hQ2i in particular will lie outside the expected range for χ2-distributed random variables. In the latter case, the moon is detectable. This approach is presented in mathematical detail in Appendix A. 4. Choice of scenarios This section presents the assumptions that we use for our sim- ulations. We discuss the astrophysical parameter space that is available for simulations of a microlensing system consisting of Fig. 5. Illustration of how the magnification pattern of a given star-planet-moon configuration is evaluated. In order to have an unbiased statistical sample, the source trajectories are chosen in- dependently of the lunar caustic features, though all light curves are required to pass through or close to the planetary caustic. This is realised through generating a grid of source trajectories that is only oriented at the planetary caustic. In the actual evalua- tion a 10 times denser grid than shown here was used. The mag- nification map has to be substantially larger than the caustic to ensure an adequate baseline for the light curve fitting. The side length of this pattern is 0.3 θE, the position angle of the moon φ = 90◦, all other parameters are chosen as listed in Table 1. Source sizes of 1 R⊙, 2 R⊙, 3 R⊙, and 5 R⊙ are indicated in the lower left. consecutively fit the two theoretical curves to the artificial data with some free parameters. Two 'χ2-values' Q2 2 will re- sult. Their difference, ∆Q2 = Q2 2 can then be calculated and a threshold value ∆Q2 thresh is chosen. The condition for reported detectability of the deviation is then ∆Q2 > ∆Q2 thresh. A detailed description of the algorithm and its application to microlensing was presented by Gaudi & Sackett (2000). 1 and Q2 1 − Q2 We decided to use a different approach here: In particular, we were led to consider other possible techniques by the wish to avoid "feeding" our knowledge of the data distribution to a ran- dom number generator in order to get an unbiased and random χ2-distributed sample of Q2, when at the same time we have all the necessary information to calculate a much more representa- tive χ2-value. We have developed a method to quantify signifi- cance of deviations between two theoretical functions that avoids the steps of data simulation and subsequent fitting. Put simply, we calculate the mean hQ2i of all 'χ2-values' that would result, if data with a Gaussian scatter drawn from a triple-lens light curve is fitted with a binary light curve. If the triple-lens case and the best-fit binary-lens case are indistin- guishable, then Q2 will be χ2-distributed and hQ2i will lie at or very close to hχ2i = number of degrees of freedom, the mean of the χ2 probability density function. If the triple-lens light curve and the binary-lens light curve differ significantly, then Q2 is not C. Liebig & J. Wambsganss: Exomoons as gravitational lenses 7 star, planet and moon. By choosing the most probable or most pragmatic value for each of the parameters, we create a standard scenario, summarised in Table 1 and visualised in Figure 10, that all other parameters are compared against during the analysis. The observational search for microlensing events caused by extrasolar planets is carried out towards the Galactic bulge, currently limited to the fields of the wide-field surveys OGLE (Udalski et al. 1992) and MOA (Muraki et al. 1999). This leads to some natural assumptions for the involved quantities. The source stars are typically close to the centre of the Galactic bulge at a distance of DS = 8 kpc, where the surface density of stars is very high. For the distance to the lens plane we adopt a value of DL = 6 kpc. We assume our primary lens mass to be an M- dwarf star with a mass of MStar = 0.3M⊙, because that is the most abundant type, cf. Figure 5 of Dominik et al. (2008). Direct lens mass determination is only possible if additional observ- ables, such as parallax, can be measured, cf. Gould (2009). We derive the corresponding Einstein radius of θE(DS , DL, MStar) = 0.32 milliarcseconds. Five parameters describe our lens configuration, cf. Figure 1: They are the mass ratios between planet and star qPS and be- tween moon and planet qMP, the angular separations θPS be- tween planet and star and θMP between moon and planet, and as the last parameter φ, the position angle of the moon with respect to the planet-star axis. These parameters are barely constrained by the physics of a three-body system, even if we do require mass ratios very different from unity and separations that allow for stable orbits. The apparent source size RSource and the sampling rate fsampled affect the shape of the simulated light curves. We have to define an observational uncertainty σ for the significance test in the final analysis of the light curves. 4.1. Massratioofplanetandstar qPS We are interested in planets (and not binary stars). Accordingly, we want a small value for the mass ratio of planet and star qPS = MPlanet . Our standard value will be 10−3 which is the mass MStar ratio of Jupiter and Sun, or a Saturn-mass planet around an M- dwarf of 0.3 M⊙. At a given projected separation between star and planet, this mass ratio determines the size of the planetary caustic, with a larger qPS leading to a larger caustic. qPS is var- ied to also examine scenarios with mass ratios corresponding to a Jovian mass around an M-dwarf and to a Saturn mass around the Sun. Summarised: For the mass ratio between planet and star we use the three values qPS = 3.3 × 10−3, 10−3, 3 × 10−4. 4.2. MassratioofmoonandplanetqMP To be classified as a moon, the tertiary body must have a mass considerably smaller than the secondary. The standard case in our examination corresponds to the Moon/Earth mass ratio of MMoon = 10−2. We are generous towards the higher mass end, MEarth and include a mass ratio of 10−1 in our analysis, corresponding to the Charon/Pluto system. Both examples are singular in the solar system, but we argue that a more massive moon is more interesting, since it can effectively stabilise the obliquity of a planet, which is thought to be favouring the habitability of the planet (Benn 2001). At the low mass end of our analysis we ex- amine qMP = 10−3. In Figure 7, three different caustic interfer- ences resulting from the three adopted mass ratios are shown. Summarised: For the mass ratio between moon and planet we use the three values qMP = 10−3, 10−2, 10−1. (a) qMP = 10−3 (b) qMP = 10−2 (c) qMP = 10−1 Fig. 7. Caustic topology with different lunar mass ratios. From left to right, the lunar mass ratio increases from qMP = 10−3 to 10−1. The side length of the displayed patterns is 0.05 θE. All other parameters as given in Table 1, in particular qPS = 10−3 and θMP = 1.0 θP E. Source sizes of 1 R⊙ to 5 R⊙ are indicated in the lower left. 4.3. Angularseparationofplanetandstar θPS The physical separation of a planet and its host star, i.e. the semi-major axis aPS , is not directly measurable in gravitational lensing. Only the angular separation θPS can directly be inferred from an observed light curve, where 0 ≤ θPS DL ≤ aPS is valid for a given physical separation aPS at a lens distance DL. The angular separation θPS of a binary will evoke a certain topology of caustics, illustrated e.g. in Figure 1 of Cassan (2008). They gradually evolve from the close separation case with two small triangular caustics on the far side of the star and a small central caustic at the star position, to a large central caustic for the inter- mediate case, when the planet is situated near the stellar Einstein ring, θPS = 1.0 θE. This is also called the "resonant lensing" case, as described in Wambsganss (1997). If the planet is moved fur- ther out, one obtains a small central caustic and a larger isolated, roughly kite-shaped planetary caustic that is elongated towards the primary mass in the beginning, but becomes more and more symmetric and diamond-shaped if the planet is placed further outwards. In analogy, a "lunar resonant lensing zone" can be defined, when measuring the angular separation of moon and planet in E = √qPS θE. Viewing our so- units of planetary Einstein radii θP lar system as a gravitational lens system at DL = 6 kpc and with DS = 8 kpc, Jupiter's Einstein radius in physical lengths would be 0.11 AU in the lens plane, that is almost 10 times the semi- major axis of Callisto. However, the smaller moons of Jupiter cover the region out to 2 Jovian Einstein radii. If there is a moon E . θMP . 3.0 θP at an angular separation of the planet of 0.6 θP E, it will show its influence in any of the caustics topologies. But we focus our study on the wide-separation case for the follow- ing reasons: Regarding the close-separation triangular caustics, we argue that the probability to cross one of them is vanish- ingly small and, furthermore, the magnification substantially de- creases as they move outwards from the star. The intermediate or resonant caustic is well "visible" because it is always located close to the peak of the single-lens curve. But all massive bodies of a given planetary system affect the central caustic with mi- nor or major perturbations and deformations (Gaudi et al. 1998). There is no reason to expect that extrasolar systems are gener- ally less densely populated by planets or moons than the solar system. Since we are looking for very small deviations, identi- fying the signature of the moon among multiple features caused by low-mass planets and possibly more moons would be increas- ingly complex. The wide-separation case is most favourable, be- cause the planetary caustic is typically caused by a single planet and interactions due to the close presence of other planets are 8 C. Liebig & J. Wambsganss: Exomoons as gravitational lenses highly unlikely (Bozza 1999, 2000). Any observed interferences can therefore be attributed to satellites of the planet. Planets with only one dominant moon are at least common in the solar sys- tem (Saturn & Titan, Neptune & Triton, Earth & Moon). Such a system would be the most straightforward to detect in real data. This makes a strong case for concentrating on the wide- separation planetary caustic. As the standard we choose a sepa- ration of θPS = 1.3 θE and also test a scenario with θPS = 1.4 θE. For comparison: the maximum projected separation of Jupiter at the chosen distances DS and DL corresponds to 1.5 solar Einstein radii. Summarised: For the angular separation between planet and star we use the two values θPS = 1.3θE , 1.4θE. 4.4. Angularseparationofmoonandplanet θMP As the moon by definition orbits the planet, there is an upper limit to the semi-major axis of the moon aMP. It must not exceed the distance between the planet and its inner Lagrange point. This distance is called Hill radius and is calculated as rHill = aPS MP 3MS ! 1 3 for circular orbits, with the semi-major axis aPS of the plane- tary orbit. This can be translated to "lensing parameters", but the equation changes into the inequality rHill ≥ θPS DL 1 3 qPS! 1 3 , (4) because for a physical distance aPS , the apparent (i.e. projected) separation lies in the range 0 ≤ θPS DL ≤ aPS depending on the inclination towards the line of sight. Thus, the Hill radius con- straint aMP ≤ rHill cannot be regarded as a strict limit. Though, turning this argument around, a minimum region of secured sta- bility exists for given mass ratios and angular planet-star separa- tion θPS . Bodies in prograde motion with an orbit below 0.5 rHill can have long term stability, for retrograde motion the limit is somewhat higher at 0.75 rHill (see Domingos et al. (2006) and references therein, particularly Hunter (1967)). The region of se- cured stability for our standard scenario has a radius of θMP ≤ 0.5 × θPS DL 1 3 qPS! 1 3 = 0.045 θE = 1.53 θP E, E . θMP . 3.0 θP which leads to aMP = 0.09AU in our setting. Our choice of θMP is also motivated by a desire to have caustic interactions between lunar and planetary caustic, a resonance, because only under this condition the moon will give rise to light curve fea- tures that are distinctly different from those of a low-mass planet. If the moon is located at an angular separation from the planet of – depending on the position angle of the moon – roughly 0.6 θP E, it will influence the planetary caus- tic. For angular separations larger than 3.0 θP E, the lunar caustic barely influences the planetary caustic, and if it is detected, it will be detected as a distinct secondary caustic. Han (2008) al- ready pointed out that the lunar signal does not vanish with an increased separation from the planet. But the lunar mass might be wrongly identified as a second, lower-mass planet rather than a satellite of the first planet. For comparison, the major moons of the solar system are all at considerably smaller angular sep- arations than 1.0 θP E, if observed as a gravitational lens system from DL = 6 kpc with DS = 8 kpc, with our Moon being most favourable at 0.4 of Earth's Einstein radii at maximum projected (a) θMP = 0.8 θP E (b) θMP = 1.0 θP E (c) θMP = 1.2 θP E Fig. 8. Magnification patterns with side lengths of 0.05 θE show- ing the lunar perturbation of the planetary caustic for in- creasing angular separations of moon and planet of θMP = 0.8 θP E, while all other parameters are as in Table 1. The topology of interaction varies with the angular separation. Source sizes of 1 R⊙ to 5 R⊙ are indicated in the lower left. E, 1.2 θP E, 1.0 θP separation. But as we have learnt through the discovery of hun- dreds of exoplanets, we do not have to expect solar system prop- erties to be mirrored in exoplanetary systems. We focus on configurations with strong interferences of plan- etary and lunar caustic, what one might call the "resonant case of lunar lensing", examples of which are displayed in Figure 8. Therefore, our standard case will have an angular separation of moon and planet of one planetary Einstein radius. In total, we evaluate four settings: θMP = 0.8θP E, in the chosen setting (Table 1), this corresponds to physical projected separations from 0.5 to 0.8 AU. E, 1.2θP E, 1.0θP E, 1.4θP 4.5. Positionangleofmoon φ The fifth physical parameter for determining the magnification patterns is the position angle of the moon φ with respect to planet-star axis, as depicted in Figure 1. It is the only parame- ter that is not fixed in our standard scenario. Instead, we eval- uate this parameter in its entirety by varying it in steps of 30◦ to complete a full circular orbit of the moon around the planet. By doing this, we are aiming at getting complete coverage of a selected mass/separation scenario. Figure 2 visualises how the position of the moon affects the planetary caustic. It is not to be expected that we will ever have an exactly frontal view of a per- fectly circular orbit, but it serves well as a first approximation. Summarised: we choose 12 equally spaced values for the posi- tion angle φ of the moon relative to the planet-star axis, and the results for the various parameter sets are averaged over these 12 geometries. 4.6. Sourcesize RSource The source size strongly influences the amplitudes of the light curve fluctuations and the "time resolution". Larger sources blur out finer caustic structures, compare Figure 9. For our source star assumptions, we have to take into account not only stel- lar abundances, but also the luminosity of a given stellar type. Main sequence dwarfs are abundant in the Galactic bulge, but they are faint, with apparent magnitudes of ∼ 15 mag to 25 mag at DS = 8 kpc. In the crowded microlensing fields, they are difficult to observe with satisfying photometry by ground- based telescopes, even if they are lensed and magnified. Giant stars, with apparent baseline magnitudes ∼ 13 mag to 17 mag, are more likely to allow precise photometric measurements from ground and, therefore, today's follow-up observations are biased towards these source stars. C. Liebig & J. Wambsganss: Exomoons as gravitational lenses 9 The finite source size constitutes a serious limitation to the discovery of extrasolar moons. In fact, Han & Han (2002) stated that detecting satellite signals in the lensing light curves will be close to impossible, because the signals are smeared out by the severe finite-source effect. They tested various source sizes (and planet-moon separations) for an Earth-mass planet with a Moon-mass satellite. They find that even for a K0-type source star (RSource = 0.8 R⊙) any light curve modifications caused by the moon are washed out. Han (2008) increased lens masses and by assuming a solar source size RSource = R⊙, he finds that "non- negligible satellite signals occur" in the light curves of planets of 10 to 300 Earth masses, "when the planet-moon separation is similar to or greater than the Einstein radius of the planet" and the moon has the mass of Earth. We use a solar sized source as our standard case, and present results for four more source radii: RSource = 1R⊙, 2R⊙, 3R⊙, 5R⊙, 10R⊙. In terms of stellar Einstein radii this corresponds to five different source size set- tings from 1.8 × 10−3 θE to 18 × 10−3 θE. The ability to monitor smaller sources – dwarfs, low-mass main sequence stars – with good photometric accuracy is one of the advantages of space-based observations. Hence a satellite mission is the ideal tool for this aspect of microlensing, rou- tine detections of moons around planets can be expected with a space-based monitoring program on a dedicated satellite. 4.7. Samplingrate fsampled Typical exposure times for the microlensing light curves are 10 to 300 seconds at a 1.5m telescope. The frequency of observa- tions of an individual interesting event varies between once per 2 hours to once every few minutes, for a single observing site only limited by exposure time and read-out time of the instrument. One example of this is the peak coverage of MOA-2007-BLG- 400, see Dong et al. (2009), where frames were taken every six minutes. Higher observing frequency equals better coverage of the resulting light curve. A normal microlensing event is seen as a transient brightening that can last for a few days to weeks or even months. A planet will alter the light curve from fraction of a day to about a week; the effect of a low-mass planet or a moon will only last for a few hours or shorter, see e.g. Beaulieu et al. (2006). This duration is inversely proportional to the transverse with which the lens moves relative to the line of sight velocity v between the background source or the relative proper motion µ⊥ and the lensing system. We fix this at a typical v ⊥ = 200 km/s for DL = 6 kpc or µ⊥ = 7.0 mas/year. For the simulations, a realistic, non-continuous observing rate is mimicked by evaluating the simulated light curve at in- tervals that correspond to typical frequencies, but allowing for a constant sampling rate, which in practise is often prohibited by observing conditions. We sample the triple-lens light curve in equally sized steps. With the assumed constant relative veloc- ity, this translates to equal spacing in time. Aiming for a rate of fsampled ≃ 1 frame 15 minutes , we choose a step size of 6 × 10−4 θE. To see the effect of a decreased sampling rate, we have also examined the standard scenario (see Table 1) with sampling rates of factors of 1.5 and 2 longer. Summarised: we use three different sampling rates, fsampled = 1/15 min, 1/22 min, 1/30 min. ⊥ 4.8. Assumedphotometricuncertainty σ The standard error σ of a data point in a microlensing light curve can vary significantly. Values between 5 and 100 mmag seem realistic, see, e.g., data plots in the recent event analyses Triple Lens Binary Lens Difference 0 0.01 0.02 0.03 0.04 t/tE 0.05 0.06 0.07 0.08 (a) Rsource = 2R⊙. Triple Lens Binary Lens Difference 0 0.01 0.02 0.03 0.04 t/tE 0.05 0.06 0.07 0.08 (b) Rsource = 3R⊙. Triple Lens Binary Lens Difference g a m ∆ − g a m ∆ − g a m ∆ − 1.6 1.4 1.2 1 0.8 0.6 0.4 0.2 0 -0.2 1.6 1.4 1.2 1 0.8 0.6 0.4 0.2 0 -0.2 1.6 1.4 1.2 1 0.8 0.6 0.4 0.2 0 -0.2 0 0.01 0.02 0.03 0.04 t/tE 0.05 0.06 0.07 0.08 (c) Rsource = 5R⊙. Fig. 9. Illustration of the increasing source radius, compare Figure 6(b): (a) Rsource = 2 R⊙, (b) 3 R⊙, and (c) 5 R⊙, all other parameters as given in Table 1. The triple-lens light curve (solid) is obtained from the magnification pattern in Figure 4, the best- fit binary-lens light curve (dashed) from the corresponding mag- nification pattern without a moon. The difference between the two curves is plotted as a dotted line. 10 C. Liebig & J. Wambsganss: Exomoons as gravitational lenses of Gaudi et al. (2008) or Janczak et al. (2009). The photometric uncertainty directly enters the statistical evaluation of each light curve as described in appendix A. We have drawn results for an (unrealistically) broad σ-range from 0.5 mmag to 500 mmag. An error as small as σ = 0.5 mmag for a V = 12.3 star has recently been reached with high-precision photometry of an ex- oplanetary transit event at the Danish 1.54 m telescope at ESO La Silla (Southworth et al. 2009), which is one of the telescopes presently used for Galactic microlensing monitoring observa- tions, but this low uncertainty cannot be transferred to standard microlensing observations, where crowded star fields make e.g. the use of defocussing techniques impossible. In our discussion, we regard only the more realistic range from σ = 5 to 100 mmag. We adopt an ideal scenario with a fixed σ-value for all sampled points on the triple-lens light curve and set σ = 20 mmag to be our standard for the comparison of different mass and separa- tion scenarios. Summarised: we explore 5 values as photometric uncertainty: σ = 5, 10, 20, 50, 100 mmag. Parameter DS DL MStar µ⊥ * qPS * θPS * qMP * θMP * RSource Standard value 8 kpc 6 kpc 0.3 M⊙ 7 mas/year 10−3 1.3 θE 10−2 1.0 θP E R⊙ ≃ 1 frame 15 minutes σ 20 mmag fsampled * * Comment distance to Galactic bulge ⊥ = 200 km/s at DL = 6 kpc most abundant type of star = v Jupiter/Sun mass ratio wide separation caustic Moon/Earth mass ratio planetary Einstein radius brightness requirements vs. stellar abundance high-cadence observation typical value in past observations Table 1. Parameter values of our standard scenario. Parameters marked with an asterisk (*) are varied in our simulations in order to evaluate their influence on the lunar detection rate and to com- pare different triple-lens scenarios. The fixed parameters lead to values for the Einstein ring radius, θE = 0.32 mas, i.e. 1.9 AU in the lens plane, and the Einstein time, tE ≃ 17 days. The lensed system is a Saturn-mass planet at a projected separation of 2.5 AU from its 0.3 M⊙ M-dwarf host, the Earth-mass satel- lite orbits the planet at 0.06 AU, i.e. 0.01 mas angular separation, cf. Figure 10. 5. Results: detectability of extrasolar moons in microlensing light curves The numerical results of our work are presented and a first inter- pretation is made. We start by considering the result for a single magnification pattern in Section 5.1. We then present results that cover specific physical triple-lens scenarios (Section 5.2). We also evaluate the parameter dependence of the detectability rate, by varying each parameter separately. 5.1. Evaluationofasinglemagnificationpattern As described in Section 3, all magnification maps are evaluated by taking a large sample of unbiased2 light curves and then de- 2 I.e. independent of lunar features, but required to pass close to or through the planetary caustic. termining the fraction of light curves with significant lunar sig- nals. The results for a single magnification map (here the pattern from Figure 5), are displayed in Table 2. The percentages can σ in mmag 5 10 20 50 100 Proportion cantly curves of deviating signifi- light Detectability 190/228 132/228 77/228 39/228 1/228 83.3 % 57.9 % 33.8 % 17.1 % 0.4 % Table 2. Detectability of the third mass in the magnification pat- tern of Figure 5, for different assumed photometric uncertainties. be interpreted as the "detection probability" of the moon in a random observed light curve displaying a signature of the plan- etary caustic – ignoring the possible complication to fully char- acterise the signal. In this map, the detectability of the moon is about one third, provided we have an observational uncertainty of σ = 20 mmag. Obviously, with a larger photometric error, one is less sensitive to small deviations caused by the moon. 5.2. Resultsforselectedscenarios In order to be able to make general statements about a given planetary/lunar system, we average the results for 12 mag- nification patterns representing a full lunar orbit with φ = 0◦, 30◦, . . . , 330◦. 5.2.1. Variation of the photometric uncertainty The scenario described in Table 1 is evaluated for different pho- tometric uncertainties by analysing the 12 magnification maps shown in Figure 2. The results are displayed in Table 3. For σ in mmag Detectability 90.2 % 59.8 % 30.6 % 12.8 % 1.5 % 5 10 20 50 100 Table 3. Detectability of the moon in our standard scenario as a function of assumed photometric uncertainty. an photometric uncertainty of σ = 20mmag, the result is that about 30 % of the light curves show significant deviations from a best-fit binary-lens light curve and display detectable signs of the moon. We recall again our requirements: – The host planet of the satellite has been detected. – A small source size, up to a few solar radii, is required. – The moon is massive compared to satellites in the solar sys- tem. – We have assumed light curves with a constant sampling rate of one frame every 15 minutes for about 50 to 70 hours. M−Dwarf host star C. Liebig & J. Wambsganss: Exomoons as gravitational lenses 11 Inner boundary of planet lensing zone 0.6 θE Stellar Einstein radius θE Planet orbit 1.3 θE Saturn−mass planet Earth−mass moon 3.0 θP E 0.6 θP E Moon orbit θP E Fig. 10. Visualisation of the approximate lunar resonance zone (dark-grey ring), where planetary and lunar caustic overlap and interact. Also depicted is the planet lensing zone (light grey), which is the area of planet positions for which the planetary caustic(s) lie within the stellar Einstein radius. Distances and caustic inset are to scale. The physical scale depends not only on the masses of the lensing bodies, but also on the distances to the lens system and the background star. In our adopted standard scenario, cf. Table 1, the Einstein radius θE corresponds to 1.9 AU in the lens plane. This places the planet at a projected orbit radius of 2.5 AU, whereas the moon's orbit has a radius of 0.06 AU (corresponding to one planetary Einstein radius). The lunar resonance zone covers lunar orbit radii from 0.04 AU to 0.18 AU. As would be expected, the detection rate increases with higher photometric sensitivity and decreases with a larger uncertainty. At a photometric error of 20 mmag or worse, the detection of a moon depends on whether the source trajectory passes through or close to the lunar caustic. The finite magnification resolution of the magnification patterns induces an uncertainty in the de- tectability results that increases with smaller σ-values, but it is negligible for the examined range. 5.2.2. Changing the angular separation of planet and moon Table 4 lines up the changing detectability for different values of the projected planet-moon separation θMP, with all other param- eters as in standard scenario (Table 1). With the chosen range θMP in θP 0.8 1.0 1.2 1.4 E Detectability 29.6 % 30.6 % 30.4 % 28.1 % Table 4. Detectability of the moon as a function of the planet- moon separation for our standard scenario. E to 1.4 θP from θMP = 0.8 θP E, we are within the regime of caustic interactions of planetary and lunar caustic. We find that the per- centage of detectability is almost constant at a 30 % level. Han (2008) defines regions of possible satellite detections as the re- gion of angular moon-planet separations between the lower limit of one planetary Einstein radius and an upper limit at the pro- jected Hill-radius average. From our results, we conclude that detections with θMP < 1.0 θP E are well possible. We have not fully probed the lower limit region, but for θMP < 0.5 θP E the size of the lunar caustic perturbation decreases substantially. 5.2.3. Variation of the moon mass Table 5 shows the changing detectability for a varying mass ratio qMP of lunar and planetary mass, with all other parameters as in our standard scenario (Table 1). Over the range of the moon- qMP 10−3 10−2 10−1 Detectability 2.4 % 30.6 % 85.6 % Table 5. Detectability of the moon depending on moon-planet mass ratio. planet mass ratio, qMP = 10−3, 10−2, and 10−1 the detection rate increases from 2.4 % to 85.6 %, compare also Figure 7. Not surprisingly: the more massive the moon, the easier it is to detect. 5.2.4. Changing the planetary mass ratio Table 6 lists the changing detectability for a varying mass ratio between planet and star qPS . We adjusted the parameters of the qPS 3 × 10−4 1 × 10−3 3 × 10−3 Detectability 19.6 % 30.6 % 42.7 % Table 6. Dependence of lunar detectability as a function of the planet-star mass ratio, all other parameters as in our standard scenario. grid of source trajectories, so that the grid spacing scales with the total caustic size. For a decreasing mass of the planet, the ap- parent source size increases relative to the caustic size, so finer features will be blurred out more in the case of a smaller planet. Also the absolute change in magnification scales with the plan- etary mass. As expected, the detection efficiency is roughly pro- portional to the planetary caustic size. 12 C. Liebig & J. Wambsganss: Exomoons as gravitational lenses 5.2.5. Different sampling rates Table 7 presents the changing lunar detectability for different sampling rates. A lower sampling frequency lowers the detection Sampling step size in θE 6 × 10−4 9 × 10−4 12 × 10−4 fsampled minutes 1/15 1/22 1/30 per Detectability 30.6 % 26.8 % 24.9 % Table 7. Detectability of the moon depending on the sampling rate of the simulated light curves. probability, but the effect is less pronounced than expected. This means it may be favourable to monitor a larger number of plan- etary microlensing events with high sampling frequency, rather than a very small number of them with ultra-high sampling. Our assumed sampling rates can easily be met by follow-up obser- vations, as they are presently performed for anomalous Galactic microlensing events. 5.2.6. Source size variations We analysed simulated light curves of the standard scenario with five different source star radii. They cover the range between a star the size of our Sun and a small giant star with RSource = 10 R⊙ at DS = 6 kpc. RSource in θE 1.8 × 10−3 3.6 × 10−3 5.4 × 10−3 9.0 × 10−3 18.0 × 10−3 RSource in R⊙ 1.0 2.0 3.0 5.0 10.0 Detectability 30.6 % 20.9 % 18.1 % 4.1 % 0.0 % Table 8. Detectability of the moon depending on the size of the source star, all other parameters as in our standard scenario. A larger source blurs out all sharp caustic-crossing features of a light curve, compare Figure 9. We see from our results in Table 8 that the discovery of a moderately massive moon (see standard scenario assumptions in Table 1) is impossible for a source star size of RSource = 10 R⊙ or larger within our assump- tions. 5.2.7. Increased separation between star and planet For reasons given above in Section 4.3, we did not vary the sep- aration between star and planet significantly, but stayed in the outer ranges of the planet lensing zone, corresponding to a well- detectable, wide-separation planetary caustic. For θPS = 1.4 θE, we get a detectability of 33.3 %, which is similar to our standard case (θPS = 1.3 θE) with 30.6 %. 6. Conclusion and outlook that massive extrasolar moons can principally be detected and identified via the technique of Galactic microlensing. From our results it can be concluded that the detection of an extrasolar moon – under favourable conditions – is within close reach of available observing technology. The unambiguous characteri- sation of observed lunar features however will be challenging. The examined lens scenarios model realistic triple-lens config- urations. An observing rate of about one frame per 15 minutes is desirable, which is high, but well within the range of what is now regularly performed in microlensing follow-up observations of anomalous events. Similarly, an observational uncertainty of about 20 mmag can be met with today's ground-based telescopes and photometric reduction techniques for sufficiently bright tar- gets. Bright microlensing targets are mostly giant stars, which is an impediment to the detection of moons: Bulge giants with radii of order 10 R or larger smooth out any lunar caustic fea- ture. Therefore, in order to find moons, it is crucial to be able to perform precise photometry on small sources with angular sizes of the order of 10−3 Einstein radii, corresponding to a few solar radii or smaller at a distance of 8 kpc. This means dwarf stars rather than giants need to be monitored in order to identify moons in the intervening planetary systems. Some improvement in resolution can be gained with the lucky-imaging technique on medium-sized ground telescopes (Grundahl et al. 2009). Under very favourable circumstances, exomoons might already be de- tectable from ground. However, future space-based telescopes, such as ESA's proposed Euclid3 mission (cf. R´efr´egier et al. (2010), Chapter 17) or the dedicated Microlensing Planet Finder (MPF) mission proposed to NASA (cf. Bennett et al. (2009)), will surely have the potential to find extrasolar moons through gravitational lensing. ⊙ Acknowledgements. We thank our referee, Jean-Philippe Beaulieu, for his con- structive comments, which helped to improve the manuscript. CL would like to thank Sven Marnach for his helpful input on the significance test. This research has made use of NASA's Astrophysics Data System and the arXiv e-print service operated by Cornell University. References Beaulieu, J.-P., Albrow, M., Bennett, D. P., et al. 2006, Nature, 439, 437 Beaulieu, J.-P., Kerins, E., Mao, S., et al. 2008, ESA white paper, arXiv:0808.0005v1 Benn, C. R. 2001, Earth, Moon and Planets, 85, 61 Bennett, D. P., Anderson, J., Beaulieu, J.-P., et al. 2009, white paper, arXiv:0902.3000v1 Bennett, D. P. & Rhie, S. H. 2002, ApJ, 574, 985 Bond, I. A., Udalski, A., Jaroszy´nski, M., et al. 2004, The Astrophysical Journal, 606, L155 Bozza, V. 1999, Astronomy and Astrophysics, 348, 311 Bozza, V. 2000, A&A, 355, 423 Cabrera, J. & Schneider, J. 2007, A&A, 464, 1133 Cassan, A. 2008, A&A, 491, 587 Domingos, R. C., Winter, O. C., & Yokoyama, T. 2006, MNRAS, 373, 1227 Dominik, M., Jørgensen, U. G., Horne, K., et al. 2008, ESA white paper, arXiv:0808.0004v1 Dong, S., Bond, I. A., Gould, A. P., et al. 2009, ApJ, 698, 1826 Dyson, F. W., Eddington, A. S., & Davidson, C. 1920, Royal Society of London Philosophical Transactions Series A, 220, 291 Einstein, A. 1916, Annalen der Physik, 354, 769 Gaudi, B. S., Beaulieu, J.-P., Bennett, D. P., et al. 2009, white paper, arXiv:0903.0880v1 Gaudi, B. S., Bennett, D. P., Udalski, A., Gould, A., et al. 2008, Science, 319, 927 Gaudi, B. S., Naber, R. M., & Sackett, P. D. 1998, ApJ, 502, L33 Gaudi, B. S. & Sackett, P. D. 2000, ApJ, 528, 56 Gould, A. P. 2009, in Astronomical Society of the Pacific Conference Series, ed. In this work we provide probabilities for the detection of moons around extrasolar planets with gravitational lensing. We showed K. Z. Stanek, Vol. 403, 86 3 sci.esa.int/euclid C. Liebig & J. Wambsganss: Exomoons as gravitational lenses 13 Grundahl, F., Christensen-Dalsgaard, J., Kjeldsen, H., et al. 2009, The Stellar Observations Network Group - the Prototype, arXiv:0908.0436v1 Han, C. 2008, ApJ, 684, 884 Han, C. & Han, W. 2002, ApJ, 580, 490 Holman, M. J. & Murray, N. W. 2005, Science, 307, 1288 Hunter, R. B. 1967, MNRAS, 136, 245 Janczak, arXiv:0908.0529v1 J., Fukui, A., Dong, S., et al. 2009, Submitted to ApJ, Kayser, R., Refsdal, S., & Stabell, R. 1986, A&A, 166, 36 Kipping, D. M. 2009a, MNRAS, 392, 181 Kipping, D. M. 2009b, MNRAS, 396, 1797 Kipping, D. M., Fossey, S. J., & Campanella, G. 2009, MNRAS, 400, 398 Lewis, K. M., Sackett, P. D., & Mardling, R. A. 2008, ApJ, 685, L153 Liebig, C. 2009, Diploma thesis, Universitat Heidelberg Mayor, M. & Queloz, D. 1995, Nature, 378, 355 Muraki, Y., Sumi, T., Abe, F., et al. 1999, Progress of Theoretical Physics Supplement, 133, 233 Paczy´nski, B. 1996, Annual Review of Astronomy and Astrophysics, 34, 419 R´efr´egier, A., Amara, A., Kitching, T. D., et al. 2010, arXiv:1001.0061v1 Rhie, S. H. 1997, The Astrophysical Journal, 484, 63 Rhie, S. H. 2002, How Cumbersome is a Tenth Order Polynomial?: The Case of Gravitational Triple Lens Equation, arXiv:astro-ph/0202294v1 Sartoretti, P. & Schneider, J. 1999, A&AS, 134, 553 Scharf, C. A. 2006, ApJ, 648, 1196 Schneider, P., Kochanek, C. S., & Wambsganss, J. 2006, Gravitational Lensing: Strong, Weak and Micro, Saas-Fee Advanced Courses 33 (Springer-Verlag Berlin Heidelberg) Schneider, P. & Weiss, A. 1986, A&A, 164, 237 Southworth, J., Hinse, T. C., Jørgensen, U. G., et al. 2009, MNRAS, 396, 1023 Udalski, A., Szyma´nski, M., Kałuzny, J., Kubiak, M., & Mateo, M. 1992, Acta Astronomica, 42, 253 Wambsganss, J. 1990, PhD thesis, Ludwig-Maximilians-Universitat Munchen Wambsganss, J. 1997, MNRAS, 284, 172 Wambsganss, J. 1999, Journal of Computational and Applied Mathematics, 109, 353 Williams, D. M. & Knacke, R. F. 2004, Astrobiology, 4, 400 Witt, H. J. 1990, A&A, 236, 311 Appendix A: Significance of deviation This section explains the method that we use to determine the significance of deviation between the two simulated light curves of the triple lens and the best-fit binary lens model. When we want to decide whether real observational data is better described with a triple-lens or with a binary-lens model, we can fit both models to the data and use the χ2-test to assess the significance of the deviation between the data and the models and pick the model with the better fit. We do not have observational data, but numerically com- puted light curves. If we observed one of the simulated triple- lens systems, we would expect the data points to have a Gaussian distribution around the theoretical values. We could produce arti- ficial data by randomly scattering the simulated triple-lens light curve data points around their theoretical values and then fit a binary-lens model and the triple-lens model to this artificial data set, as described in Section 3.4. This approach allows us to use the usual χ2-test, but is computationally intensive, because the models have to be fitted numerous times to get a statistically sound sample of χ2-values. In our approach, we directly calcu- late the χ2-value which is the expectation value of the above method. We hypothesise, every simulated point on our triple-lens light curve is in fact the mean µt i of the distribution of a random variable Xi that is distributed according to the Gaussian probabil- ity density function fi(xi) with a standard deviation of σt is the variance of the distribution.4 i. (σt i)2 We now want to examine whether the Xi could be described i . We introduce the new equally well with a binary-lens model µb χ2-distributed5 random variable Q2 = n Xi=1 i Xi − µb σb i  2 .  At this step, we could simulate data Xi in order to find a some- what representative, randomly drawn value of Q2, but instead we simply calculate what the mean value of all possible Q2 would be. We use the definitions, to find hQ2i = * n Xi=1 i Xi − µb σb i  2  + = n Xi=1 1 (σb i )2E i )2 D(Xi − µb Here, we keep in mind that, in general, hx2i , hxi2. Using the parameters µt i and σt i of the distribution fi(xi), we reduce the equation by calculating (xi − µb i )2 fi(xi)dxi. hQ2i = = = n n Xi=1 Xi=1 Xi=1 n 1 (σb 1 (σb 1 (σb −∞ i )2 Z ∞ i )2 Z ∞ i )2 (cid:16)(σt (xi − µt −∞ i)2 + (µt i + µt i − µb i )2 fi(xi)dxi i − µb i )2(cid:17) . i and (σt We used the definitions of µt probability density functionR fi(xi)dxi = 1. In our case σt i)2 and the property of the i = σb i holds true without loss of accuracy, and indeed, we simplify σi = σ for all i, that is we assume that the photometric uncer- tainty is the same for all data points, and argue that this does not pose a problem as long as σ is chosen to match the maximum photometric uncertainty. So we can further reduce to i − µb µt (1 + i )2) = n + 2 n n . 1 σ2 (µt i − µb hQ2i = Xi=1 Xi=1 This is the mean of all Q2 possibly resulting, when compar- ing the simulated binary-lens light curve to randomly scattered triple-lens light curve points. It shall be our measure of devia- tion. Fortunately, we can provide all remaining parameters:  i σ  4 We recall the definitions of mean and variance. For the random vari- able X with probability density function f (x), they are in general µ = Z ∞ σ2 = Z ∞ −∞ −∞ and x f (x) dx = hXi (x − µ)2 f (x) dx = h(X − µ)2i, where we also introduce the notation hXi for the expectation value of X. 5 The χ2-distribution can be applied to the sum of n independent, normally distributed random variables Zi with the mean µi = 0 and the standard deviation σi = 1 for all i, χ2 ∼ n Xi=1 Z2 i . See any standard book with an introduction to calculus of probability. 14 n is the number of degrees of freedom, equal to the number of compared data points. C. Liebig & J. Wambsganss: Exomoons as gravitational lenses avoid refitting the binary-lens model light curve and always com- pare to the one we got as best-fit to the exact (simulated) triple- lens model light curve. This overestimation effect vanishes for n → ∞ and is negligible for a sufficiently large n. The high sam- pling frequency we use, ensures that the number of data points is always large enough (n > 250 in all cases). While we cannot exactly quantify the uncertainty of the re- sults, we know that they pose strict upper limits for the lunar detectability in the various scenarios. This knowledge would en- able us to infer a tentative census of extrasolar moons, once the first successful microlensing detections have been achieved. σ is the assumed standard deviation or photometric uncertainty of observations. i )2 equals the difference between the two compared light (µt i − µb curves squared, which we already use for our least square fit. → y t i s n e d y t i l i b a b o r p χ2 probability density function hχ2i = degrees of freedom hQ2i P ≤ 0.01 X → Fig. A.1. The χ2 probability density function plotted for illus- tration only. For increasing degrees of freedom the central limit theorem takes effect and the curve will very closely resemble a Gaussian distribution. If P(X ≥ hQ2i) ≤ 1%, i.e. if the proba- bility P for a χ2-distributed random variable X to be larger or as large as the mean of Q2 is less than 0.01, we consider the deviation between the two compared curves to be significant. Now we have to decide whether the deviation between our two examined light curves is significant. By taking hQ2i as χ2, one can evaluate the cumulative distribution function Fn(z) = R z f (x)dx of the χ2-distribution - we relied on the GNU −∞ Scientific Library6 - to find the corresponding probability for any given χ2-distributed random variable X to be as large as or larger than hQ2i, P(X ≥ hQ2i) = 1 − Fn(hQ2i). This probability is just the integral of the χ2 probability density function to the right of hQ2i as illustrated in Figure A.1. If this probability is very small, hQ2i is outside the expected range for a χ2-distributed random variable Q2. In that case, Q2 is obviously not χ2-distributed with n degrees of freedom, so we must con- clude that there is a significant deviation between the triple lens model light curve and the binary-lens model light curve. We say we have a significant deviation between our two curves, if the mean value hQ2i is so high that the probability P for any random variable X being larger or equally large is less than 1%. We interpret this to say, only if the probability for a given triple lens light curve with independent, normally distributed data points to be random fluctuation of the compared binary-lens light curve is less than 1%, we consider it to be prin- cipally detectable. There are two known sources for overestimating the detec- tion rate coming with our method. First, systematic errors are not accounted for. A possible solution could be to add a further term to hQ2i as in hQ2 , where σsys can be as- sumed to lie in the few percent region for real data. Secondly, we σ (cid:17)2 newi = hQ2i−n(cid:16) σsys 6 http://www.gnu.org/software/gsl/
1202.5412
1
1202
2012-02-24T10:28:16
Minimum Dust Abundances for Planetesimal Formation via Secular Gravitational Instabilities
[ "astro-ph.EP" ]
We estimate minimum dust abundances required for secular gravitational instability (SGI) to operate at the midplane dust layer of protoplanetary disks. For SGI to be a viable process, the growth time of the instability T_grow must be shorter than the radial drift time of the dust T_drift. The growth time depends on the turbulent diffusion parameter alpha, because the modes with short wavelengths are stabilized by turbulent diffusion. Assuming that turbulence is excited via the Kelvin-Helmholtz or streaming instabilities in the dust layer, and that its strength is controlled by the energy supply rate from dust accretion, we estimate the diffusion parameter and the growth time of the instability. The condition T_grow < T_drift requires that the dust abundance must be greater than a critical abundance Z_min, which is a function of the Toomre parameter Q_g and aspect ratio h_g / r of the gas disk. For a wide range of parameter space, the required dust abundance is less than 0.1. A slight increase in dust abundance opens a possible route for the dust to directly collapse to planetesimals.
astro-ph.EP
astro-ph
Accepted by the Astrophysical Journal Preprint typeset using LATEX style emulateapj v. 5/2/11 2 1 0 2 b e F 4 2 . ] P E h p - o r t s a [ 1 v 2 1 4 5 . 2 0 2 1 : v i X r a MINIMUM DUST ABUNDANCES FOR PLANETESIMAL FORMATION VIA SECULAR GRAVITATIONAL Department of Earth and Planetary Sciences, Tokyo Institute of Technology, Meguro-ku, Tokyo, 152-8551, Japan INSTABILITIES Taku Takeuchi1 and Shigeru Ida Accepted by the Astrophysical Journal ABSTRACT We estimate minimum dust abundances required for secular gravitational instability (SGI) to op- erate at the midplane dust layer of protoplanetary disks. For SGI to be a viable process, the growth time of the instability Tgrow must be shorter than the radial drift time of the dust Tdrift. The growth time depends on the turbulent diffusion parameter α, because the modes with short wavelengths are stabilized by turbulent diffusion. Assuming that turbulence is excited via the Kelvin-Helmholtz or streaming instabilities in the dust layer, and that its strength is controlled by the energy supply rate from dust accretion, we estimate the diffusion parameter and the growth time of the instability. The condition Tgrow < Tdrift requires that the dust abundance must be greater than a critical abundance Zmin, which is a function of the Toomre parameter Qg and aspect ratio hg/r of the gas disk. For a wide range of parameter space, the required dust abundance is less than 0.1. A slight increase in dust abundance opens a possible route for the dust to directly collapse to planetesimals. Subject headings: planets and satellites: formation -- protoplanetary disks 1. INTRODUCTION Formation processes of planetesimals have not been well understood in planet formation theory. The grav- itational instability of the dust layer at the midplane of a protoplanetary disk has been proposed as a pos- sible route to planetesimal formation. The classical sce- nario of gravitational instability requires high densities of dust layers to surpass the Roche limit (Goldreich & Ward 1973; Sekiya 1983). Dust sedimentation only in the vertical direction hardly achieves such a high density (Sekiya 1998). The dust layer becomes turbulent via Kelvin-Helmholtz (KH) or streaming instabilities when its dust density exceeds the gas density, which is much less than the Roche density, and further dust settling is significantly suppressed (Chiang & Youdin 2010 for re- view). Radial drift of the dust caused by gas drag may provide a possible route for further accumulation of the dust. When the dissipative effects of gas drag are in- cluded, the dust layer is secularly gravitationally unsta- ble to the modes of dust accumulation in the radial di- rection (Ward 2000; Youdin 2005a, 2005b). This secular gravitational instability (SGI) occurs in any dust layers even if their densities are small (i.e., there is no crite- rion on their Toomre's Q values for SGI; Youdin 2011; Shariff & Cuzzi 2011; Michikoshi et al. 2012; these pa- pers are referred hereafter as Y11, SC11, and MKI12, respectively). However, the growth rate of SGI is greatly suppressed by gas drag, particularly for small particles, and could be too slow to have any effect on planetesimal formation. Y11, SC11, and MKI12 argue that the growth timescale must be shorter than the lifetime of the dust drifting toward the star owing to gas drag. The wavelength of the most unstable mode is determined by the balance between the radial accumulation of dust particles due to self-gravity and their diffusion due to gas turbulence. 1 [email protected] Only perturbations with wavelengths sufficiently long are unstable. For stronger turbulence, wavelengths of the unstable modes, as well as growth timescales, are longer. Y11 and SC11 have calculated the growth timescales for various values of the turbulent diffusion parameter α, and have derived an upper limit on α required for the growth timescale to be shorter than the drift timescale. Their approach is general and can be applied for any disk turbulence, provided that the value of α is known. In this paper, we focus on disk turbulence induced in the dust layer, assuming that the gas disk itself is a glob- ally laminar flow. In such a disk, the concentration of dust particles at its midplane triggers turbulence. The velocity difference between the dust and gas induces the KH or streaming instabilities and causes turbulence in the dust layer (Chiang & Youdin 2010). The ultimate energy source for turbulence is the accretion energy of the dust drifting toward the star, regardless of the kind of instability that occurs in the dust layer. Thus, a simple energetics can be applied to estimate the turbu- lence strength, provided that the dust accretion velocity is known. Takeuchi et al. (2012, hereafter T12) calculate the energy supply rate to turbulence, using the classical formulae on the drift velocity of the dust (Nakagawa et al. 1986; Weidenschilling 2003; Youdin & Chiang 2004), and then estimate the turbulence strength or the value of α. T12 has shown that, from the comparison between the estimated value of α and the recent results of numer- ical simulations of turbulence by Johansen et al. (2006) and Bai & Stone (2010), the turbulence strength excited via KH and/or streaming instabilities can be estimated from the dust accretion rate. While Y11, SC11, and MKI12 have derived the growth time of SGI as a function of α and other disk parameters (Equation (51) of Y11), T12 has obtained α values of tur- bulence in the dust layer (Equation (5) below). Combin- ing these results gives an expression of the growth time without the unknown parameter α, as shown in Section 2.1. Consequently, in Section 2.3, the criterion for SGI 2 Takeuchi & Ida to operate is obtained as a condition on the dust abun- dance Z in the disk. For SGI to operate faster than the dust drift timescale, the dust abundance must be greater than a specific critical value Zmin, which is a function of the Toomre Q value of the gas disk and the deviation fraction η of the gas velocity from the Keplerian velocity. The Zmin value derived in this paper gives the minimum dust abundance required for SGI, because we consider only turbulence induced in the dust layer. If turbulence were caused by another mechanism such as magneto- rotational instability (MRI) of the gas disk, α would be larger than the value we adopted, and consequently the required value of Zmin would increase. Hence, Zmin de- rived in this paper is considered as the minimum value required for SGI in realistic disks. 2. CONDITION FOR SECULAR GRAVITATIONAL INSTABILITY We consider a protoplanetary disk initially in a laminar flow state. Sedimentation of dust particles to the mid- plane of the disk induces turbulence via KH or streaming instabilities caused by velocity differences between the dust and gas. A (quasi-)steady dust layer forms when turbulent diffusion matches dust settling, in which steady state turbulence is maintained. We neglect intermittency of turbulence. If an extra source of turbulence is present such as global turbulence of the gas disk via MRI, the turbulence would be stronger than that excited only by the dust-gas velocity difference. Thus, in this paper, we consider the minimum strength of turbulence. The dust particles are characterized by their stopping time ts, which is the timescale of damping the velocity difference from the gas. We define the non-dimensional stopping time, Ts = tsΩK, normalized by the Keplerian frequency ΩK. In this study, we focus on the dynamics of relatively small particles such that Ts . 1. Particles with Ts ∼ 1 experience high speed collisions and the fastest radial drift, which likely hinder particle growth. SGI of the dust layer is a possible route for directly forming large bodies (Ts ≫ 1) from small particles (Ts ≪ 1). In the following subsections, we compare the growth time of SGI with the orbital drift time of the dust parti- cles due to gas drag in order to determine the condition for SGI to be a relevant process for planetesimal forma- tion. 2.1. Timescale of Secular Gravitational Instability In his Equation (51), Y11 shows that the growth time of SGI for small particles with Ts . 1, normalized by the Keplerian time Ω−1 K , is Tgrow ≈ αQ2 g Z 2T 2 s . (1) Here α is the turbulent diffusion parameter (see below), Qg is the Toomre stability parameter of the gas disk, Qg = cgΩK πGΣg , (2) growth time depends on the turbulent diffusion parame- ter α through its stabilizing effect for perturbations with short wavelengths. Turbulence in the dust layer is induced by velocity dif- ferences between the dust and gas. If the drag force be- tween the gas and dust were not effective, the dust would orbit with the Keplerian velocity vK, while the gas would orbit with a sub-Keplerian velocity vg = (1 − η)vK , (3) K)−1∂P/∂r ∼ (cg/vK)2 ∼ 10−3 − where η = −(2ρgrΩ2 10−2, ρg is the gas density, and P is the gas pressure. In the calculation of α, a slightly different definition of η is useful (see T12 or Appendix A in this paper). We introduce η defined by cg (cid:19)2 η =(cid:18) vK η2 = Cηη , (4) where Cη is a factor of the order of unity, which depends on the radial profile of the gas disk. In this paper, we adopt Cη = 1 (i.e., η = η) for simplicity, except in Section 3 where specific disk models are considered. T12 shows that the dust particles accrete toward the star due to gas drag and supply energy to turbulence via KH and/or streaming instabilities. Dust accretion is either caused by the gas drag acting on individual par- ticles or by turbulent drag acting on the surface of the dust layer. The former (individual drag) is effective if the dust-to-gas ratio at the disk midplane, fmid = ρd/ρgz=0, is less than unity, and the latter (collective drag) is ef- fective if fmid & 1. Appendix A briefly summarizes dust accretion due to both drag types (see T12 for detailed discussions). The turbulence strength or the parameter α is determined by the energy supply rate from dust ac- cretion toward the star. The approximate expression of α for Ts . 1 particles is given by Ts , (5) α ≈h(C1Ceff ηZ)− 2 3 +(cid:0)C2Ceff ηZ −1(cid:1)−2i−1 where Ceff = 0.19 is the energy supply efficiency (see T12), and C1 = 1.0 and C2 = 1.6 are the numerical factors. This expression connects the approximate for- mulae of α for the two limiting regimes of fmid ≪ 1 and fmid ≫ 1 (Equations (A2) and (A3)). Note that the condition fmid ≪ 1 (or fmid ≫ 1) corresponds to the condition Z ≪ (Ceff η)1/2 (or Z ≫ (Ceff η)1/2) (see Equa- tion (A5)). The numerical factors C1 and C2 are ad- justed to make an appropriate fit to the overall behavior of the numerical result. A comparison of this approxi- mate expression to the numerically calculated α is shown in Appendix A. Substituting Equation (5) in Equation (1) gives Tgrow ≈ Q2 g Z 2Ts h(C1Ceff ηZ)− 2 3 +(cid:0)C2Ceff ηZ −1(cid:1)−2i−1 , (6) showing that Tgrow is a function of the disk parameters (η, Z, Qg) and is inversely proportional to Ts. where cg is the sound speed of the gas, G is the gravita- tional constant, and Σg is the surface density of the gas disk. The disk "metallicity" Z is the ratio of the surface densities between the dust and gas, Z = Σd/Σg. The 2.2. Timescale of Radial Drift According to T12, the non-dimensional timescale of the radial drift of dust particles due to gas drag is estimated t f i r d T / w o r g T t f i r d T / w o r g T 104 103 102 101 100 10-1 10-2 10-3 10-4 104 103 102 101 100 10-1 10-2 10-3 10-4 10-4 10-4 Secular Gravitational Instability of the Dust Layer 3 h =3x10-3 (a) Qg = 30 h =10-2 h =10-3 103 102 101 g Q 10-3 10-2 Z 10-1 100 100 -5 10-4 Qg = 30 (b) η= 3 x 10-3 Qg = 100 Qg = 10 103 102 101 g Q 10-3 10-2 Z 10-1 100 -4 100 10-4 (a) log(Zmin) for Tgrow < Tdrift 0 -0.5 -1 -1.5 -2 -2.5 -3 -3.5 -4 -4.5 10-3 h~ 10-2 10-1 (b) log(Zmin) for 10Tgrow < Tdrift 0 -0.5 -1 -1.5 -2.5 -2 -3 -3.5 10-3 h~ 10-2 10-1 Fig. 1. -- Ratio of the growth time to the radial drift time, Tgrow/Tdrift, against dust abundance Z. (a) For models with a fixed Toomre parameter of the gas disk, Qg = 30. (b) For models with a fixed deviation fraction of the gas velocity from the Keple- rian velocity, η = 3 × 10−3. The horizontal dotted line indicates Tgrow = Tdrift. as Tdrift ≈ fmid + 1 2ηTs , (7) where the midplane dust-to-gas ratio fmid is approxi- mately given by 1 3 fmid ≈(cid:18) Z 2 C1Ceff η(cid:19) +(cid:18) Z 2 C2Ceff η(cid:19) . (8) In the above estimate, both individual and collective drags are considered. Derivation of the above expres- sions are described in Appendix A. The timescale of the radial drift is a function of (η, Z) and is inversely pro- portional to Ts. 2.3. Comparison of Timescales 2.3.1. Minimum Dust Abundances Y11, SC11, and MKI12 show that dust layers are al- ways unstable and subject to SGI, but the growth rate for particles tightly coupled to the gas is strongly suppressed by gas drag. The growth timescale can be larger than the other timescales, in which the dust layer evolves signif- icantly by other processes such as the dispersal of the gas disk and the radial drift of the dust to the star. The condition for SGI to be a relevant process for planetesi- mal formation dictates that its growth timescale must be Fig. 2. -- Contour map of the minimum dust abundance Zmin on the η-Qg plane. (a) Zmin required for the condition Tgrow < Tdrift. The contours are labeled by log(Zmin). The dashed line shows the locus of Zmin = (Ceff η)1/2 (i.e., fmid ≈ 1 when Z = Zmin). (b) Zmin required for the condition 10Tgrow < Tdrift. shorter than the other evolution timescales. Y11 shows that the radial drift imposes the most stringent condition for a wide range of disk parameters. We discuss the con- dition for Tgrow < Tdrift. Because both Tgrow and Tdrift are inversely proportional to Ts, the ratio Tgrow/Tdrift is independent of Ts; that is, the condition Tgrow < Tdrift is not affected by particle size. We consider dependence of Tgrow/Tdrift on the parameters (η, Z, Qg). First, we examine the limiting cases of Z ≪ (Ceff η)1/2 and Z ≫ (Ceff η)1/2 (fmid ≪ 1 and fmid ≫ 1). In both limits, Tgrow Tdrift ≈(2C −1 (C1Ceff )2/3 η5/3Q2 η (C2Ceff )3 η4Q2 η 2C −1 gZ −4/3 for Z ≪ (Ceff η)1/2 for Z ≫ (Ceff η)1/2 . gZ −6 (9) For Z ≪ (Ceff η)1/2, Tgrow/Tdrift is proportional to Z −4/3, while for Z ≫ (Ceff η)1/2 it rapidly decreases as Tgrow/Tdrift ∝ Z −6. In Figure 1, Tgrow/Tdrift is plot- ted against Z for various values of η and Qg. For Z ≫ (Ceff η)1/2 ∼ 10−2, it is evident that Tgrow/Tdrift rapidly decreases. For sufficiently large values of Z, Tgrow/Tdrift is less than unity, indicating that SGI oper- 4 103 102 101 g Q log(TgrowTs) 4 3.5 3 2.5 2 Takeuchi & Ida 103 102 101 g Q -4.5 -5 -4 log(a / Ts) -3.5 -3 -2.5 -2 100 10-4 10-3 h~ 10-2 1.5 1 10-1 -5.5 100 -6 10-4 10-3 h~ 10-2 10-1 Fig. 3. -- Contour map of the growth time multiplied by the stop- ping time, TgrowTs, on the η-Qg plane. The contours are labeled by log(Tgrow Ts). In calculating Tgrow(η, Qg), the minimum dust abundances Zmin(η, Qg) are used. The dashed line shows the locus of Zmin = (Ceff η)1/2. ates faster than the radial drift of the dust. From Figure 1, the minimum dust abundance required for SGI, Zmin, is determined as Z satisfying Tgrow/Tdrift = 1. The min- imum dust abundance Zmin ranges between 10−2 and 10−1 for parameters (η, Qg) adopted in Figure 1. In Figure 2(a), the minimum dust abundance Zmin re- quired for Tgrow < Tdrift is plotted as contours on the η- Qg plane. The locus of Zmin = (Ceff η)1/2 is represented as a dashed line on the η-Qg plane. On this line, the midplane dust-to-gas ratio fmid becomes close to unity when Z = Zmin. Substitution of Z = (Ceff η)1/2 into Equation (9) shows that the condition Tgrow/Tdrift = 1 (i.e., Z = Zmin) becomes Qg ≈ η−1/2. Thus, the dashed line in Figure 2(a) is linear with a slope of −1/2. Above this line, Zmin > (Ceff η)1/2 (and fmid > 1 for Z = Zmin); below it, Zmin < (Ceff η)1/2 (and fmid < 1). The depen- dence of Zmin on (η, Qg) is expressed as Zmin ≈(cid:26) (8C −3 (2C −1 η C2 η C3 1 C2 2 C3 eff η5Q6 eff η4Q2 g)1/4 for Qg ≪ η−1/2 g)1/6 for Qg ≫ η−1/2 . (10) Larger dust abundances are required for larger η (i.e., hotter gas disks) and larger Qg (i.e., less massive gas disks). In standard models for protoplanetary disks, η = 10−3 − 10−2, and Qg = 10 − 1000 (see Figure 6 below). Thus, if the dust abundance Z is larger than 0.1, such dusty disks operate SGI. If the gas disk is so cold that η is as small as 10−3 and if it is so massive that Qg is as small as 10, then even a standard value Z = 0.01 is sufficiently large to operate SGI. 2.3.2. Dependence of Zmin on Some Model Parameters The growth time Tgrow is just a time for density per- turbations to increase e-fold. Because many e-folds are needed for a significant density increase, actual condi- tion for SGI would be RTgrow < Tdrift, where R > 1 is the required number of e-folding. From Equa- tion (9) the minimum dust abundance Zmin scales as Zmin ∝ R3/4 for Qg ≪ (Rη)−1/2 and Zmin ∝ R1/6 for Qg ≫ (Rη)−1/2. Figure 2(b) shows Zmin derived for the condition 10Tgrow < Tdrift. For less massive gas Fig. 4. -- Contour map of the diffusion parameter divided by the stopping time, α/Ts, on the η-Qg plane for the dust layer with Zmin. The contours are labeled by log(α/Ts). The dashed line shows the locus of Zmin = (Ceff η)1/2. disks (Qg ≫ (Rη)−1/2; above the dashed line in Fig- ure 2), Zmin depends only weakly on R. For massive gas disks (Qg ≪ (Rη)−1/2), required Zmin is higher for larger R (Zmin ∝ R3/4), but would not be too high (i.e., Zmin . 0.1 even for R = 10; see the region below the dashed line in Figure 2(b)). In the following discussions, we use the condition Tgrow < Tdrift. In this paper, it is assumed that about 20% of the dust accretion energy is used for turbulence excitation (Ceff = 0.19). This value is determined by compari- son with the numerical simulations of turbulence excited via KH instability by Johansen et al. (2006). Because their simulation was two-dimensional, the realistic value of Ceff could be different. For example, three-dimensional simulations by Lee et al. (2010) show that the critical Richardson number for KH instability increases with Z, implying that Ceff also increases with Z (see discussion in Section 5.2.1 of T12). As shown in Equation (10), Zmin ∝ C1/2 eff . For the maximum efficiency (Ceff = 1), Zmin would be about 2 times larger than our estimate. 2.4. Conditions on Tgrow, α, and λ In the previous subsection, the condition for Tgrow < Tdrift was discussed. Next, we discuss other constraints required for SGI operation. First, we consider the con- dition for the growth time to be less than the disk life time. The growth time multiplied by the stopping time, TgrowTs, is plotted on the η-Qg plane in Figure 3. The values of Tgrow(η, Qg) are calculated using the minimum dust abundance Z = Zmin(η, Qg), the values of which differ according to (η, Qg) as shown in Figure 2. For larger values of Z, Tgrow is shorter (Tgrow ∝ Z −4/3 for Z < (Ceff η)1/2 and Tgrow ∝ Z −4 for Z > (Ceff η)1/2). Be- cause Tgrow is inversely proportional to Ts, the contours are labeled by log(TgrowTs). Thus, the growth timescale is obtained by dividing the value in Figure 3 by Ts. For s Ω−1 η > 10−3 and Qg < 102, Tgrow is less than 103T −1 K . Thus at 1AU, the dust layer composed of Ts & 10−3 particles (size a & 1mm) with the dust abundance Zmin operates SGI within a disk life time of ∼ 1Myr. Figure 4 shows the diffusion parameter α due to tur- bulence induced in the dust layer with the abundance Secular Gravitational Instability of the Dust Layer 5 2λ r ≈ 4παQg η1/2 CηZTs , (11) g Q Zmin. Because α is proportional to Ts, the contours are labeled by log(α/Ts). Thus, the α value is obtained by multiplying the value in Figure 4 by Ts. For η > 10−3 and Qg < 102, α is larger than 10−4Ts, except for very small Qg . 10 and η . 3 × 10−3. If the disk is turbu- lent due to other mechanisms than that originating from the dust layer and if turbulent diffusion is stronger than that shown in Figure 4, then the required value for Z would be higher. Even in the dead zone where MRI is not active, the gas could have turbulent motion, causing diffusion of the dust (Fleming & Stone 2003). Okuzumi & Hirose (2011) show that lower values of the diffusion coefficient in the dead zone are realized for a wider dead zone or weaker vertical magnetic fields. For example, if the plasma β (the ratio of gas pressure to magnetic pres- sure) is greater than 3× 106, the diffusion coefficient due to MRI turbulence is less than 10−5 (see model X1b of Okuzumi & Hirose 2011). In such a weakly magnetized disk and for Ts > 0.1 particles, turbulence originating in the dust layer is stronger than MRI turbulence in the dead zone. The wavelength of the most unstable mode must be smaller than the disk radius. The wavelength λ is given in Equation (56) of Y11. The ratio of λ to half of the disk radius r is and is plotted in Figure 5 for disks with the minimum dust abundances Zmin. Note that 2λ/r does not depend on Ts, because α is proportional to Ts. Figure 5 shows that λ is smaller than r/2 if η is less than 3 × 10−2, al- though it is still comparable to r/2 even for η as small as 3 × 10−3. In the analysis by Y11, SC11, and MKI12, the gas disk is assumed to behave as a stationary back- ground, and its velocity profiles are not affected by the gravity of the dust layer. However, the response of the gas to perturbations longer than the gas scale height is not clear. For perturbations with smaller wavelengths to be unstable, larger values of Z than those shown in Fig- ure 2 are required. The wavelength varies as λ ∝ Z −1/3 for Z < (Ceff η)1/2 and λ ∝ Z −3 for Z > (Ceff η)1/2. In the latter regime (Z > (Ceff η)1/2), an increase by a factor of 2 in Z results in an order of magnitude decrease in λ, while in the former regime, reduction in the wavelength requires a large increase in Z. 3. MINIMUM DUST ABUNDANCES FOR VARIOUS DISK MODELS In the previous section, minimum dust abundances for SGI were obtained for given parameters (η, Qg). In this section, we consider several disk models. We adopt power-law profiles for density and temperature distribu- tions of the gas disk, such that Σg = Σg,0r−p AU , T = T0r−q AU , (12) (13) where rAU is the distance from the star measured in AU. In such a disk, Qg ∝ r(2p−q−3)/2 AU . We consider two cases of hot and cold gas disks, T0 = 300K and 150K, while we fix the power-law index q = 1/2. and η ∝ r1−q AU log(2l / r) -0.5 0 103 102 101 g Q -1 -1.5 -2 100 10-4 10-3 h~ 10-2 10-1 Fig. 5. -- Contour map of the normalized wavelength of the most unstable mode 2λ/r on the η-Qg plane for the dust layer with Zmin. The contours are labeled by log(2λ/r). The dashed line shows the locus of Zmin = (Ceff η)1/2. 103 102 101 100 -5 10-4 log(Zmin) S (cid:181) S (cid:181) S (cid:181) S (cid:181) r-1, T0=150K r-1, T0=300K r-1.5, T0=150K r-1.5, T0=300K 0 Mdisk=0.03Msun -1 10AU 1AU 10AU -2 100AU 100AU 0.1AU 1AU -3 0.1AU -4 10-3 h~ 10-2 10-1 Fig. 6. -- Models of the gas disks are plotted on the η-Qg plane. Red and blue lines represent models with shallow and steep density profiles, Σ ∝ r−1 and Σ ∝ r−1.5, respectively. Solid and dashed lines indicate models of cold and hot gas disks, T0 = 150K and T0 = 300K, respectively. Dots on the lines show locations on the disks, 0.1, 1, 10, and 100AU. The disk mass is Mdisk = 0.03M⊙. Contours show log(Zmin). For the density profile, p = 1.0 and 1.5 are adopted and the mass of the gas disk inside 100 AU is varied within Mdisk = 10−2 − 10−1M⊙. Figure 6 shows the variability in the values of η and Qg between the models and with r. In this figure, models with Mdisk = 3 × 10−2M⊙ are shown. For various values of Mdisk, Qg scales as Qg ∝ Mdisk. Figure 7 shows the required minimum dust abun- dances Zmin against rAU for various disk models. For the steep density profile (p = 1.5; blue dashed lines), Zmin ∝ r1/4 AU . (In the regime of Qg ≪ η−1/2 of Equa- tion (10), Zmin ∝ η5/4Q3/2 AU ; in the Qg ≫ η−1/2 regime, Zmin ∝ η2/3Q1/3 AU ). SGI is more viable at the inner part of the disk. For p = 1.0 (red solid lines), Zmin ∝ r1/12 AU only weakly depends on r (in the Qg ≫ η−1/2 regime). The outer part of the disk with Mdisk = 10−1M⊙ and p = −1 is in the Qg ≪ η−1/2 regime, and Zmin decreases as Zmin ∝ r−1/2. Even in g ∝ r1/4 g ∝ r1/4 6 10-1 i n m Z 10-2 (a) T0=150K Mdisk=0.01 Msun Mdisk=0.03 Msun Mdisk=0.1 Msun 0.01 0.03 0.1 10-3 0.1 1 r (AU) r-1 r-1.5 S (cid:181) S (cid:181) 10 100 10-1 (b) T0=300K Mdisk=0.01 Msun Mdisk=0.03 Msun i n m Z 10-2 Mdisk=0.1 Msun 0.1 1 r (AU) 0.01 0.03 0.1 r-1 r-1.5 S (cid:181) S (cid:181) 10 100 Fig. 7. -- Minimum dust abundances Zmin for various disk mod- els. (a) Cold disks (T0 = 150K). (b) Hot disks (T0 = 300K). Red solid and blue dashed lines represent models with shallow and steep density profiles, Σg ∝ r−1 and Σg ∝ r−1.5, respectively. the hot gas disks (T0 = 300K) with a mass as small as Mdisk = 10−2M⊙, the required dust abundances for SGI are less than 0.1. If the temperature at the disk midplane is as cold as T0 ≈ 150K, as expected in passive disks that are heated only by stellar radiation (Chiang & Goldreich 1997; Tanaka et al. 2005), SGI operates for Zmin < 0.05 at 10AU of disks with Mdisk = 10−2M⊙. 4. DISCUSSION Analysis in this paper assumes that turbulence is in- duced in the dust layer. If the dust abundance at the midplane fmid is larger than unity (Z & (Ceff η)1/2), tur- bulence weakens with increasing Z as α ∝ Z −2 (Equa- tion (5)). The growth timescale of SGI rapidly decreases as Tgrow ∝ Z −4 (Equation (6)). A slight increase in Z makes SGI viable. Y11 also estimated the minimum dust abundance Zη required for SGI (Equation (69) of Y11), assuming that the turbulent diffusion parameter of a dense dust layer (fmid & 1) was αη ≈ Ts η (Equation (68) of Y11), independent of Z. This estimate for αη adopted by Y11 corresponds to the maximum value of α(Z) at Z ≈ (Ceff η)1/2 in our estimate (Equation (5)). If α had a constant value αη for all Z, the condition Tgrow < Tdrift would be Z > Zη ∼ η5/6Q2/3 g , (14) Takeuchi & Ida where we used fmid = Zhg/hd and hd/hg ∼ η1/2. Note that this Equation (14) differs from Equation (69) of Y11, because we use Equation (7), in which Tdrift ∝ fmid, while Y11 assumes Tdrift ∝ f 2 mid (see discussion following Equation (A6) below). As discussed in T12, accumula- tion of the dust increases the inertia of the dust layer and decelerates the inward drift of the dust, resulting in weaker turbulence. Considering α ∝ Z −2, the condition Tgrow < Tdrift becomes Z > Zmin ∼ η2/3Q1/3 g , (15) as shown in Equation (10). We ignore numerical co- efficients such as Ceff and Cη. These two conditions (Equations (14) and (15)) coincide when fmid = 1 (i.e., Zη = Zmin ≈ η1/2) or when Qg = Qg,0(η) ≡ η−1/2 (i.e., on the dashed line in Figures 2-5). If the disk's Qg is greater than Qg,0, then the latter condition (Equa- tion (15)) permits smaller dust abundances for SGI by a factor of Zmin/Zη = (Qg/Qg,0)−1/3. For example, at 1AU on a disk with Mdisk = 0.03M⊙, Σg ∝ r−1, and T0 = 300K, Equation (15) gives Zmin = 0.11, which is smaller by a factor of 2 than the value Zη = 0.23 pre- dicted by Equation (14). Considering the Z dependence of α reduces the required minimum value Zmin for SGI as compared with that derived by Y11. 5. SUMMARY We analyze the condition for SGI occurrence in the dust layer. The growth timescale of the instability must be shorter than the radial drift timescale of the dust. The growth timescale decreases as turbulence weakens. The necessary condition for SGI is obtained by consider- ing turbulence induced in the dust layer. Using the tur- bulence strength estimated from the energy supply rate from the accreting dust, the minimum dust abundances for SGI, Zmin, are derived as a function of the Toomre Qg parameter of the gas disk and the deviation fraction η of the gas velocity from the Keplerian value. If the dust particles are small and their stopping time is less than the Keplerian time (Ts < 1), Zmin is independent of Ts. For disks with Qg ∼ 102 and η ∼ 10−2, SGI occurs if Z & 0.1. The required Z decreases with decreasing Qg and η, be- coming as small as 0.01 for Qg ∼ 10 and η ∼ 10−3. Such an increase in Z from the solar abundance is expected to occur through several processes, including the radial drift of the dust and dispersal of the gas from the disk (Youdin & Shu 2002; Takeuchi & Lin 2002; Takeuchi et al. 2005). Therefore, SGI provides a possible route for small particles to directly collapse to planetesimals in an initially laminar disk. If the gas disk were globally tur- bulent via MRI, for example, the required Z would be larger than Zmin estimated in this paper. We thank Takayuki Muto, Satoshi Okuzumi, Hidekazu Tanaka, Shugo Michikoshi, and Naoki Ishitsu for use- ful discussions. We also thank an anonymous referee for helpful comments. This work was supported by Grants-in-Aid for Scientific Research, Nos. 20244013 and 20540232 from the Ministry of Education, Culture, Sports, Science, and Technology, Japan. Secular Gravitational Instability of the Dust Layer 7 10-5 10-6 a 10-7 10-8 Ts=10-1 Ts=10-2 Ts=10-3 (a) diffusion parameter 101 100 10-1 ) s / m c ( y t i c o e v l (b) drift velocity Numerical Approximation 10-9 10-4 Approximation Numerical 10-3 10-2 Z 10-1 100 10-2 10-3 10-2 10-1 100 Z Fig. 8. -- Comparison of approximate expressions with numerical results for (a) turbulent diffusion parameter α and (b) dust radial drift velocity vd,r. Solid lines represent approximate expressions (A4) and (A6), and dashed lines are numerically calculated by T12. (a) Diffusion parameter α is plotted for Ts = 10−3,10−2, and 10−1. (b) Radial drift velocity vd,r is plotted for Ts = 10−3. APPROXIMATE EXPRESSIONS OF DIFFUSION PARAMETER AND RADIAL DRIFT RATE APPENDIX In this Appendix, we evaluate an approximate expression of the turbulent diffusion parameter α for small particle limit (Ts ≪ 1) according to T12. We consider turbulence excited via KH or streaming instabilities of the dust layer. The strength of the turbulence is controlled by the dust accretion rate. The dust particles accrete toward the star either due to the gas drag on individual particles (Nakagawa et al. 1986) or collective drag exerted on the entire dust layer (Weidenschilling 2003; Youdin & Chiang 2004). Individual drag dominates if the dust-to-gas ratio in the dust layer is less than unity (fmid . 1), while collective drag is important if fmid & 1. First, we consider individual drag (for fmid . 1). Using the formula derived by Nakagawa et al. (1986; or Equation (9) of T12, hereafter Equation (T9)), the accretion velocity of small dust particles (Ts . 1) is vd,r = −2TsηvK . K and gg = −(1− 2η)rΩ2 (A1) To satisfy angular momentum conservation, the gas moves outward with the velocity vg,r = −vd,rρd/ρg, where ρd (ρg) is the dust (gas) density. The effective gravities (including the pressure gradient force) acting on the dust and gas are gd = −rΩ2 K, respectively. The net work done by these effective gravities on a unit surface area of the dust layer is calculated as ∆Edrag ∼ R (ρdgdvd,r + ρgggvg,r)dz ∼ η2v2 KΩKTsΣd (Equation (T12)). The energy dissipation due to turbulence is estimated as ∆Eturb ∼ Σlayeru2 eddy/τeddy, where Σlayer is the surface density of the dust layer, and ueddy and τeddy are the typical values for the velocity and turnover time of the largest eddies, respectively. Adopting the so-called "α prescription", we evaluate ueddy ∼ √αcg and τeddy ∼ Ω−1 K (Cuzzi et al. 2001). For fmid . 1, the column density of the dust layer is dominated by the gas, and thus Σlayer ∼ Σghd/hg. Here, hd/hg is the ratio of the scale heights of the dust layer and gas disk, and is given by hd/hg ≈ pα/Ts (Equation (T4)). Then, the energy dissipation rate is ∆Eturb ∼ pα3/Tsh2 an efficiency parameter for the energy supply, ∆Eturb ∼ Ceff ∆Edrag, gives KΣg. Equating the energy supply and dissipation rates of turbulence, with gΩ3 α ∼ (Ceff ηZ)2/3Ts . (A2) The collective drag force in the θ-direction on a unit surface area is estimated by the plate drag approximation for fmid & 1 as Pθz ∼ ρgν∂vg,θ/∂z ∼ −ρgνηvK/hd, where ν is the turbulent viscosity, and the velocity shear in the vertical direction ∂vg,θ/∂z is estimated as ηvK/hd. This drag causes accretion of the dust layer and an outward motion of the upper gas layer. The dust layer loses energy ΩKrPθz and the upper gas layer gains energy (1−η)ΩKrPθz, where the work done by gas pressure is taken into account. The net energy liberation rate is ∆Evis ∼ −ηΩKrPθz ∼ η2v2 KΩKTsΣd/fmid (Equation (T18)), where we used ν ∼ h2 dTsΩK and hd = hgΣd/(Σgfmid) (Equations (T16) and (T6)). The energy dissipation rate is ∆Eturb ∼ Σlayeru2 KΣd, where Σlayer ∼ Σd is dominated by the dust. Equating ∆Eturb ∼ Ceff ∆Evis gives (A3) where we use fmid ≈ ZpTs/α (Equation (T6)). Expressions (A2) and (A3) are the same as Equation (T28), except for numerical factors and the adoption of a slightly steeper power-law index (δ = 1 in Equation (A3) instead of δ = 0.94 in Equation (T28)) to simplify analytical expressions. Connecting these expressions gives α ∼ (Ceff ηZ −1)2Ts , eddy/τeddy ∼ αh2 gΩ3 α ≈h(C1Ceff ηZ)− 2 3 +(cid:0)C2Ceff ηZ −1(cid:1)−2i−1 Ts , (A4) 8 Takeuchi & Ida where the numerical factors C1 = 1.0 and C2 = 1.6 are set to fit the numerical result of T12. In Figure 8(a), the approximate expression (A4) is compared to the numerical result of T12. The approximation overestimates α around Z = √Ceff η and underestimates it for large Z. However, the error is less than 2 for 10−4 < Z < 10−1, and the overall behavior appears to be acceptable. The midplane dust-to-gas ratio fmid = Zhg/hd ≈ ZpTs/α is, from Equations (A2) and (A3), fmid ∼ [Z 2/(Ceff η)]1/3 for fmid . 1, and fmid ∼ Z 2/(Ceff η) for fmid & 1. Connecting these two expressions gives an approximate formula for fmid as C2Ceff η(cid:19) . From the above Equation (A5), it is seen that fmid ≈ 1 when Z ≈ (Ceff η)1/2. The drift velocity of the dust in the limit of fmid ≪ 1 and Ts ≪ 1 is given in Equation (A1). In the limit of fmid ≫ 1, vd,r ∼ rPθz/(vKΣd) ∼ ηvKTs/fmid (Equation (T51)). Connecting these limits, we estimate fmid ≈(cid:18) Z 2 +(cid:18) Z 2 (A5) 1 3 C1Ceff η(cid:19) vd,r ≈ − 2TsηvK fmid + 1 , (A6) which gives the dust drift timescale given in Equation (7). This approximate expression is compared with the numerical calculation by T12 in Figure 8(b), which shows that the above expression provides a good approximation. Note that vd,r ∝ f −1 mid for fmid ≫ 1. Equation (2.11) of Nakagawa et al. (1986) shows that, due to individual drag, the radial velocity of the dust particles at the midplane is proportional to f −2 mid. However, the average velocity of the dust layer, R ρdvd,rdz/Σd, is approximately proportional to f −1 mid, provided that the dust density profile ρd(z) is Gaussian. The drift velocity due to collective drag is also proportional to f −1 mid, as shown in Equation (A6). Thus, for fmid > 1, the radial drift time derived from the particle velocity at the midplane (Tdrift ∝ f 2 mid), which is adopted in Y11, SC11, and MKI12, overestimates the average drift time (Tdrift ∝ fmid), resulting in a less strict condition for Tgrow < Tdrift than reality. Bai, X.-N., & Stone, J. M. 2010, ApJ, 722, 1437 Chiang, E. I., & Goldreich, P. 1997, ApJ, 490, 368 Chiang, E., & Youdin, A. N. 2010, Annual Review of Earth and Planetary Sciences, 38, 493 Shariff, K., & Cuzzi, J. N. 2011, ApJ, 738, 73 (SC11) Takeuchi, T., & Lin, D. N. C. 2002, ApJ, 581, 1344 Takeuchi, T., Clarke, C. J., & Lin, D. N. C. 2005, ApJ, 627, 286 Takeuchi, T., Muto, T., Okuzumi, S., Ishitsu, N., & Ida, S. 2012, Cuzzi, J. N., Hogan, R. C., Paque, J. M., & Dobrovolskis, A. R. ApJ, 744, 101 (T12) REFERENCES 2001, ApJ, 546, 496 Fleming, T., Stone, J. M. 2003, ApJ, 585, 908 Goldreich, P., & Ward, W. R. 1973, ApJ, 183, 1051 Johansen, A., Henning, Th., & Klahr, H. 2006, ApJ, 643, 1219 Lee, A. T., Chiang, E., Asay-Davis, X., & Barranco, J. 2010, ApJ, 718, 1367 Michikoshi, S., Kokubo, E., & Inutsuka S.-i. 2012, ApJ, 746, 35 (MKI12) Nakagawa, Y., Sekiya, M., & Hayashi, C. 1986, Icarus, 67, 375 Okuzumi, S., & Hirose, S. 2011, ApJ, 742, 65 Sekiya, M. 1983, Prog. Theor. Phys., 69, 1116 Sekiya, M. 1998, Icarus, 133, 298 Tanaka, H., Himeno, Y., & Ida, S. 2005, ApJ, 625, 414 Ward, W. R. 2000, in Origin of the Earth and Moon, ed. R. M. Canup & K. Righter (Tucson, AZ: Univ. Arizona Press), 75 Weidenschilling, S. J. 2003, Icarus, 165, 438 Youdin, A. N. 2005a, arXiv:astro-ph/0508659 Youdin, A. N. 2005b, arXiv:astro-ph/0508662 Youdin, A. N. 2011, ApJ, 731, 99 (Y11) Youdin, A. N., & Chiang, E. I. 2004, ApJ, 601, 1109 Youdin, A. N., & Shu, F. H. 2002, ApJ, 580, 494
0912.2019
1
0912
2009-12-10T15:27:31
Giant Planets
[ "astro-ph.EP" ]
We review the interior structure and evolution of Jupiter, Saturn, Uranus and Neptune, and extrasolar giant planets with particular emphasis on constraining their global composition.
astro-ph.EP
astro-ph
Giant Planets Tristan Guillot Observatoire de la Cote d’Azur Laboratoire Cassiop´ee, CNRS UMR 6202 BP 4229 06304 Nice Cedex 4 France [email protected] Daniel Gautier Observatoire de Paris LESIA, CNRS FRE 2461 5 pl. J. Janssen 92195 Meudon Cedex France [email protected] To be published in Treatise on Geophysics, G. Shubert, T. Spohn Eds 9 0 0 2 c e D 0 1 . ] P E h p - o r t s a [ 1 v 9 1 0 2 . 2 1 9 0 : v i X r a 1 Contents 1 Introduction 2 Observations and global properties 2.1 Visual appearances . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.2 Gravity fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.3 Magnetic fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.4 Atmospheric compositions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.4.1 Hydrogen and helium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.4.2 Heavy elements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.5 Energy balance and atmospheric temperature profiles 2.6 Atmospheric dynamics: winds and weather 2.7 Moons and rings 2.8 Extrasolar planets 3 The calculation of interior and evolution models 3.1 Basic equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2 High pressure physics & equations of state . . . . . . . . . . . . . . . . . . . . . . . 3.3 Heat transport . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.4 The contraction and cooling histories of giant planets . . . . . . . . . . . . . . . . . 3.5 Mass-radius relation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.6 Rotation and the figures of planets . . . . . . . . . . . . . . . . . . . . . . . . . . . 4 Interior structures and evolutions 4.1 Jupiter and Saturn . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2 Uranus and Neptune . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Irradiated giant planets 5 Implications for planetary formation models 6 Future prospects 3 4 4 5 6 7 8 8 10 11 12 13 14 14 15 18 18 22 23 24 24 27 28 34 35 2 The Giant Planets Tristan Guillot Observatoire de la Cote d’Azur Daniel Gautier Observatoire de Paris We review the interior structure and evolution of Jupiter, Saturn, Uranus and Neptune, and extrasolar giant planets with particular emphasis on constraining their global composition. Abstract KEYWORDS: Giant planets, extrasolar planets, Jupiter, Saturn, Uranus, Neptune, planet formation 1 Introduction In our Solar System, four planets stand out for their sheer mass and size. Jupiter, Saturn, Uranus and Neptune indeed qualify as “giant planets” because they are larger than any terrestrial planet and much more massive than all other objects in the Solar System, except the Sun, put together (fig. 1). Because of their gravitational might, they have played a key role in the formation of the Solar System, tossing around many objects in the system, preventing the formation of a planet in what is now the asteroid belt, and directly leading to the formation of the Kuiper belt and Oort cloud. They also retain some of the gas (in particular hydrogen and helium) that was present when the Sun and its planets formed and are thus key witnesses in the search for our origins. Because of a massive envelope mostly made of hydrogen helium, these planets are fluid, with no solid or liquid surface. In terms of structure and composition, they lie in between stars (gaseous and mostly made of hydrogen and helium) and smaller terrestrial planets (solid and liquid and mostly made of heavy elements), with Jupiter and Saturn being closer to the former and Uranus and Neptune to the latter. The discovery of many extrasolar planets of masses from a few hundreds down to a few Earth masses and the possibility to characterize them by the measurement of their mass and size prompts a more general definition of giant planets. For this review, we will adopt the following: “a giant planet is a planet mostly made of hydrogen and helium and too light to ignite deuterium fusion”. This is purposedly relatively vague – depending on whether the inventory is performed by mass or by atom or molecule, Uranus and Neptune may be included or left out of the category –. Note that Uranus and Neptune are indeed relatively different in structure than Jupiter and Saturn and are generally referred to as “ice giants”, due to an interior structure that is consistent with the presence of mostly “ices” (a mixture formed from the condensation in the protoplanetary disk of low refractivity materials such as H2O, CH4 and NH3, and brought to the high pressure conditions of planetary interiors –see below). 3 Figure 1: An inventory of hydrogen and helium and all other elements (“heavy elements”) in the Solar System excluding the Sun (the Sun has a total mass of 332, 960 M⊕, including about 5000 M⊕ in heavy elements, 1 M⊕ being the mass of the Earth). The precise amount of heavy elements in Jupiter (10 − 40 M⊕) and Saturn (20 − 30 M⊕) is uncertain (see § 4.1). Globally, this definition encompasses a class of objects that have similar properties (in par- ticular, a low viscosity and a non-negligible compressibility) and inherited part of their material directly from the same reservoir as their parent star. These objects can thus be largely studied with the same tools, and their formation is linked to that of their parent star and the fate of the circumstellar gaseous disk present around the young star. We will hereafter present some of the key data concerning giant planets in the Solar System and outside. We will then present the theoretical basis for the study of their structure and evolution. On this basis, the constraints on their composition will be discussed and analyzed in terms of consequences for the models of planet formation. 2 Observations and global properties 2.1 Visual appearances In spite of its smallness, the sample of four giant planets in our Solar System exhibits a large variety of appearances, shapes, colors, variability...etc. As shown by fig. 2, all four giant planets are flattened by rotation and exhibit a more or less clear zonal wind pattern, but the color of their visible atmosphere is very different (this is due mostly to minor species in the high planetary atmosphere), their clouds have different compositions (ammonia for Jupiter and Saturn, methane for Uranus and Neptune) and depths, and their global meteorology (number of vortexes, long- lived anticyclones such as Jupiter’s Great Red Spot, presence of planetary-scale storms, convective activity) is different from one planet to the next. We can presently only wonder about what is in store for us with extrasolar giant planets since 4 Figure 2: Photographs of Jupiter, Saturn, Uranus and Neptune. we cannot image them. But with orbital distances that can be as close as 0.02 AU, a variety of masses, sizes, and parent stars, we should expect to be surprised! 2.2 Gravity fields The mass of our giant planets can be obtained with great accuracy from the observation of the motions of their natural satellites: 317.834, 95.161, 14.538 and 17.148 times the mass of the Earth (1 M⊕ = 5.97369 × 1027 g) for Jupiter, Saturn, Uranus and Neptune, respectively. More precise measurements of their gravity field can be obtained through the analysis of the trajectories of spacecrafts during flyby, especially when they come close to the planet and preferably in a near-polar orbit. The gravitational field thus measured departs from a purely spherical function due to the planets’ rapid rotation. The measurements are generally expressed by expanding the components of the gravity field on Legendre polynomials Pi of progressively higher orders: Vext(r, θ) = − GM r (1 − ∞ r (cid:19)i Xi=1(cid:18) Req JiPi(cos θ)) , (1) where Vext(r, θ) is the gravity field evaluated outside the planet at a distance r and colatitude θ, Req is the equatorial radius, and Ji are the gravitational moments. Because the giant planets are very close to hydrostatic equilibrium the coefficients of even order are the only ones that are not negligible. We will see how these gravitational moments, as listed in table 1, help us constrain the planets’ interior density profiles. Table 1 also indicates the radii obtained with the greatest accuracy by radio-occultation ex- periments. An important consequence obtained is the fact that these planets have low densities, from 0.688 g cm−3 for Saturn to 1.64 g cm−3 for Neptune, to be compared with densities of 3.9 to 5.5 g cm−3 for the terrestrial planets. Considering the compression that strongly increases with mass, one is led naturally to the conclusion that these planets contain an important proportion of light materials including hydrogen and helium. It also implies that Uranus and Neptune which are less massive must contain a relatively larger proportion of heavy elements than Jupiter and Saturn. This may lead to a sub-classification between the hydrogen-helium giant planets Jupiter and Saturn, and the “ice giants” or “sub giants” Uranus and Neptune. The planets are also relatively fast rotators, with periods of ∼ 10 hours for Jupiter and Saturn, and ∼ 17 hours for Uranus and Neptune. The fact that this fast rotation visibly affects the figure (shape) of these planets is seen by the significant difference between the polar and equatorial radii. It also leads to gravitational moments that differ significantly from a nul value. However, it is 5 Table 1: Characteristics of the gravity fields and radii Jupiter Saturn Uranus Neptune −5.84(5)a 0.31(20)a 3.57297(41)j 0.08923(5) 0.258 −0.354(41)c −0.28(22)d . . . 6.206(4)l 0.02951(5) 0.230 1.0243542(31)d 2.4766(15)g 2.4342(30)g 2.4625(20)i 1.6377(40) 0.3539(10)d . . . 5.800(20)m 0.02609(23) 0.241 5.684640(30)b 6.0268(4)f 5.4364(10)f 5.8210(6)h 0.6880(2) 1.632425(27)b −9.397(28)b 0.867(97)b 3.83624(47) ?j,k 0.15491(10) 0.220 0.8683205(34)c 2.5559(4)g 2.4973(20)g 2.5364(10)i 1.2704(15) 0.35160(32)c 18.986112(15)a 7.1492(4)e 6.6854(10)e 6.9894(6)h 1.3275(4) 1.4697(1)a M × 10−26 [kg] Req × 10−7 [m] Rpol × 10−7 [m] R × 10−7 [m] ρ × 10−3 [kg m−3] J2 × 102 J4 × 104 J6 × 104 Pω × 10−4 [s] q C/M R2 eq The numbers in parentheses are the uncertainty in the last digits of the given value. The value of the gravitational constant used to calculate the masses of Jupiter and Saturn is G = 6.67259 × 10−11 N m2 kg−2 (Cohen and Taylor, 1987). The values of the radii, density and gravitational moments correspond to the one bar pressure level (1 bar= 105 Pa). a Campbell and Synnott (1985) b Jacobson et al. (2006) c Anderson et al. (1987) d Tyler et al. (1989) e Lindal et al. (1981) f Lindal et al. (1985) g Lindal (1992a) h From 4th order figure theory i (2Req + Rpol)/3 (Clairaut’s approximation) j Davies et al. (1986) k This measurement from the Voyager era is now in question and values up to 38826 s have been proposed (see § 2.3) l Warwick et al. (1986) m Warwick et al. (1989) important to stress that there is no unique rotation frame for these fluid planets: atmospheric zonal winds imply that different latitude rotate at different velocities (see § 2.6), and the magnetic field provides an other rotation period. Because the latter is tied to the deeper levels of the planet, it is believed to be more relevant when interpreting the gravitational moments. The rotation periods listed in Table 1 hence correspond to that of the magnetic field. The case of Saturn appears to be complex and is discussed in the next section. 2.3 Magnetic fields As the Earth, the Sun and Mercury, our four giant planets possess their own magnetic fields, as shown by the Voyager 2 measurements. The structures of these magnetic fields are very different from one planet to another and the dynamo mechanism that generates them is believed to be related to convection in their interior but is otherwise essentially unknown (see Stevenson, 1983, for a review). The magnetic field B is generally expressed in form of a development in spherical harmonics of 6 the scalar potential W , such that B = −∇W : W = a ∞ Xn=1(cid:16) a r(cid:17)n+1 n Xm=0 {gm n cos(mφ) + hm n sin(mφ)} P m n (cos θ). (2) r is the distance to the planet’s center, a its radius, θ the colatitude, φ the longitude and P m n the associated Legendre polynomials. The coefficients gm n and hm n are the magnetic moments that characterize the field. They are expressed in magnetic field units. One can show that the first coefficients of relation (2) (for n = 0 and n = 1) correspond to the potential of a magnetic dipole such that W = M · r/r3 of moment: M = a3n(cid:0)g0 1(cid:1)2 1(cid:1)2 +(cid:0)g1 +(cid:0)h1 1(cid:1)2o1/2 . (3) Jupiter and Saturn have magnetic fields of essentially dipolar nature, of axis close to the rotation axis (g0 1 is much larger than the other harmonics); Uranus and Neptune have magnetic fields that are intrinsically much more complex. To provide an idea of the intensity of the magnetic fields, the value of the dipolar moments for the four planets are 4.27 Gauss R3 S, 0.23 Gauss R3 U, 0.133 Gauss R3 N, respectively (Connerney et al., 1982; Acuna et al., 1983; Ness et al., 1986, 1989). A true surprise from Voyager that has been confirmed by the Cassini-Huygens mission is that Saturn’s magnetic field is axisymetric to the limit of the measurement accuracy: Saturn’s magnetic and rotation axes are perfectly aligned. Voyager measurements indicated nevertheless a clear signature in the radio signal at 10h 39min 22s believed to be a consequence of the rotation of the magnetic field. Determinations of a magnetic anomaly and new measurements by Cassini have since considerably blurred the picture, and the interpretation of the measurements have become unclear, with a deep rotation period evaluated between that of Voyager and as slow as 10h 47min 6s (Gurnett et al., 2005; Giampieri et al., 2006) (see also Galopeau and Lecacheux, 2000; Cecconi and Zarka, 2005). Note that models discussed hereafter have not yet included this additional uncertainty. J, 0.21 Gauss R3 2.4 Atmospheric compositions In fluid planets, the distinction between the atmosphere and the interior is not obvious. We name “atmosphere” the part of the planet which can directly exchange radiation with the exterior environment. This is also the part which is accessible by remote sensing. It is important to note that the continuity between the atmosphere and the interior does not guarantee that compositions measured in the atmosphere can be extrapolated to the deep interior, even in a fully convective environment: Processes such as phase separations (e.g. Salpeter, 1973; Stevenson and Salpeter, 1977b; Fortney and Hubbard, 2003), phase transitions (e.g. Hubbard, 1989), chemical reactions (e.g. Fegley and Lodders, 1994) can occur and decouple the surface and interior compositions. Furthermore, imperfect mixing may also occur, depending on the initial conditions (e.g. Stevenson, 1985). The conventional wisdom is however that these processes are limited to certain species (e.g. helium) or that they have a relatively small impact on the global abundances, so that the hydrogen- helium envelopes may be considered relatively uniform, from the perspective of the global abun- dance in heavy elements. We first discuss measurements made in the atmosphere before inferring interior compositions from interior and evolution models. 7 2.4.1 Hydrogen and helium The most important components of the atmospheres of our giant planets are also among the most difficult to detect: H2 and He have a zero dipolar moment and hence absorb very inefficiently visible and infrared light. Their infrared absorption becomes important only at high pressures when collision-induced absorption becomes significant (e.g. Borysow et al., 1997). On the other hand, lines due to electronic transitions correspond to very high altitudes in the atmosphere, and bear little information on the structure of the deeper levels. The only robust result concerning the abundance of helium in a giant planet is by in situ measurement by the Galileo probe in the atmosphere of Jupiter (von Zahn et al., 1998). The helium mole fraction (i.e. number of helium atoms over the total number of species in a given volume) is qHe = 0.1359 ± 0.0027. The helium mass mixing ratio Y (i.e. mass of helium atoms over total mass) is constrained by its ratio over hydrogen, X: Y /(X + Y ) = 0.238 ± 0.05. This ratio is by coincidence that found in the Sun’s atmosphere, but because of helium sedimentation in the Sun’s radiative zone, it was larger in the protosolar nebula: Yproto = 0.275 ± 0.01 and (X + Y )proto ≈ 0.98 (e.g. Bahcall et al., 1995). Less helium is therefore found in the atmosphere of Jupiter than inferred to be present when the planet formed. We will discuss the consequences of this measurement later: let us mention that the explanation invokes helium settling due to a phase separation in the interiors of massive and cold giant planets. Helium is also found to be depleted compared to the protosolar value in Saturn’s atmo- sphere. However, in this case the analysis is complicated by the fact that Voyager radio oc- cultations apparently led to a wrong value. The current adopted value is now Y = 0.18 − 0.25 (Conrath and Gautier, 2000), in agreement with values predicted by interior and evolution models (Guillot, 1999b; Hubbard et al., 1999). Finally, Uranus and Neptune are found to have near- protosolar helium mixing ratios, but with considerable uncertainty. 2.4.2 Heavy elements The abundance of other elements than hydrogen and helium (that we will call hereafter “heavy el- ements”) bears crucial information for the understanding of the processes that led to the formation of these planets. The most abundant heavy elements in the envelopes of our four giant planets are O, C, N, S. It is possible to model the chemistry of gases in the tropospheres from the top of the convective zone down to the 2000 K temperature level (Fegley and Lodders, 1994). Models conclude that, what- ever the initial composition in these elements of planetesimals which collapsed with hydrogen onto Jupiter and Saturn cores during the last phase of the planetary formation, C in the upper tropo- spheres of giant planets is mainly in the form of gaseous CH4, N in the form of NH3, S in the form of H2S, and O in the form of H2O. All these gases, but methane in Jupiter and Saturn, condense in the upper troposphere, but vaporize at deeper levels when the temperature increases. Interest- ingly enough, noble gases are not expected to condense even at the cold tropopause temperatures of Uranus and Neptune. The mass spectrometer aboard the Galileo atmospheric probe has performed in situ measure- ments of Ar, Kr, Xe, CH4, NH3, H2S, and H2O in the troposphere of Jupiter. C, N, and S were found to be oversolar by a factor 3 to 4 (Wong et al., 2004), which was not unexpected because condensation of nebula gases results in enriching icy grains and planetesimals. The surprise came from Ar, Kr, Xe, which were expected to be solar because they are difficult to condense, but turned out to be oversolar by factors 2 to 4 (Owen et al., 1999; Wong et al., 2004). One exception among these enriched species was neon, which was found to be significantly undersolar, but was 8 Table 2: Main gaseous components of heavy elements measured in the tropo- sphere of giant planets Species Mixing ratio/H2 References Comments Jupiter Saturn Uranus Neptune CH4 NH3 H2S H20 36Ar 84Kr 132Xe CH4 NH3 H2S CH4 H2S CH4 H2S H2O (2.37 ± 0.37) × 10−3 Wong et al. (2004) (6.64 ± 2.34) × 10−3 Wong et al. (2004) (8.9 ± 2.1) × 10−3 Wong et al. (2004) (4.9 ± 1.6) × 10−4 Wong et al. (2004) (6.1 ± 1.2) × 10−6 Atreya et al. (1999) (1.84 ± 0.37) × 10−9 Atreya et al. (1999) (4.9 ± 1.0) × 10−11 Atreya et al. (1999) GPMS on Galileo(1) idem idem idem; region not well mixed idem idem idem (4.3 ± 1) × 10−3 (1 ± 1) × 10−4 (2.2 ± 0.3) × 10−4 Flasar et al. (2005) Briggs and Sackett (1989) Briggs and Sackett (1989) CIRS on Cassini (3) Ground-based microwave (4) idem (3.3 ± 1.1) × 10−2 (1 ± 1) × 10−4 Gautier et al. (1995) Briggs and Sackett (1989) Compilation from ground-based observations Ground-based microwave (4) (3.3 ± 1.1) × 10−2 (7.5 ± 3.25) × 10−4 7.7 × 10(−1) Gautier et al. (1995) de Pater et al. (1991) Lodders and Fegley (1994) Compilation from ground-based observations Ground based microwave (5) Inferred from CO (6) (1) Galileo Probe Mass Spectrometer aboard the atmospheric probe in Jupiter (2) The signal stopped at the 22 bar levels prior to have reached a constant value. It is currently believed that the region where the probe made measurementds was atypically dry and that the bulk abundance of H2O in Jupiter has not been measured. (3) Composite Infra Red Spectrometer aboard the Cassini spacecraft (4) Ground based measurements of the microwave continuum. The result is somewhat uncertain due to the difficulty to precisely estimate opacities of absorbing species (5) Ground based microwave measurements. An oversolar H2S abundance is required to interpret the depletion of NH3 in the upper troposphere (6) Inferred from the microwave detection of CO in the troposphere. Note that the validity of the approach is questionned by B´ezard et al. (2002). A large amount of water seems to be present anyway in the deep atmosphere of Neptune. The case of Uranus is still uncertain because it is not known so far if CO is present in the troposphere of the planet. predicted to be so because of a capture by the falling helium droplets (Roulston and Stevenson, 1995). Another exception was water, but this molecule is affected by meteorological processes, and the probe was shown to have fallen into a dry region of Jupiter’s atmosphere. Specifically, CH4/H2 has been found oversolar in the four giant planets: the C/H ratio corre- sponding to the measured abundances is always higher than the solar C/H ratio, and in fact appears to be increasing with distance to the Sun. C/H is 3, 7.5, 45 and 45 times solar, in Jupiter, Saturn, Uranus and Neptune, respectively. Note that the quoted enrichments are subject to changes when the solar abundances tables are revised, which happens surprisingly frequently. Except for Jupiter, the determination of the NH3 abundance is more uncertain than that of CH4 because it is model dependant. It is derived from fitting microwave spectra of giant planets which exhibit a continuum opacity, more difficult to model than absorption spectral lines. However, the N/H enrichment seems to be, so far, fairly constant from a planet to another, around a factor 2, so that C/N is higher in Saturn than in Jupiter, and still higher in Uranus and Neptune. H2S has been measured in situ in Jupiter, but in the three other giant planets its large abun- dance is derived from the requirement to deplete NH3 at deeper levels than the saturation one; 9 This scenario has been proposed a long time ago by Gulkis et al. (1978). It implies that S/H is substantially oversolar in Uranus and in Neptune. H2O is difficult to measure in all four giant planets because of its condensation relatively deep. It was hoped that the Galileo probe would provide a measurement of its deep abundance, but the probe fell into one of Jupiter’s 5-microns hot spot, what is now believed to be a dry region mostly governed by downwelling motions (e.g. Showman and Ingersoll, 1998). As a result, and although the probe provided measurements down to 22 bars, well below water’s canonical 5 bar cloud base, it is believed that this measurement of a water abundance equal to a fraction of the solar value is only a lower limit. 2.5 Energy balance and atmospheric temperature profiles Jupiter, Saturn and Neptune are observed to emit significantly more energy than they receive from the Sun (see Table 3). The case of Uranus is less clear. Its intrinsic heat flux Fint is significantly smaller than that of the other giant planets. Detailed modeling of its atmosphere however indicate that Fint ∼> 60 erg cm−2 s−1 (Marley and McKay, 1999). With this caveat, all four giant planets can be said to emit more energy than they receive from the Sun. Hubbard (1968) showed in the case of Jupiter that this can be explained simply by the progressive contraction and cooling of the planets. Table 3: Energy balance as determined from Voyager IRIS dataa. Absorbed power [1016 J s−1] Emitted power [1016 J s−1] Intrinsic power [1016 J s−1] Intrinsic flux [J s−1 m−2] Bond albedo [] Effective temperature [K] 1-bar temperatureb [K] a After Pearl and Conrath (1991) b Lindal (1992b) 0.343±0.032 124.4±0.3 165±5 Jupiter 50.14±2.48 83.65±0.84 33.5±2.6 5.44±0.43 Saturn 11.14±0.50 19.77±0.32 8.63±0.60 2.01±0.14 0.342±0.030 95.0±0.4 135±5 Uranus 0.526±0.037 0.560±0.011 0.034 +0.038 −0.034 0.042 +0.047 −0.042 0.300±0.049 59.1±0.3 76±2 Neptune 0.204±0.019 0.534±0.029 0.330±0.035 0.433±0.046 0.290±0.067 59.3±0.8 72±2 A crucial consequence of the presence of an intrinsic heat flux is that it requires high internal temperatures (∼ 10, 000 K or more), and that consequently the giant planets are fluid (not solid) (Hubbard (1968); see also Hubbard et al. (1995)). Another consequence is that they are essentially convective, and that their interior temperature profile are close to adiabats. We will come back to this in more detail. The deep atmospheres (more accurately tropospheres) of the four giant planets are indeed ob- served to be close to adiabats, a result first obtained by spectroscopic models (Trafton, 1967), then verified by radio-occultation experiments by the Voyager spacecrafts, and by the in situ measure- ment from the Galileo probe (fig. 3). The temperature profiles show a temperature minimum, in a region near 0.2 bar called the tropopause. At higher altitudes, in the stratosphere, the temperature gradient is negative (increasing with decreasing pressure). In the regions that we will be mostly concerned with, in the troposphere and in the deeper interior, the temperature always increases with depth. It can be noticed that the slope of the temperature profile in fig 3 becomes almost constant when the atmosphere becomes convective, at pressures of a few tens of bars, in the four 10 giant planets. Figure 3: Atmospheric temperatures as a function of pressure for Jupiter, Saturn, Uranus and Neptune, as obtained from Voyager radio-occultation experiments (see Lindal, 1992b). The dotted line corresponds to the temperature profile retrieved by the Galileo probe, down to 22 bar and a temperature of 428 K (Seiff et al., 1998). It should be noted that the 1 bar temperatures listed in table 3 and the profiles shown in Fig. 3 are retrieved from radio-occultation measurements using a helium to hydrogen ratio which, at least in the case of Jupiter and Saturn, was shown to be incorrect. The new values of Y are found to lead to increased temperatures by ∼ 5 K in Jupiter and ∼ 10 K in Saturn (see Guillot, 1999a). However, the Galileo probe found a 1 bar temperature of 166 K (Seiff et al., 1998), and generally a good agreement with the Voyager radio-occultation profile with the wrong He/H2 value. When studied at low spatial resolution, it is found that all four giant planets, in spite of their inhomogeneous appearances, have a rather uniform brightness temperature, with pole-to-equator latitudinal variations limited to a few kelvins (e.g. Ingersoll et al., 1995). However, in the case of Jupiter, some small regions are known to be very different from the average of the planet. This is the case of hot spots, which cover about 1% of the surface of the planet at any given time, but contribute to most of the emitted flux at 5 microns, due to their dryness (absence of water vapor) and their temperature brightness which can, at this wavelength, peak to 260 K. 2.6 Atmospheric dynamics: winds and weather The atmospheres of all giant planets are evidently complex and turbulent in nature. This can for example be seen from the mean zonal winds (inferred from cloud tracking), which are very rapidly varying functions of the latitude (see e.g. Ingersoll et al., 1995): while some of the regions rotate 11 at the same speed as the interior magnetic field (in the so-called “system III” reference frame), most of the atmospheres do not. Jupiter and Saturn both have superrotating equators (+100 and +400 m s−1 in system III, for Jupiter and Saturn, respectively), Uranus and Neptune have subrotating equators, and superrotating high latitude jets. Neptune, which receives the smallest amount of energy from the Sun has the largest peak-to-peak latitudinal variations in wind velocity: about 600 m s−1. It can be noted that, contrary to the case of the strongly irradiated planets to be discussed later, the winds of Jupiter, Saturn, Uranus and Neptune, are significantly slower than the planet itself under its own spin (from 12.2 km s−1 for Jupiter to 2.6 km s−1 for Neptune, at the equator). The observed surface winds are believed to be related to motions in the planets’ interiors, which, according to the Taylor-Proudman theorem, should be confined by the rapid rotation to the plane perpendicular to the axis of rotation (e.g. Busse, 1978). Unfortunately, no convincing model is yet capable of modeling with sufficient accuracy both the interior and the surface layers. Our giant planets also exhibit planetary-scale to small-scale storms with very different temporal variations. For example, Jupiter’s great red spot is a 12000 km-diameter anticyclone found to have lasted for at least 300 years (e.g. Simon-Miller et al., 2002). Storms developing over the entire planet have even been observed on Saturn (Sanchez-Lavega et al., 1996). Uranus and Neptune’s storm system has been shown to have been significantly altered since the Voyager era (Rages et al., 2002; Hammel et al., 2005). On Jupiter, small-scale storms related to cumulus-type cloud systems have been observed (e.g. Gierasch et al., 2000; Hueso et al., 2002), and lightning strikes have been monitored by Galileo (e.g. Little et al., 1999). These represent only a small arbitrary subset of the work concerning the complex atmospheres of these planets. It is tempting to extrapolate these observations to the objects outside our Solar System as well. However, it is important to stress that an important component of the variability in the atmospheres of our giant planets is the presence of relatively abundant condensing chemical species: ammonia and water in the case of Jupiter and Saturn, and methane for Uranus and Neptune. These species can only condense in very cold atmospheres, thus providing latent heat to fuel important storms. Depending on their temperatures and compositions, extrasolar planets may or may not possess such important condensing species (e.g. Guillot, 1999b). 2.7 Moons and rings A discussion of our giant planets motivated by the opportunity to extrapolate the results to objects outside our solar system would be incomplete without mentioning the moons and rings that these planets all possess (see chapters by Breuer & Moore, by Peale and by Husmann et al.). First, the satellites/moons can be distinguished from their orbital characteristics as regular or irregular. The first ones have generally circular, prograde orbits. The latter tend to have eccentric, extended, and/or retrograde orbits. These satellites are numerous: After the Voyager era, Jupiter was known to possess 16 satellites, Saturn to have 18, Uranus 20 and Neptune 8. Recent extensive observation programs have seen the number of satellites increase considerably, with a growing list of satellites presently reaching 62, 56, 27 and 13 for Jupiter, Saturn, Uranus and Neptune, respectively. All of the new satellites discovered since Voyager are classified as irregular. The presence of regular and irregular satellites is due in part to the history of planet formation. It is believed that the regular satellites have mostly been formed in the protoplanetary subnebulae that surrounded the giant planets (at least Jupiter and Saturn) at the time when they accreted their envelopes. On the other hand, the irregular satellites are thought to have been captured by 12 the planet. This is for example believed to be the case of Neptune’s largest moon, Triton, which has a retrograde orbit. A few satellites stand out by having relatively large masses: it is the case of Jupiter’s Io, Europa, Ganymede and Callisto, of Saturn’s Titan, and of Neptune’s Triton. Ganymede is the most massive of them, being about twice the mass of our Moon. However, compared to the mass of the central planet, these moons and satellites have very small weights: 10−4 and less for Jupiter, 1/4000 for Saturn, 1/25000 for Uranus and 1/4500 for Neptune. All these satellites orbit relatively closely to their planets. The farthest one, Callisto revolves around Jupiter in about 16 Earth days. The four giant planets also have rings, whose material is probably constantly resupplied from their satellites. The ring of Saturn stands out as the only one directly visible with binoculars. In this particular case, its enormous area allows it to reflect a sizable fraction of the stellar flux arriving at Saturn, and makes this particular ring as bright as the planet itself. The occurrence of such rings would make the detection of extrasolar planets slightly easier, but it is yet unclear how frequent they can be, and how close to the stars rings can survive both the increased radiation and tidal forces. 2.8 Extrasolar planets Figure 4: Masses and orbital distances of the extrasolar planets discovered by 2006. The size of the symbols is proportional to the mass of the parent star (from 0.1 to 4 stellar masses). The color (from white to red and black) is proportional to the stellar metallicity. Stars with metallicities [Fe/H] < 0.2 are shown in white. The radial velocimetry thresholds at 1 and 10 m/s, encompassing the detection limits of most current surveys, are indicated as dashed lines. Transiting planets are highlighted with white crosses. 13 Huge progresses have been made in the field of extrasolar planets since the detection of the first giant planet orbiting a solar-type star by Mayor and Queloz (1995). More than 200 planets are known at the time of this review, and importantly, 14 planets that transit their star at each orbital revolution have been identified (see fig 4). These transiting planets are especially interesting because of the possibility to measure both their mass and size and thus obtain constraints on their global composition. In spite of their particular location just a few stellar radii away from their stars, the transiting planets that have been discovered bear some resemblance with their Solar System cousins in the sense that they are also mostly made of hydrogen and helium (e.g. Burrows et al., 2000; Guillot, 2005; Baraffe et al., 2005). They are, however, much hotter due to the intense irradiation that they receive. Although obtaining direct informations on these planets represent a great observational chal- lenge, several key steps have been accomplished: Atomic Sodium, predicted to be detectable (Seager and Sasselov, 2000), has indeed been detected by transit spectroscopy around one planet (Charbonneau et al., 2002). Hydrodynamically escaping hydrogen, oxygen and carbon have also been detected from the same planet (Vidal-Madjar et al., 2003, 2004). The measurement of the sec- ondary eclipse of several planets by the Spitzer satellite allowed a constraint on spectral and hence thermal properties of the planetary atmospheres (Charbonneau et al., 2005; Deming et al., 2005). Recently, the light curve of a non-transiting planet was detected in the infrared, also with Spitzer, providing preliminary indications of a strong day/night temperature variation (Harrington et al., 2006) perhaps even larger than predicted (see Showman and Guillot, 2002). Obviously, there is a big potential for growth in this young field, and the comparison between fine observations made for giant planets in our Solar System and the more crude, but also more statistically significant data obtained for planets around other stars promises to be extremely fruitful to better understand these objects. 3 The calculation of interior and evolution models 3.1 Basic equations The structure and evolution of a giant planet is governed by the following hydrostatic, thermody- namic, mass conservation and energy conservation equations: ∂P ∂r ∂T ∂r ∂m ∂r ∂L ∂r = −ρg = ∇T . ∂P ∂r T P = 4πr2ρ. = 4πr2ρ(cid:18) ǫ − T ∂S ∂t(cid:19) , (4) (5) (6) (7) where P is the pressure, ρ the density, and g = Gm/r2 the gravity (m is the mass, r the radius and G the gravitational constant). The temperature gradient ∇T ≡ (d ln T /d ln P ) depends on the process by which the internal heat is transported. L is the intrinsic luminosity, t the time, S the specific entropy (per unit mass), and ǫ accounts for the sources of energy due e.g. to radioactivity or more importantly nuclear reactions. Generally it is a good approximation to assume ǫ ∼ 0 for 14 objects less massive than ∼ 13 MJ, i.e. too cold to even burn deuterium (but we will see that in certain conditions this term may be useful, even for low mass planets). The boundary condition at the center is trivial: r = 0; (m = 0, L = 0). The external boundary condition is more difficult to obtain because it depends on how energy is transported in the atmo- sphere. One possibility is to use the Eddington approximation, and to write (e.g. Chandrasekhar, 1939): r = R; (T0 = Teff, P0 = 2/3 g/κ), where Teff is the effective temperature (defined by L = 4πRσT 4 eff, with σ being the Stephan-Boltzmann constant), and κ is the opacity. Note for example that in the case of Jupiter Teff = 124 K, g = 26 m s−2 and κ ≈ 5 × 10−3(P/1 bar) m2 kg−1. This implies P0 ≈ 0.2 bar (20,000 Pa), which is actually close to Jupiter’s tropopause, where T ≈ 110 K. More generally, one has to use an atmospheric model relating the temperature and pressure at a given level to the radius R, intrinsic luminosity L and incoming stellar luminosity L∗p: r = R; (T0 = T0(R, L, L∗p), P0 = P0(R, L, L∗p)). P0 is chosen to satisfy the condition that the corresponding optical depth at that level should be much larger than unity. If the stellar flux is absorbed mostly in a convective zone, then the problem can be simplified by using T0(R, L, L∗p) ≈ T0(R, L + L∗p, 0) (e.g. Hubbard, 1977). An example of such a model is described by Saumon et al. (1996) and Hubbard et al. (2002) and is used hereafter to model the planets in the low irradiation limit. 3.2 High pressure physics & equations of state In terms of pressures and temperatures, the interiors of giant planets lie in a region for which accu- rate equations of state (EOS) are extremely difficult to calculate. This is because both molecules, atoms, and ions can all coexist, in a fluid that is partially degenerate (free electrons have energies that are determined both by quantum and thermal effects) and partially coupled (coulomb inter- actions between ions are not dominant but must be taken into account). The presence of many elements and their possible interactions further complicate matters. For lack of space, this section will mostly focus on hydrogen whose EOS has seen the most important developments in recent years. A phase diagram of hydrogen (fig. 5) illustrates some of the important phenomena that occur in giant planets. The photospheres of giant planets are generally relatively cold (50 to 3000 K) and at low pressure (0.1 to 10 bar, or 104 to 106 Pa), so that hydrogen is in molecular form and the perfect gas conditions apply. As one goes deeper into the interior hydrogen and helium progressively become fluid. (The perfect gas relation tends to underestimate the pressure by 10% or more when the density becomes larger than about 0.02 g cm−3 (P ∼> 1 kbar in the case of Jupiter)). Characteristic interior pressures are considerably larger however: as implied by Eqs. 4 and 6, Pc ≈ GM 2/R4, of the order of 10-100 Mbar for Jupiter and Saturn. At these pressures and the corresponding densities, the Fermi temperature TF is larger than 105 K. This implies that electrons are degenerate. Figure 5 shows that inside Jupiter, Saturn, the extrasolar planet HD 209458 b, but also for giant planets in general for most of their history, the degeneracy parameter θ = T /TF is between 0.1 and 0.03. Therefore, the energy of electrons in the interior is expected to be only slightly larger than their non-relativistic, fully degenerate limit: ue ≥ 3/5 kTF = 15.6 (ρ/µe)2/3 eV, where k is Boltzmann’s constant, µe is the number of electrons per nucleon and ρ is the density in g cm−3. For pure hydrogen, when the density reaches ∼ 0.8 g cm−3, the average energy of electrons becomes larger than hydrogen’s ionization potential, even at zero temperature: hydrogen pressure- ionizes and becomes metallic. This molecular to metallic transition occurs near Mbar pressures, but exactly how this happens remains unclear because of the complex interplay of thermal, coulomb, and degeneracy effects (in particular, whether hydrogen metallizes into an atomic state H+ — as 15 Figure 5: Phase diagram for hydrogen with the main phase transitions occurring in the fluid or gas phase. The temperature-pressure profiles for Jupiter, Saturn, Uranus, Neptune, and the exoplanet HD 209458 b are shown. The dashed nearly vertical line near 1 Mbar is indicative of the molecular to metallic transition (here it represents the so-called plasma phase transition as calculated by Saumon et al. (1995)). The region in which hydrogen is in solid phase (Datchi et al., 2000; Gregoryanz et al., 2003) is represented as a hatched area. The three phases (I,II,III) of solid hydrogen are shown (see Mao and Hemley, 1994). Values of the degeneracy parameter θ are indicated as dotted lines to the upper right corner of the figure. 16 suggested in Fig. 5 — or first metallizes in the molecular state H2 remains to be clarified). Recent laboratory measurements on fluid deuterium have been able to reach pressures above ∼> 1 Mbar, and provide new data in a region where the EOS remains most uncertain. Gas- guns experiments have been able to measure the reshock temperature (Holmes et al., 1995), near T ∼ 5000 K, P ∼ 0.8 Mbar, and a rise in the conductivity of molecular hydrogen up to T ∼ 3000 K, P ∼ 1.4 Mbar, a sign that metallicity may have been reached (Weir et al., 1996). The following few years have seen the development of laser-induced shock compression (da Silva et al., 1997; Collins et al., 1998), pulsed-power shock compression (Knudson et al., 2004), and convergent shock wave experiments (Belov et al., 2002; Boriskov et al., 2005) in a high-pressure (P = 0.3 − 4 Mbar) high-temperature (T ∼ 6000 − 105 K) regime. Unfortunately, experimental results along the prin- cipal Hugoniot of deuterium do not agree in this pressure range. Laser compression data give a maximum compression of ∼ 6 while both the pulsed-power compression experiments and the convergent shock wave experiments find a value of ∼ 4 . Models that are partly calibrated with experimental data (Saumon et al., 1995; Ross, 1998; Ross and Yang, 2001) obtain a generally good agreement with the laser-compression data. However, the fact that independant models based on first principles (Militzer et al., 2001; Desjarlais, 2003; Bonev et al., 2004) yield low compressions strongly favors this solution. The question of the existence of a first-order molecular to metallic transition of hydrogen (i.e. both molecular dissociation and ionisation occur simultaneously and discontinuously at the so-called plasma phase transition, or PPT) remains however. The critical line shown in fig. 5 corresponds to calculations by Saumon et al. (1995), but may be caused by artefacts in the free energy calculation. Recent Density Functional Theory (DFT) simulations by Bonev et al. (2004) indicate the possibility of a first order liquid-liquid transition but other path-integral calculations (Militzer et al., 2001) do not. It is crucial to assess the existence of such a PPT because it would affect both convection and chemical composition in the giant planets. A clear result from fig. 5 at least is that, as first shown by Hubbard (1968), the interiors of the hydrogen-helium giant planets are fluid, whatever their age: of course, they avoid the critical point for the liquid gas transition in hydrogen and helium, at very low temperatures, but they also lie comfortably above the solidification lines for hydrogen and helium. (An isolated Jupiter should begin partial solidification only after at least ∼ 103 Ga of evolution.) They are considered to be fluid because at the high pressures and relatively modest temperatures in their interiors, coulomb interactions between ions play an important role in the EOS and yield a behavior that is more reminiscent of that of a liquid than that of a gas, contrary to what is the case in e.g. solar-like stars. For Uranus and Neptune, the situation is actually more complex because at large pressures they are not expected to contain hydrogen, but numerical simulations show that ices in their interior should be fluid as well (Cavazzoni et al., 1999). Models of the interiors of giant planets require thermodynamically consistent EOSs calculated over the entire domain of pressure and temperature spanned by the planets during their evolution. Elements other than hydrogen, most importantly helium, should be consistently included. Such a calculation is a daunting task, and the only recent attempt at such an astrophysical EOS for substellar objects is that by Saumon et al. (1995). Another set of EOSs reproducing either the high- or low-compression results was calculated by Saumon and Guillot (2004) specifically for the calculation of present-day models of Jupiter and Saturn. These EOSs have so far included other elements (including helium), only in a very approximative way, i.e. with EOSs for helium and heavy elements that are based on interpolations between somewhat ideal regimes, using an additive volume law, and neglecting the possibility of existence of phase separations (see Hubbard et al., 2002; Guillot et al., 2004, for further discussions). 17 3.3 Heat transport Giant planets possess hot interiors, implying that a relatively large amount of energy has to be transported from the deep regions of the planets to their surface. This can either be done by radi- ation, conduction, or, if these processes are not sufficient, by convection. Convection is generally ensured by the rapid rise of the opacity with increasing pressure and temperature. At pressures of a bar or more and relatively low temperatures (less than 1000 K), the three dominant sources of opacities are water, methane and collision-induced absorption by hydrogen molecules. However, in the intermediate temperature range between ∼ 1200 and 1500 K, the Rosseland opacity due to the hydrogen and helium absorption behaves differently: the absorption at any given wavelength increases with density, but because the temperature also rises, the photons are emitted at shorter wavelengths, where the monochromatic absorption is smaller. As a consequence, the opacity can decrease. This was shown by Guillot et al. (1994) to potentially lead to the presence of a deep radiative zone in the interiors of Jupiter, Saturn and Uranus. This problem must however be reanalyzed in the light of recent observations and analyses of brown dwarfs. Their spectra show unexpectedly wide sodium and potassium absorption lines (see Burrows, Marley & Sharp 2000), in spectral regions where hydrogen, helium, water, methane and ammonia are relatively transparent. It thus appears that the added contribution of these elements (if they are indeed present) would wipe out any radiative region at these levels (Guillot et al., 2004). At temperatures above 1500 ∼ 2000 K two important sources of opacity appear: (i) the rising number of electrons greatly enhances the absorption of H− 2 and H−; (ii) TiO, a very strong absorber at visible wavelengths is freed by the vaporization of CaTiO3. Again, the opacity rises rapidly which ensures a convective transport of the heat. Still deeper, conduction by free electrons becomes more efficient, but the densities are found not to be high enough for this process to be significant, except perhaps near the central core (see Hubbard, 1968; Stevenson and Salpeter, 1977b). While our giant planets seem to possess globally convective interiors, strongly irradiated ex- trasolar planets must develop a radiative zone just beneath the levels where most of the stellar irradiation is absorbed. Depending on the irradiation and characteristics of the planet, this zone may extend down to kbar levels, the deeper levels being convective. In this case, a careful deter- mination of the opacities is necessary (but generally not possible) as these control the cooling and contraction of the deeper interior (see Ferguson et al., 2005, for a discussion of opacities and tables for substellar atmospheres and interiors). 3.4 The contraction and cooling histories of giant planets The interiors of giant planets is expected to evolve with time from a high entropy, high θ value, hot initial state to a low entropy, low θ, cold degenerate state. The essential physics behind can be derived from the well-known virial theorem and the energy conservation which link the planet’s internal energy Ei, gravitational energy Eg and luminosity through: ξEi + Eg = 0, L = − ξ − 1 ξ dEg dt , (8) (9) where ξ = R M 0 udm ≈< 3P/ρu >, the brackets indicating averaging, and u is the specific internal energy. For a diatomic perfect gas, ξ = 3.2; for fully-degenerate non-relativistic electrons, ξ = 2. 0 3(P/ρ)dm/R M 18 Thus, for a giant planet or brown dwarf beginning its life mostly as a perfect H2 gas, two third of the energy gained by contraction is radiated away, one third being used to increase Ei. The internal energy being proportional to the temperature, the effect is to heat up the planet. This represents the slightly counter-intuitive but well known effect that a star or giant planet initially heats up while radiating a significant luminosity (e.g. Kippenhahn and Weigert, 1994). Let us now move further in the evolution, when the contraction has proceeded to a point where the electrons have become degenerate. For simplicity, we will ignore coulomb interactions and exchange terms, and assume that the internal energy can be written as Ei = Eel + Eion, and that furthermore Eel ≫ Eion (θ is small). Because ξ ≈ 2, we know that half of the gravitational potential energy is radiated away and half of it goes into internal energy. The problem is to decide how this energy is split into an electronic and an ionic part. The gravitational energy changes with some average value of the interior density as Eg ∝ 1/R ∝ ρ1/3. The energy of the degenerate Eel ≈ 2(Eel/Eg) Eg. Using the electrons is essentially the Fermi energy: Eel ∝ ρ2/3. Therefore, virial theorem, this yields: Eel ≈ − Eg ≈ 2L L ≈ − Eion ∝ − T . (10) (11) The gravitational energy lost is entirely absorbed by the degenerate electrons, and the observed luminosity is due to the thermal cooling of the ions. Several simplifications limit the applicability of this result (that would be valid in the white dwarf regime). In particular, the coulomb and exchange terms in the EOS introduce negative contributions that cannot be neglected. However, the approach is useful to grasp how the evolution proceeds: in its very early stages, the planet is very compressible. It follows a standard Kelvin- Helmoltz contraction. When degeneracy sets in, the compressibility becomes much smaller (αT ∼ 0.1, where α is the coefficient of thermal expansion), and the planet gets its luminosity mostly from the thermal cooling of the ions. The luminosity can be written in terms of a modified Kelvin- Helmoltz formula: L ≈ η , (12) GM 2 Rτ where τ is the age, and η is a factor that hides most of the complex physics. In the approximation that coulomb and exchange terms can be neglected, η ≈ θ/(θ+1). The poor compressibility of giant planets in their mature evolution stages imply that η ≪ 1 (η ∼ 0.03 for Jupiter): the luminosity is not obtained from the entire gravitational potential, but from the much more limited reservoir constituted by the thermal internal energy. Equation 12 shows that to first order, log L ∝ − log τ : very little time is spent at high luminosity values. In other words, the problem is (in most cases) weakly sensitive to initial conditions. However, it is to be noticed that with progresses in our capabilities to detect very young objects, i.e. planets and brown dwarfs of only a few million years of age, the problem of the initial conditions does become important (Marley et al., 2006). Figure 6 shows more generally how giant planets, but also brown dwarfs and small stars see their luminosities evolve as a function of time. The 1/τ slope is globally conserved, with some variations for brown dwarfs during the transient epoch of deuterium burning, and of course for stars, when they begin burning efficiently their hydrogen and settle on the main sequence: in that case, the tendency of the star to contract under the action of gravity is exactly balanced by thermonuclear hydrogen fusion. 19 Figure 6: Evolution of the luminosity (in L⊙) of solar-metallicity M dwarfs and substellar objects vs. time (in yr) after formation. In this figure, ”brown dwarfs” are arbitrarily designated as those objects that burn deuterium, while those that do not are tentatively labelled ”planets”. Stars are objects massive enough to halt their contraction due to hydrogen fusion. Each curve is labelled by its corresponding mass in M⊙, with the lowest three corresponding to the mass of Saturn, half the mass of Jupiter, and the mass of Jupiter. [From Burrows et al. (1997)]. 20 n o c o r e n o core 100 M⊕ core T eq=2000K Teq=1000K H - H e , n o c o r e e o l a c o r e d t i s 5 M 1 ⊕ core ⊕ 100 M planets brown dwarfs stars Figure 7: Theoretical and observed mass-radius relations. The black line is applicable to the evolution of solar composition planets, brown dwarfs and stars, when isolated or nearly isolated (as Jupiter, Saturn, Uranus and Neptune, defined by diamonds and their respective symbols), after 5 Ga of evolution. The dotted line shows the effect of a 15 M⊕ core on the mass-radius relation. Orange and yellow curves represent the mass-radius relations for heavily irradiated planets with equilibrium temperatures of 1000 and 2000 K, respectively, and assuming that 0.5% of the incoming stellar luminosity is dissipated at the center (see section 4.3). For each irradiation level, two cases are considered: a solar-composition planet with no core (top curve), and one with a 100 M⊕ central core (bottom curve). The transiting extrasolar giant planets for which a mass and a radius was measured are shown with points that are color-coded in function of the planet’s equilibrium temperature. The masses and radii of very low mass stars are also indicated as blue points with error bars. 21 3.5 Mass-radius relation The relation between mass and radius has very fundamental astrophysical applications. Most importantly it allows one to infer the gross composition of an object from a measurement of its mass and radius. This is especially relevant in the context of the discovery of extrasolar planets with both radial velocimetry and the transit method, as the two techniques yield relatively accurate determination of M and R. Figure 7 shows mass-radius relations for compact degenerate objects from giant planets to brown dwarfs and low-mass stars. The right-hand side of the diagram shows a rapid increase of the radius with mass in the stellar regime which is directly due to the onset of stable thermonuclear reactions. In this regime, observations and theoretical models agree (see however Ribas, 2006, for a more detailed discussion). The left-hand side of the diagram is obviously more complex, and this can be understood by the fact that planets have much larger variations in compositions than stars, and because external factors such as the amount of irradiation they receive do affect their contraction in a significant manner. Let us first concentrate on isolated or nearly-isolated gaseous planets. The black curves have a local maximum near 4 MJ: at small masses, the compression is small so that the radius increases with mass. At large masses, degeneracy sets in and the radius decreases with mass. This can be understood on the basis of polytropic models based on the assumption that P = Kρ1+1/n, where K and n are constants. Because of degeneracy, a planet of large mass will tend to have n → 1.5, while a planet a smaller mass will be less compressible (n → 0). Indeed, it can be shown that in their inner 70 to 80% in radius isolated solar composition planets of 10, 1 and 0.1 MJ have n = 1.3, 1.0 and 0.6, respectively. From polytropic equations (e.g. Chandrasekhar, 1939): R ∝ K n 3−n M 1−n 3−n . (13) Assuming that K is independant of mass, one gets R ∝ M 0.16, M 0, and M −0.18 for M = 10, 1 and 0.1 MJ, respectively, in relatively good agreement with fig. 7 (the small discrepancies are due to the fact that the intrinsic luminosity and hence K depend on the mass considered). Figure 7 shows already that the planets in our Solar System are not made of pure hydrogen and helium and require an additional fraction of heavy elements in their interior, either in the form of a core, or distributed in the envelope (dotted line). For extrasolar planets, the situation is complicated by the fact that the intense irradiation that they receive plays a major role in their evolution. The present sample is already quite diverse, with equilibrium temperature (defined as the effective temperature corresponding to the stellar flux received by the planet) ranging from 1000 to 2500 K. Their composition is also quite variable, with some planets having large masses of heavy elements (Sato et al., 2005; Guillot et al., 2006). The orange and yellow curves in fig. 7 show theoretical results for equilibrium temperatures of 1000 and 2000 K, respectively. Two extreme models have been plotted: assuming a purely solar composition planet (top curve), and assuming the presence of a 100 M⊕ central core (bottom curve). In each case, an additional energy source proportional to 0.5% of the incoming luminosity was also assumed (see discussion in § 4.3 hereafter). The increase in radius for decreasing planetary mass for irradiated, solar-composition planets with little or no core can be understood using the polytropic relation (eq. 13), but accounting for variations of K as defined by the atmospheric boundary condition. Using the Eddington approximation, assuming κ ∝ P and a perfect gas relation in the atmosphere, one can show 2−n . With n = 1, one finds R ∝ M −1/2. that K ∝ (M/R2)−1/2n and that therefore R ∝ M Strongly irradiated hydrogen-helium planets of small masses are hence expected to have the largest 1/2−n 22 radii which qualitatively explain the positions of the extrasolar planets in fig. 7. Note that this estimate implicitly assumes that n is constant throughout the planet. The real situation is more complex because of the growth of a deep radiative region in most irradiated planets, and because of structural changes between the degenerate interior and the perfect gas atmosphere. In the case of the presence of a fixed mass of heavy elements, the trend is inverse because of the increase of mean molecular mass (or equivalently core/envelope mass) with decreasing total mass. Thus, small planets with a core are much more tightly bound and less subject to evaporation than those that have no core. 3.6 Rotation and the figures of planets The mass and radius of a planet informs us on its global composition. Because planets are also rotating, one is allowed to obtain more information on their deep interior structure. The hydrostatic equation becomes more complex however: ∇P ρ = ∇(cid:18)GZZZ ρ(r′) r − r′ d3r′(cid:19) − Ω × (Ω × r), (14) where Ω is the rotation vector. The resolution of eq. (14) is a complex problem. It can however be somewhat simplified by assuming that Ω ≡ ω is such that the centrifugal force can be derived from a potential. The hydrostatic equilibrium then writes ∇P = ρ∇U , and the figure of the rotating planet is then defined by the U = constant level surface. One can show (e.g. Zharkov and Trubitsyn, 1978) that the hydrostatic equation of a fluid planet can then be written in terms of the mean radius r (the radius of a sphere containing the same volume as that enclosed by the considered equipotential surface): 1 ρ ∂P ∂r = − Gm r2 + 2 3 ω2r + GM R 3 rϕω, (15) where M and R are the total mass and mean radius of the planet, and ϕω is a slowly varying function of r. (In the case of Jupiter, ϕω varies from about 2 × 10−3 at the center to 4 × 10−3 at the surface.) Equations (5-7) remain the same with the hypothesis that the level surfaces for the pressure, temperature, and luminosity are equipotentials. The significance of rotation is measured by the ratio of the centrifugal acceleration to the gravity: q = ω2R3 eq GM . (16) As discussed in section 2.2, in some cases, the external gravity field of a planet can be accu- rately measured in the form of gravitational moments Jk (with zero odd moments for a planet in hydrostatic equilibrium) that measure the departure from spherical symmetry. Together with the mass, this provides a constraint on the interior density profile (see Zharkov and Trubitsyn (1974) -see also chapters by Van Hoolst and Sohl & Schubert): M = ZZZ ρ(r, θ)d3τ, 1 eq ZZZ ρ(r, θ)r2iP2i(cos θ)d3τ, J2i = − M R2i where dτ is a volume element and the integrals are performed over the entire volume of the planet. 23 Figure 8 shows how the different layers inside a planet contribute to the mass and the grav- itational moments. The figure applies to Jupiter, but would remain relatively similar for other planets. Note however that in the case of Uranus and Neptune, the core is a sizable fraction of the total planet and contributes both to J2 and J4. Measured gravitational moments thus provide information on the external levels of a planet. It is only indirectly, through the constraints on the outer envelope that the presence of a central core can be infered. As a consequence, it is impossible to determine this core’s state (liquid or solid), structure (differentiated, partially mixed with the envelope) and composition (rock, ice, helium...). Figure 8: Contribution of the level radii to the gravitational moments of Jupiter. J0 is equivalent to the planet’s mass. The small discontinuities are caused by the following transitions, from left to right: core/envelope, helium rich/helium poor (metallic/molecular). Diamonds indicate the median radius for each moment. For planets outside the solar system, although measuring their gravitational potential is utopic, their oblateness may be reachable with future space transit observations (Seager and Hui, 2002). Since the oblateness e is, to first order, proportionnal to q: e = Req Req − Rpol ≈(cid:18) 3 2 Λ2 + 1 2(cid:19) q (17) (where Λ2 = J2/q ≈ 0.1 to 0.2), it may be possible to obtain their rotation rate, or with a rotation measured from another method, a first constraint on their interior structure. 4 Interior structures and evolutions 4.1 Jupiter and Saturn As illustrated by fig. 9, the simplest interior models of Jupiter and Saturn matching all observational constraints assume the presence of three main layers: (i) an outer hydrogen-helium envelope, whose 24 Figure 9: Schematic representation of the interiors of Jupiter and Saturn. The range of tempera- tures is estimated using homogeneous models and including a possible radiative zone indicated by the hashed regions. Helium mass mixing ratios Y are indicated. The size of the central rock and ice cores of Jupiter and Saturn is very uncertain (see text). In the case of Saturn, the inhomogeneous region may extend down all the way to the core which would imply the formation of a helium core. [Adapted from Guillot (1999b)]. 25 global composition is that of the deep atmosphere; (ii) an inner hydrogen-helium envelope, enriched in helium because the whole planet has to fit the H/He protosolar value; (iii) a central dense core. Because the planets are believed to be mostly convective, these regions are expected to be globally homogeneous. (Many interesting thermochemical transformations take place in the deep atmosphere, but they are of little concern to us). The transition from a helium-poor upper envelope to a helium-rich lower envelope is thought to take place through the formation of helium-rich droplets that fall deeper into the planet due to their larger density. These droplets form because of an assumed phase transition of helium in hydrogen at high pressures and low temperatures. Three-layer models implicitely make the hypothesis that this region is narrow. Indeed, calculations of such a phase separation in a fully- ionized plasma indicate a rapid decrease of the critical temperature with increasing pressure, with the consequence that helium would be unsoluble in a relatively small region in low-pressure metallic hydrogen. This region would progressively grow with time (e.g. Stevenson, 1982). However, DFT calculations have indicated that the critical temperature for helium demixing may rise with pressure (Pfaffenzeller et al., 1995), presumably in the regime where hydrogen is only partially ionized and bound states remain. This opens up the possibility that the inhomogeneous regions may be more extended, and that models more complex than the three-layer models may be needed, in particular in the case of Saturn (see below). In the absence of these calculations, the three-layer models can be used as a useful guid- ance to a necessarily hypothetical ensemble of allowed structures and compositions of Jupiter and Saturn. These relatively extensive exploration of the parameter space have been performed by Saumon and Guillot (2004). The calculations assume that only helium is inhomogeneous in the en- velope (the abundance of heavy elements is supposed to be uniform accross the molecular/metallic hydrogen transition). Many sources of uncertainties are taken into account however; among them, the most significant are on the equations of state of hydrogen and helium, the uncertain values of J4 and J6, the presence of differential rotation deep inside the planet, the location of the helium-poor to helium-rich region, and the uncertain helium to hydrogen protosolar ratio. Their results indicate that Jupiter’s core is smaller than ∼ 10 M⊕, and that its global com- position is pretty much unknown (between 10 to 42 M⊕ of heavy elements in total). The models indicate that Jupiter is enriched compared to the solar value by a factor 1.5 to 8 times the solar value. This enrichment is compatible with a global uniform enrichment of all species near the atmospheric Galileo values, but include many other possibilities. In the case of Saturn, the solutions depend less on the hydrogen EOS because the Mbar pressure region is comparatively smaller. The total amount of heavy elements present in the planet can therefore be estimated with a better accuracy than for Jupiter, and is between 20 and 30 M⊕. In three-layer models with a discontinuity of the helium abundance at the molecular-metallic hydrogen interface but continuity of all other elements, the core masses found are between 10 and 22 M⊕. However, because Saturn’s metallic region is deeper into the planet, it mimics the effect that a central core would have on J2. If we allow for variations in the abundance of heavy elements together with the helium discontinuity, then the core mass can become much smaller, and even solutions with no core can be found (Guillot 1999a). These solutions depend on the hypothetic phase separation of an abundant species (e.g. water), and generally cause an energy problem because of the release of considerable gravitational energy. However, another possibility is through the formation of an almost pure helium shell around the central core, which could lower the core masses by up to 7 M⊕ (Fortney and Hubbard, 2003, Hubbard, personnal communication). Concerning the evolutions of Jupiter and Saturn, the three main sources of uncertainty are, by order of importance: (1) the magnitude of the helium separation; (2) the EOS; (3) the at- 26 mospheric boundary conditions. Figure 10 shows an ensemble of possibilities that attempt to bracket the minimum and maximum cooling. In all cases, helium sedimentation is needed to ex- plain Saturn’s present luminosity (see Salpeter, 1973; Stevenson and Salpeter, 1977a; Hubbard, 1977). Recent models of Saturn’s evolution appear to favor a scenario in which helium settles down almost to the central core (Hubbard et al., 1999; Fortney and Hubbard, 2003). In the case of Jupiter, the sedimentation of helium that appears to be necessary to explain the low atmospheric helium abundance poses a problem for evolution models because it appears to generally prolong its evolution beyond 4.55 Ga, the age of the Solar System. However, different solutions are possible, including improvements of the EOS and atmospheric boundary conditions, or even the possible progressive erosion of the central core that would yield a lower Jupiter’s luminosity at a given age (Guillot et al., 2004). Figure 10: Final stages of evolution of Jupiter and Saturn. The present effective temperatures, reached after ∼ 4.55 Ga of evolution, are indicated as horizontal orange lines. For each planet two models represent attempts to bracket the ensemble of possibilities, with the faster evolu- tion corresponding to that of an homogeneous planet, while the slowest evolution includes the effect of helium settling in the last evolution phase. [Adapted from Hubbard et al. (1999) and Fortney and Hubbard (2003)]. 4.2 Uranus and Neptune Although the two planets are relatively similar, fig. 7 already shows that Neptune’s larger mean density compared to Uranus has to be due to a slightly different composition: either more heavy elements compared to hydrogen and helium, or a larger rock/ice ratio. The gravitational moments impose that the density profiles lie close to that of “ices” (a mixture initially composed of e.g. H2O, CH4 and NH3, but which rapidly becomes a ionic fluid of uncertain chemical composition in the planetary interior), except in the outermost layers, which have a density closer to that of hy- drogen and helium (Marley et al., 1995; Podolak et al., 2000). As illustrated in fig. 11, three-layer models of Uranus and Neptune consisting of a central “rocks” core (magnesium-silicate and iron material), an ice layer and a hydrogen-helium gas envelope have been calculated (Podolak et al., 1991; Hubbard et al., 1995). 27 Figure 11: Schematic representation of the interiors of Uranus and Neptune. [Adapted from Guillot (1999b)]. The fact that models of Uranus assuming homogeneity of each layer and adiabatic temperature profiles fail in reproducing its gravitational moments seem to imply that substantial parts of the planetary interior are not homogeneously mixed (Podolak et al., 1995). This could explain the fact that Uranus’ heat flux is so small: its heat would not be allowed to escape to space by convection, but through a much slower diffusive process in the regions of high molecular weight gradient. Such regions would also be present in Neptune, but much deeper, thus allowing more heat to be transported outward. The existence of these non-homogeneous, partially mixed regions are further confirmed by the fact that if hydrogen is supposed to be confined solely to the hydrogen-helium envelope, models predict ice/rock ratios of the order of 10 or more, much larger than the protosolar value of ∼ 2.5. On the other hand, if we impose the constraint that the ice/rock ratio is protosolar, the overall composition of both Uranus and Neptune is, by mass, about 25% rock, 60 − 70% ice, and 5 − 15% hydrogen and helium (Podolak et al., 1991, 1995; Hubbard et al., 1995). Assuming both ice and rock are present in the envelope, an upper limit to the amount of hydrogen and helium present is ∼ 4.2 M⊕ for Uranus and ∼ 3.2 M⊕ for Neptune (Podolak et al., 2000). A lower limit of ∼ 0.5 M⊕ for both planets can be inferred by assuming that hydrogen and helium are only present in the outer envelope at P ∼< 100 kbar. 4.3 Irradiated giant planets Although all extrasolar giant planets are in principle interesting, we focus here on the ones that orbit extremely close to their star because of the possibility to directly characterise them and measure their mass, radius and some properties of their atmosphere. Two planets are proxies for this new class of objects: the first extrasolar giant planet discovered, 51 Peg b, with an orbital period of P = 4.23 days, and the first transiting extrasolar giant planet, HD 209458 b, with P = 3.52 days. Both planets belong to the Pegasus constellation, and following astronomical conventions (e.g. Cepheids, named after δ Cephei), we choose to name giant planets orbiting close to their stars with periods shorter than 10 days “Pegasids” (alternatively, “hot Jupiters” is also found in the litterature). With such a short orbital period, these planets are for most of them subject to an irradiation 28 from their central star that is so intense that the absorbed stellar energy flux can be about ∼ 104 times larger than their intrinsic flux. The atmosphere is thus prevented from cooling, with the consequence that a radiative zone develops and governs the cooling and contraction of the interior (Guillot et al., 1996). Typically, for a planet like HD 209458 b, this radiative zone extends to kbar levels, T ∼ 4000 K, and is located in the outer 5% in radius (0.3% in mass) (Guillot and Showman, 2002). Figure 12: Conjectured dynamical structure of Pegasids (strongly irradiated extrasolar giant plan- ets): At pressures larger than 100–800 bar, the intrinsic heat flux must be transported by convec- tion. The convective core is at or near synchronous rotation with the star and has small latitudinal and longitudinal temperature variations. At lower pressures a radiative envelope is present. The top part of the atmosphere is penetrated by the stellar light on the day side. The spatial variation in insolation should drive winds that transport heat from the day side to the night side. [From Showman and Guillot (2002)]. Problems in the modeling of the evolution of Pegasids arise mostly because of the uncertain outer boundary condition. The intense stellar flux implies that the atmospheric temperature profile is extremely dependant upon the opacity sources considered. Depending on the chosen composi- tion, the opacity data used, the assumed presence of clouds, the geometry considered, resulting temperatures in the deep atmosphere can differ by up to ∼ 600 K (Seager and Sasselov, 2000; Goukenleuque et al., 2000; Barman et al., 2001; Sudarsky et al., 2003; Iro et al., 2005; Fortney et al., 2006). Furthermore, as illustrated by fig. 12, the strong irradiation and expected synchronisa- tion of the planets implies that strong inhomogeneities should exist in the atmosphere with in particular strong (∼ 500 K) day-night and equator-to-pole differences in effective temperatures (Showman and Guillot, 2002; Iro et al., 2005; Cooper and Showman, 2005; Barman et al., 2005), further complicating the modeling of the planetary evolution (see fig. fig. 13). Finally, another related problem is the presence of the radiative zone. Again, the composition is unknown and the opacity data are uncertain in this relatively high temperature (T ∼ 1500 − 3000 K) and high 29 pressure (up to ∼ 1 kbar) regime. We have seen in fig. fig. 7 that the measured masses and radii of transiting planets can be globally explained in the framework of an evolution model including the strong stellar irradiation and the presence of a variable mass of heavy elements, either in the form of a central core, or spread in the planet interior. However, when analyzing the situation for each planet, it appears that several planets are too large to be reproduced by standard models, i.e. models using the most up-to-date equations of state, opacities, atmospheric boundary conditions and assuming that the planetary luminosity governing its cooling is taken solely from the lost gravitational potential energy (see Section 3.1). Figure 14 illustrates the situation for the particular case of HD209458b: unless using an unre- alistically hot atmosphere, or arbitrarily increasing the internal opacity, or decreasing the helium content, one cannot reproduce the observed radius which is 10 to 20% larger than calculated (Bodenheimer et al., 2001, 2003; Guillot and Showman, 2002; Baraffe et al., 2003). The fact that the measured radius corresponds to a low-pressure (∼mbar) level while the calculated radius cor- responds to a level near 1 bar is not negligible (Burrows et al., 2003) but too small to account for the difference. This is problematic because while it is easy to invoke the presence of a massive core to explain the small size of a planet, a large size such as that of HD209458b requires an additional energy source, or significant modifications in the data/physics involved. Bodenheimer et al. (2001) proposed that this large radius may be due to a small forced ec- centricity (e ∼ 0.03) of HD209458b, and subsequent tidal dissipation in the planet interior, but detailed observations indicate that the eccentricity is small, e = 0.014 ± 0.009 (Laughlin et al., 2005), and observations of the secondary eclipse imply that this would further require a chance configuration of the orbit (Deming et al., 2005). Another proposed explanation also involving tidal dissipation of orbital energy is that the planet may be trapped in a Cassini state with a large or- bital inclination (Winn and Holman, 2005), but it appears to have a low probability of occurrence (Levrard et al., 2006). Finally, a third possibility that would apply to all Pegasids is to invoke a downward transport of kinetic energy and its dissipation by tides (Showman and Guillot, 2002). This last possibility would require the various transiting planets to have different core masses to reproduce the observed radii (Guillot, 2005). Recently, as more transiting Pegasids have been discovered, the number of anomalously large ones has increased to at least 3 for 11 planets, implying that this is not a rare event. This lends more weight to a mechanism that would apply to each planet. In this case, masses of heavy elements can be derived by imposing that all planets should be fitted by the same model with the same hypotheses. This can be done by inverting the results of fig. 7, as described by (Guillot et al., 2006). The method is applied to the known transiting Pegasids by the end of 2006 in fig. 15, a plot of the masses of heavy elements in the planets as a function of the metallicities of the parent star (which measures how rich a given star is in heavy elements compared to the Sun). Figure 15 first shows that in some cases, large masses of heavy elements (up to ∼ 100 M⊕ are necessary. This is in harmony with the composition inferred for HD149026b, i.e. around 70 M⊕ of heavy elements, a conclusion that is hard to escape because of the low total mass and high irradiation of the planet (see Ikoma et al., 2006; Fortney et al., 2006). Furthermore, there seems to be a correlation between the mass of heavy elements inferred in the transiting planets, and the metallicity of the parent stars (Guillot et al., 2006), although this correlation has to be ascertained by more measurements. A caveat is important: these results are intrinsically model-dependent as they are based on the assumption that all planets receive an additionnal tidal heat flux that is proportionnal to the energy flux that they receive in the form of photons from the star. More transiting planets are needed to confirm or infirm this model, but the large variety in core masses, 30 Figure 13: Temperature versus pressure for a sequence of locations in the atmosphere of HD209458b, assuming no horizontal redistribution of heat. Each sequence corresponds to a given direction of the incident flux relative to the surface normal. The approximate regions represented by the collection of T-P profiles are shown as solid black lines on the illustrative sphere. The top- most T-P profile corresponds to the sub-stellar point (black dot on the sphere). The terminator and night side (black hemisphere) are modeled with the non-irradiated profile (lowest T-P curve). The radiative-convective boundary at the sub-stellar point and on the night side are labeled with filled circles. The dashed lines indicate the approximate condensation curves for three common grain species. The dotted line indicates where gaseous CO and CH4 concentrations are equal (CO is dominant to the left of this line). The thick, grey, dashed lines are T-P profiles calculated for a normal incident flux equal to 0.5 (top) and 0.25 (bottom) times that at the substellar point, as often used as approximate solutions for the day side, or entire atmosphere, respectively. [From Barman et al. (2005)]. 31 HD 209458b " h o t " a t m o s p h e r e with h e lo w h o p a eliu at dissip atio n m ( Y = cities x 1 0 0.1 5) standard model Figure 14: The contraction of HD209458b as a function of time can be compared to its measured radius and inferred age shown by the black box. Standard models (blue curve) for the evolution of that 0.69 MJ planet generally yield a radius that is too small compared to the observations, even for a solar composition and no central core (a larger core and -in most cases- larger amounts of heavy elements in the planet imply an even smaller size for a given age). Unrealistically low helium abundances or high opacities models lead to evolution tracks that barely cross the observational box. A possiblity is that heat is dissipated into the deep interior by stellar tides, either related to a non-zero orbital eccentricity forced by an unseen companion, or because of a constant transfer of angular momentum from the heated atmosphere to the interior (black curve). Alternatively, the atmosphere may be hotter than predicted due to heating by strong zonal winds and shear instabilities (red curve). 32 Figure 15: Mass of heavy elements in transiting Pegasids known by 2006 as a function of the metal content of the parent star relative to the Sun. The mass of heavy elements required to fit the measured radii is calculated on the basis of evolution models including an additional heat source slowing the cooling of the planet. This heat source is assumed equal to 0.5% of the incoming stellar heat flux (Showman and Guillot, 2002). Horizontal error bars correspond to the 1σ errors on the [Fe/H] determination. Vertical error bars are a consequence of the uncertainties on the measured planetary radii and ages. Note that the results, based on Guillot et al. (2006) are intrinsically model-dependent and may be affected by further discoveries of transiting planets. 33 and the absence of Pegasids around metal-poor stars ([Fe/H] ∼< −0.07) as indicated by fig. 15 appear to be robust consequences of this work. Another intriguing possibility concerning Pegasids is that of a sustained mass loss due to the high irradiation dose that the planets receive. Indeed, this effect was predicted (Burrows and Lunine, 1995; Guillot et al., 1996; Lammer et al., 2003) and detected (Vidal-Madjar et al., 2003, 2004), but its magnitude is still quite uncertain, by at least two orders of magnitude (Lammer et al., 2003; Lecavelier des Etangs et al., 2004; Yelle, 2006). The effect on the evolution is surprisingly lim- ited, except at the final stages when an exponential mass loss appears in fully gaseous planets (Baraffe et al., 2004). Finally, it is important to note that another class of planets awaits a direct characterisation by the transit method: that of ice or rock giants. Small-mass planets around 10 M⊕ have been detected (e.g. Lovis et al., 2006; Beaulieu et al., 2006) but their radius is expected to be small (Guillot et al., 1996; Valencia et al., 2006), and we currently may not have the observational capability to test whether they transit in front of their star. This should be resolved by the space mission CoRoT (launched on 27 dec 2006) and Kepler (launch ∼ 2008). These objects are especially interesting but pose difficult problems in terms of structure because depending on their formation history, precise composition and location, they may be fluid, solid, or they may even possess a global liquid ocean (see Kuchner, 2003; L´eger et al., 2004). 5 Implications for planetary formation models The giant planets in our Solar System have in common possessing a large mass of hydrogen and helium, but they are obviously quite different in their aspect and in their internal structures. Although studies cannot be conducted with the same level of details, we can safely conclude that extrasolar planets show an even greater variety in composition and visible appearance. A parallel study of the structures of our giant planets and of giant planets orbiting around other stars should provide us with key information regarding planet formation in the next decade or so. But, already, some conclusions, some of them robust, others still tentative, can be drawn (see also the chapter by Stevenson): Giant planets formed in circumstellar disks, before these were completely dissipated: This is a relatively obvious consequence of the fact that giant planets are mostly made of hydrogen and helium: these elements had to be acquired when they were still present in the disk. Because the observed lifetime of gaseous circumstellar disks is of the order of a few million years, this implies that these planets formed (i.e. acquired most of their final masses) in a few million years also, quite faster than terrestrial planets in the Solar System. Giant planets migrated: Although not cleanly demonstrated yet, there is evidence that the observed orbital distribution of extrasolar planets requires an inward migration of planets, and various mechanisms have been proposed for that (see Ida and Lin, 2004a; Alibert et al., 2005; Moorhead and Adams, 2005, ...etc.). Separately, it was shown that several properties of our Solar System can be explained if Jupiter, Saturn, Uranus and Neptune ended up the early formation phase in the presence of a disk with quasi-circular orbit, and with Saturn, Uranus and Neptune significantly closer to the Sun than they are now, and that these three planets subsequently migrated outward (Tsiganis et al., 2005). Accretion played a key role for giant planet formation: Several indications point towards a formation of giant planets that is dominated by accretion of heavy elements: First, Jupiter, Saturn, Uranus and Neptune are all significantly enriched in heavy elements compared to the Sun. This feature can be reproduced by core-accretion models, 34 for Jupiter and Saturn at least (Alibert et al., 2005). Second, the probability to find a giant planet around a solar-type star (with stellar type F, G or K) is a strongly rising function of stellar metallicity (Gonzalez, 1998; Santos et al., 2004; Fischer and Valenti, 2005), a property that is also well-reproduced by standard core accretion models (Ida and Lin, 2004b; Alibert et al., 2005). Third, the large masses of heavy elements inferred in some transiting extrasolar planets as well as the apparent correlation between mass of heavy elements in the planet and stellar metallicity (Guillot et al. (2006); see also Sato et al. (2005) and Ikoma et al. (2006)) is a strong indication that accretion was possible and that it was furthermore efficient. It is to be noted that none of these key properties are directly explained by formation models that assume a direct gravitational collapse (see Boss, 2004; Mayer et al., 2004). Giant planets were enriched in heavy elements by core accretion, planetesimal delivery and/or formation in an enriched protoplanetary disk: The giant planets in our Solar System are unambigously enriched in heavy elements compared to the Sun, both globally, and when considering their atmosphere. This may also be the case of extrasolar planets, although the evidence is still tenuous. The accretion of a central core can explain part of the global enrichment, but not that of the atmosphere. The accretion of planetesimals may be a possible solution but in the case of Jupiter at least the rapid drop in accretion efficiency as the planet reaches appreciable masses (∼ 100 M⊕ or so) implies that such an enrichment would have originally concerned only very deep layers, and would require a relatively efficient upper mixing of these elements, and possibly an erosion of the central core (Guillot et al., 2004). Although not unambiguously explained, the fact that Jupiter is also enriched in noble gases compared to the Sun is a key observation to understand some of the processes occuring in the early Solar System. Indeed, noble gases are trapped into solids only at very low temperatures, and this tells us either that most of the solids that formed Jupiter were formed at very low temperature to be able to trap gases such as Argon, probably as clathrates (Gautier et al., 2001; Hersant et al., 2004), or that the planet formed in an enriched disk as it was being evaporated (Guillot and Hueso, 2006). 6 Future prospects We have shown that the compositions and structures of giant planets remain very uncertain. This is an important problem when attempting to understand and constrain the formation of planets, and the origins of the Solar System. However, the parallel study of giant planets in our Solar System by space missions such as Galileo and Cassini, and of extrasolar planets by both ground based and space programs has led to rapid improvements in the field, with in particular a precise determination of the composition of Jupiter’s troposphere, and constraints on the compositions of a dozen of extrasolar planets. Improvements on our knowledge of the giant planets requires a variety of efforts. Fortunately, nearly all of these are addressed at least partially by adequate projects in the next few years. The efforts that are necessary thus include (but are not limited to): • Obtain a better EOS of hydrogen, in particular near the molecular/metallic transition. This will be addessed by the construction of powerful lasers such as the NIF in the US and the M´egaJoule laser in France, and by innovative experiments such as shocks on pre-compressed samples. One of the challenges is not only obtaining higher pressures, but mostly lower tem- peratures than currently possible with single shocks. The parallel improvement of computing facilities should allow more extended numerical experiments. 35 • Calculate hydrogen-helium and hydrogen-water phase diagrams. (Other phase diagrams are desirable too, but of lesser immediate importance). This should be possible with new numerical experiments. • Have a better yardstick to measure solar and protosolar compositions. This may be addressed by the analysis of the Genesis mission samples, or may require another future mission. • Improve the values of J4 and J6 for Saturn. This will be done as part of the Cassini-Huygens mission. This should lead to better constraints, and possibly a determination of whether the interior of Saturn rotates as a solid body. • Detect new transiting extrasolar planets, and hopefully some that are further from their star. The space missions CoRoT (2006) and Kepler (2008) should provide the detection and characterization of many tens, possibly hundreds of giant planets. • Model the formation and evolution of ice giants such as Uranus, Neptune, and similar planets around other stars, in order to analyze detections of these objects and understand planetary formation. • Improve the measurement of Jupiter’s gravity field, and determine the abundance of water in the deep atmosphere. This will be done by the Juno mission (launch 2011) with a combination of an exquisite determination of the planet’s gravity field and of radiometric measurements to probe the deep water abundance. • It would be highly desirable to send a probe similar to the Galileo probe into Saturn’s atmosphere. The comparison of the abundance of noble gases would discriminate between different models of the enrichment of the giant planets, and the additional measurement of key isotopic ratio would provide further tests to understand our origins. Clearly, there is a lot of work on the road, but the prospects for a much improved knowledge of giant planets and their formation are bright. References Acuna, M. H., Connerney, J. E. P., Ness, N. F., 1983. The Z3 zonal harmonic model of Saturn’s magnetic field Analyses and implications. JGR 88, 8771–8778. Alibert, Y., Mordasini, C., Benz, W., Winisdoerffer, C., 2005. Models of giant planet formation with migration and disc evolution. A&A 434, 343–353. Anderson, J. D., Campbell, J. K., Jacobson, R. A., Sweetnam, D. N., Taylor, A. H., 1987. Radio science with Voyager 2 at Uranus - Results on masses and densities of the planet and five principal satellites. JGR 92, 14877–14883. Atreya, S. K., Wong, M. H., Owen, T. C., Mahaffy, P. R., Niemann, H. B., de Pater, I., Drossart, P., Encrenaz, T., 1999. A comparison of the atmospheres of Jupiter and Saturn: deep atmospheric composition, cloud structure, vertical mixing, and origin. Plan. Space Sci. 47, 1243–1262. Bahcall, J. N., Pinsonneault, M. H., Wasserburg, G. J., 1995. Solar models with helium and heavy-element diffusion. Reviews of Modern Physics 67, 781–808. 36 Baraffe, I., Chabrier, G., Barman, T. S., Allard, F., Hauschildt, P. H., 2003. Evolutionary models for cool brown dwarfs and extrasolar giant planets. The case of HD 209458. A&A 402, 701–712. Baraffe, I., Chabrier, G., Barman, T. S., Selsis, F., Allard, F., Hauschildt, P. H., 2005. Hot-Jupiters and hot-Neptunes: A common origin? A&A 436, L47–L51. Baraffe, I., Selsis, F., Chabrier, G., Barman, T. S., Allard, F., Hauschildt, P. H., Lammer, H., 2004. The effect of evaporation on the evolution of close-in giant planets. A&A 419, L13–L16. Barman, T. S., Hauschildt, P. H., Allard, F., 2001. Irradiated Planets. ApJ 556, 885–895. Barman, T. S., Hauschildt, P. H., Allard, F., 2005. Phase-Dependent Properties of Extrasolar Planet Atmospheres. ApJ 632, 1132–1139. Beaulieu, J.-P., Bennett, D. P., Fouqu´e, P., Williams, A., Dominik, M., Jorgensen, U. G., Kubas, D., Cassan, A., Coutures, C., Greenhill, J., Hill, K., Menzies, J., Sackett, P. D., Albrow, M., Brillant, S., Caldwell, J. A. R., Calitz, J. J., Cook, K. H., Corrales, E., Desort, M., Dieters, S., Dominis, D., Donatowicz, J., Hoffman, M., Kane, S., Marquette, J.-B., Martin, R., Meintjes, P., Pollard, K., Sahu, K., Vinter, C., Wambsganss, J., Woller, K., Horne, K., Steele, I., Bramich, D. M., Burgdorf, M., Snodgrass, C., Bode, M., Udalski, A., Szyma´nski, M. K., Kubiak, M., Wi¸eckowski, T., Pietrzy´nski, G., Soszy´nski, I., Szewczyk, O., Wyrzykowski, L., Paczy´nski, B., Abe, F., Bond, I. A., Britton, T. R., Gilmore, A. C., Hearnshaw, J. B., Itow, Y., Kamiya, K., Kilmartin, P. M., Korpela, A. V., Masuda, K., Matsubara, Y., Motomura, M., Muraki, Y., Nakamura, S., Okada, C., Ohnishi, K., Rattenbury, N. J., Sako, T., Sato, S., Sasaki, M., Sekiguchi, T., Sullivan, D. J., Tristram, P. J., Yock, P. C. M., Yoshioka, T., 2006. Discovery of a cool planet of 5.5 Earth masses through gravitational microlensing. Nature 439, 437–440. Belov, S. I., Boriskov, G. V., Bykov, A. I., Il’Kaev, R. I., Luk’yanov, N. B., Matveev, A. Y., Mikhailova, O. L., Selemir, V. D., Simakov, G. V., Trunin, R. F., Trusov, I. P., Urlin, V. D., Fortov, V. E., Shuikin, A. N., 2002. Shock Compression of Solid Deuterium. Journal of Experimental and Theoretical Physics Letteres 76, 433–+. B´ezard, B., Lellouch, E., Strobel, D., Maillard, J.-P., Drossart, P., 2002. Carbon Monoxide on Jupiter: Evidence for Both Internal and External Sources. Icarus 159, 95–111. Bodenheimer, P., Laughlin, G., Lin, D. N. C., 2003. On the Radii of Extrasolar Giant Planets. ApJ 592, 555–563. Bodenheimer, P., Lin, D. N. C., Mardling, R. A., 2001. On the Tidal Inflation of Short-Period Extrasolar Planets. ApJ 548, 466–472. Bonev, S. A., Militzer, B., Galli, G., 2004. Ab initio simulations of dense liquid deuterium: Comparison with gas-gun shock-wave experiments. Phys. Rev. B 69 (1), 014101–+. Boriskov, G. V., Bykov, A. I., Il’Kaev, R. I., Selemir, V. D., Simakov, G. V., Trunin, R. F., Urlin, V. D., Shuikin, A. N., Nellis, W. J., 2005. Shock compression of liquid deuterium up to 109 GPa. Phys. Rev. B 71 (9), 092104–+. Borysow, A., Jorgensen, U. G., Zheng, C., 1997. Model atmospheres of cool, low-metallicity stars: the importance of collision-induced absorption. A&A 324, 185–195. 37 Boss, A. P., 2004. Convective Cooling of Protoplanetary Disks and Rapid Giant Planet Formation. ApJ 610, 456–463. Briggs, F. H., Sackett, P. D., 1989. Radio observations of Saturn as a probe of its atmosphere and cloud structure. Icarus 80, 77–103. Burrows, A., Guillot, T., Hubbard, W. B., Marley, M. S., Saumon, D., Lunine, J. I., Sudarsky, D., 2000. On the Radii of Close-in Giant Planets. ApJL 534, L97–L100. Burrows, A., Lunine, J., 1995. Extrasolar Planets - Astronomical Questions of Origin and Survival. Nature 378, 333–+. Burrows, A., Marley, M., Hubbard, W. B., Lunine, J. I., Guillot, T., Saumon, D., Freedman, R., Sudarsky, D., Sharp, C., 1997. A Nongray Theory of Extrasolar Giant Planets and Brown Dwarfs. ApJ 491, 856–+. Burrows, A., Sudarsky, D., Hubbard, W. B., 2003. A Theory for the Radius of the Transiting Giant Planet HD 209458b. ApJ 594, 545–551. Busse, F. H., 1978. Magnetohydrodynamics of the Earth’s Dynamo. Annual Review of Fluid Mechanics 10, 435–462. Campbell, J. K., Synnott, S. P., 1985. Gravity field of the Jovian system from Pioneer and Voyager tracking data. AJ 90, 364–372. Cavazzoni, C., Chiarotti, G. L., Scandolo, S., Tosatti, E., Bernasconi, M., Parrinello, M., 1999. Superionic and Metallic States of Water and Ammonia at Giant Planet Conditions. Science 283, 44–+. Cecconi, B., Zarka, P., 2005. Model of a variable radio period for Saturn. Journal of Geophysical Research (Space Physics) 110, 12203–+. Chandrasekhar, S., 1939. An introduction to the study of stellar structure. Chicago, Ill., The University of Chicago press [1939]. Charbonneau, D., Allen, L. E., Megeath, S. T., Torres, G., Alonso, R., Brown, T. M., Gilliland, R. L., Latham, D. W., Mandushev, G., O’Donovan, F. T., Sozzetti, A., 2005. Detection of Thermal Emission from an Extrasolar Planet. ApJ 626, 523–529. Charbonneau, D., Brown, T. M., Noyes, R. W., Gilliland, R. L., 2002. Detection of an Extrasolar Planet Atmosphere. ApJ 568, 377–384. Cohen, E. R., Taylor, B. N., 1987. The 1986 adjustment of the fundamental physical constants. Reviews of Modern Physics 59, 1121–1148. Collins, G. W., da Silva, L. B., Celliers, P., Gold, D. M., Foord, M. E., Wallace, R. J., Ng, A., Weber, S. V., Budil, K. S., Cauble, R., 1998. Measurements of the equation of state of deuterium at the fluid insulator-metal transition. Science 281, 1178–1181. Connerney, J. E. P., Ness, N. F., Acuna, M. H., 1982. Zonal harmonic model of Saturn’s magnetic field from Voyager 1 and 2 observations. Nature 298, 44–46. 38 Conrath, B. J., Gautier, D., 2000. Saturn Helium Abundance: A Reanalysis of Voyager Measure- ments. Icarus 144, 124–134. Cooper, C. S., Showman, A. P., 2005. Dynamic Meteorology at the Photosphere of HD 209458b. ApJL 629, L45–L48. da Silva, L. B., Celliers, P., Collins, G. W., Budil, K. S., Holmes, N. C., Barbee, T. W., Jr., Hammel, B. A., Kilkenny, J. D., Wallace, R. J., Ross, M., Cauble, R., Ng, A., Chiu, G., 1997. Absolute Equation of State Measurements on Shocked Liquid Deuterium up to 200 GPa (2 Mbar). Physical Review Letters 78, 483–486. Datchi, F., Loubeyre, P., Letoullec, R., 2000. Extended and accurate determination of the melting curves of argon, helium, ice (H2O), and hydrogen (H2). Phys. Rev. B 61, 6535–6546. Davies, M. E., Abalakin, V. K., Bursa, M., Lederle, T., Lieske, J. H., 1986. Report of the IAU/IAG/COSPAR working group on cartographic coordinates and rotational elements of the planets and satellites - 1985. Celestial Mechanics 39, 103–113. de Pater, I., Romani, P. N., Atreya, S. K., 1991. Possible microwave absorption by H2S gas in Uranus’ and Neptune’s atmospheres. Icarus 91, 220–233. Deming, D., Seager, S., Richardson, L. J., Harrington, J., 2005. Infrared radiation from an extra- solar planet. Nature 434, 740–743. Desjarlais, M. P., 2003. Density-functional calculations of the liquid deuterium Hugoniot, reshock, and reverberation timing. Phys. Rev. B 68 (6), 064204–+. Fegley, B. J., Lodders, K., 1994. Chemical models of the deep atmospheres of Jupiter and Saturn. Icarus 110, 117–154. Ferguson, J. W., Alexander, D. R., Allard, F., Barman, T., Bodnarik, J. G., Hauschildt, P. H., Heffner-Wong, A., Tamanai, A., 2005. Low-Temperature Opacities. ApJ 623, 585–596. Fischer, D. A., Valenti, J., 2005. The Planet-Metallicity Correlation. ApJ 622, 1102–1117. Flasar, F. M., Achterberg, R. K., Conrath, B. J., Pearl, J. C., Bjoraker, G. L., Jennings, D. E., Romani, P. N., Simon-Miller, A. A., Kunde, V. G., Nixon, C. A., B´ezard, B., Orton, G. S., Spilker, L. J., Spencer, J. R., Irwin, P. G. J., Teanby, N. A., Owen, T. C., Brasunas, J., Segura, M. E., Carlson, R. C., Mamoutkine, A., Gierasch, P. J., Schinder, P. J., Showalter, M. R., Ferrari, C., Barucci, A., Courtin, R., Coustenis, A., Fouchet, T., Gautier, D., Lellouch, E., Marten, A., Prang´e, R., Strobel, D. F., Calcutt, S. B., Read, P. L., Taylor, F. W., Bowles, N., Samuelson, R. E., Abbas, M. M., Raulin, F., Ade, P., Edgington, S., Pilorz, S., Wallis, B., Wishnow, E. H., 2005. Temperatures, Winds, and Composition in the Saturnian System. Science 307, 1247–1251. Fortney, J. J., Hubbard, W. B., 2003. Phase separation in giant planets: inhomogeneous evolution of Saturn. Icarus 164, 228–243. Fortney, J. J., Saumon, D., Marley, M. S., Lodders, K., Freedman, R. S., 2006. Atmosphere, Interior, and Evolution of the Metal-rich Transiting Planet HD 149026b. ApJ 642, 495–504. Galopeau, P. H. M., Lecacheux, A., 2000. Variations of Saturn’s radio rotation period measured at kilometer wavelengths. JGR 105, 13089–13102. 39 Gautier, D., Conrath, B. J., Owen, T., De Pater, I., Atreya, S. K., 1995. The Troposphere of Neptune, pp. 547–611. Neptune and Triton, UofA Press. Gautier, D., Hersant, F., Mousis, O., Lunine, J. I., 2001. Erratum: Enrichments in Volatiles in Jupiter: A New Interpretation of the Galileo Measurements. ApJL 559, L183–L183. Giampieri, G., Dougherty, M. K., Smith, E. J., Russell, C. T., 2006. A regular period for Saturn’s magnetic field that may track its internal rotation. Nature 441, 62–64. Gierasch, P. J., Ingersoll, A. P., Banfield, D., Ewald, S. P., Helfenstein, P., Simon-Miller, A., Vasavada, A., Breneman, H. H., Senske, D. A., A4 Galileo Imaging Team, 2000. Observation of moist convection in Jupiter’s atmosphere. Nature 403, 628–630. Gonzalez, G., 1998. Spectroscopic analyses of the parent stars of extrasolar planetary system candidates. A&A 334, 221–238. Goukenleuque, C., B´ezard, B., Joguet, B., Lellouch, E., Freedman, R., 2000. A Radiative Equilib- rium Model of 51 Peg b. Icarus 143, 308–323. Gregoryanz, E., Goncharov, A. F., Matsuishi, K., Mao, H.-K., Hemley, R. J., 2003. Raman Spectroscopy of Hot Dense Hydrogen. Physical Review Letters 90 (17), 175701–+. Guillot, T., 1999a. A comparison of the interiors of Jupiter and Saturn. Plan. Space Sci. 47, 1183–1200. Guillot, T., 1999b. Interior of Giant Planets Inside and Outside the Solar System. Science 286, 72–77. Guillot, T., 2005. THE INTERIORS OF GIANT PLANETS: Models and Outstanding Questions. Annual Review of Earth and Planetary Sciences 33, 493–530. Guillot, T., Burrows, A., Hubbard, W. B., Lunine, J. I., Saumon, D., 1996. Giant Planets at Small Orbital Distances. ApJL 459, L35–L39. Guillot, T., Hueso, R., 2006. The composition of Jupiter: sign of a (relatively) late formation in a chemically evolved protosolar disc. MNRAS 367, L47–L51. Guillot, T., Santos, N. C., Pont, F., Iro, N., Melo, C., Ribas, I., 2006. A correlation between the heavy element content of transiting extrasolar planets and the metallicity of their parent stars. A&A 453, L21–L24. Guillot, T., Showman, A. P., 2002. Evolution of “51 pegasus b-like” planets. A&A 385, 156–165. Guillot, T., Stevenson, D. J., Hubbard, W. B., Saumon, D., 2004. The interior of Jupiter, pp. 35–57. Jupiter. The Planet, Satellites and Magnetosphere. Gulkis, S., Janssen, M. A., Olsen, E. T., 1978. Evidence for the depletion of ammonia in the Uranus atmosphere. Icarus 34, 10–19. Gurnett, D. A., Kurth, W. S., Hospodarsky, G. B., Persoon, A. M., Averkamp, T. F., Cecconi, B., Lecacheux, A., Zarka, P., Canu, P., Cornilleau-Wehrlin, N., Galopeau, P., Roux, A., Harvey, C., Louarn, P., Bostrom, R., Gustafsson, G., Wahlund, J.-E., Desch, M. D., Farrell, W. M., Kaiser, M. L., Goetz, K., Kellogg, P. J., Fischer, G., Ladreiter, H.-P., Rucker, H., Alleyne, H., Pedersen, A., 2005. Radio and Plasma Wave Observations at Saturn from Cassini’s Approach and First Orbit. Science 307, 1255–1259. 40 Hammel, H. B., de Pater, I., Gibbard, S. G., Lockwood, G. W., Rages, K., 2005. New cloud activity on Uranus in 2004: First detection of a southern feature at 2.2 µm. Icarus 175, 284–288. Harrington, J., Hansen, B. M., Luszcz, S. H., Seager, S., Deming, D., Menou, K., Cho, J. Y.-K., Richardson, L. J., 2006. The Phase-Dependent Infrared Brightness of the Extrasolar Planet υ Andromedae b. Science 314, 623–626. Hersant, F., Gautier, D., Lunine, J. I., 2004. Enrichment in volatiles in the giant planets of the Solar System. Plan. Space Sci. 52, 623–641. Holmes, N. C., Ross, M., Nellis, W. J., 1995. Temperature measurements and dissociation of shock-compressed liquid deuterium and hydrogen. Phys. Rev. B 52, 15835–15845. Hubbard, W. B., 1968. Thermal structure of Jupiter. ApJ 152, 745–754. Hubbard, W. B., 1977. The Jovian surface condition and cooling rate. Icarus 30, 305–310. Hubbard, W. B., 1989. Structure and composition of giant planet interiors, pp. 539–563. Origin and Evolution of Planetary and Satellite Atmospheres. Hubbard, W. B., Burrows, A., Lunine, J. I., 2002. Theory of Giant Planets. Ann. Rev. Astron. Astrophys. 40, 103–136. Hubbard, W. B., Guillot, T., Marley, M. S., Burrows, A., Lunine, J. I., Saumon, D. S., 1999. Comparative evolution of Jupiter and Saturn. Plan. Space Sci. 47, 1175–1182. Hubbard, W. B., Pearl, J. C., Podolak, M., Stevenson, D. J., 1995. The Interior of Neptune, pp. 109–138. Neptune and Triton, UofA Press. Hueso, R., S´anchez-Lavega, A., Guillot, T., 2002. A model for large-scale convective storms in Jupiter. Journal of Geophysical Research (Planets) 107, 5–1. Ida, S., Lin, D. N. C., 2004a. Toward a Deterministic Model of Planetary Formation. I. A Desert in the Mass and Semimajor Axis Distributions of Extrasolar Planets. ApJ 604, 388–413. Ida, S., Lin, D. N. C., 2004b. Toward a Deterministic Model of Planetary Formation. II. The Formation and Retention of Gas Giant Planets around Stars with a Range of Metallicities. ApJ 616, 567–572. Ikoma, M., Guillot, T., Genda, H., Tanigawa, T., Ida, S., 2006. On the Origin of HD 149026b. ApJ 650, 1150–1159. Ingersoll, A. P., Barnet, C. D., Beebe, R. F., Flasar, F. M., Hinson, D. P., Limaye, S. S., Sromovsky, L. A., Suomi, V. E., 1995. Dynamic Meteorology of Neptune, pp. 613–682. Neptune and Triton, UofA Press. Iro, N., B´ezard, B., Guillot, T., 2005. A time-dependent radiative model of HD 209458b. A&A 436, 719–727. Jacobson, R. A., Antreasian, P. G., Bordi, J. J., Criddle, K. E., Ionasescu, R., Jones, J. B., Mackenzie, R. A., Meek, M. C., Parcher, D., Pelletier, F. J., Owen, W. M., Jr., Roth, D. C., Roundhill, I. M., Stauch, J. R., 2006. The Gravity Field of the Saturnian System from Satellite Observations and Spacecraft Tracking Data. AJ 132, 2520–2526. 41 Kippenhahn, R., Weigert, A., 1994. Stellar Structure and Evolution. Stellar Structure and Evolu- tion, XVI, 468 pp. 192 figs.. Springer-Verlag Berlin Heidelberg New York. Also Astronomy and Astrophysics Library. Knudson, M. D., Hanson, D. L., Bailey, J. E., Hall, C. A., Asay, J. R., Deeney, C., 2004. Principal Hugoniot, reverberating wave, and mechanical reshock measurements of liquid deuterium to 400 GPa using plate impact techniques. Phys. Rev. B 69 (14), 144209–+. Kuchner, M. J., 2003. Volatile-rich Earth-Mass Planets in the Habitable Zone. ApJL 596, L105– L108. Lammer, H., Selsis, F., Ribas, I., Guinan, E. F., Bauer, S. J., Weiss, W. W., 2003. Atmospheric Loss of Exoplanets Resulting from Stellar X-Ray and Extreme-Ultraviolet Heating. ApJL 598, L121–L124. Laughlin, G., Marcy, G. W., Vogt, S. S., Fischer, D. A., Butler, R. P., 2005. On the Eccentricity of HD 209458b. ApJL 629, L121–L124. Lecavelier des Etangs, A., Vidal-Madjar, A., McConnell, J. C., H´ebrard, G., 2004. Atmospheric escape from hot Jupiters. A&A 418, L1–L4. L´eger, A., Selsis, F., Sotin, C., Guillot, T., Despois, D., Mawet, D., Ollivier, M., Lab`eque, A., Valette, C., Brachet, F., Chazelas, B., Lammer, H., 2004. A new family of planets? “Ocean- Planets”. Icarus 169, 499–504. Levrard, B., Morgado Correia, A., Chabrier, G., Baraffe, I., Selsis, F., Laskar, J., 2006. Tidal dissipation within hot Jupiters: a new appraisal. ArXiv Astrophysics e-prints. Lindal, G. F., 1992a. The atmosphere of Neptune - an analysis of radio occultation data acquired with Voyager 2. AJ 103, 967–982. Lindal, G. F., 1992b. The atmosphere of Neptune - an analysis of radio occultation data acquired with Voyager 2. AJ 103, 967–982. Lindal, G. F., Sweetnam, D. N., Eshleman, V. R., 1985. The atmosphere of Saturn - an analysis of the Voyager radio occultation measurements. AJ 90, 1136–1146. Lindal, G. F., Wood, G. E., Levy, G. S., Anderson, J. D., Sweetnam, D. N., Hotz, H. B., Buckles, B. J., Holmes, D. P., Doms, P. E., Eshleman, V. R., Tyler, G. L., Croft, T. A., 1981. The atmosphere of Jupiter - an analysis of the Voyager radio occultation measurements. JGR 86, 8721–8727. Little, B., Anger, C. D., Ingersoll, A. P., Vasavada, A. R., Senske, D. A., Breneman, H. H., Borucki, W. J., The Galileo SSI Team, 1999. Galileo Images of Lightning on Jupiter. Icarus 142, 306–323. Lodders, K., Fegley, B., Jr., 1994. The origin of carbon monoxide in Neptunes’s atmosphere. Icarus 112, 368–375. Lovis, C., Mayor, M., Pepe, F., Alibert, Y., Benz, W., Bouchy, F., Correia, A. C. M., Laskar, J., Mordasini, C., Queloz, D., Santos, N. C., Udry, S., Bertaux, J.-L., Sivan, J.-P., 2006. An extrasolar planetary system with three Neptune-mass planets. Nature 441, 305–309. 42 Mao, H.-K., Hemley, R. J., 1994. Ultrahigh-pressure transitions in solid hydrogen. Reviews of Modern Physics 66, 671–692. Marley, M. S., Fortney, J. J., Hubickyj, O., Bodenheimer, P., Lissauer, J. J., 2006. On the Luminosity of Young Jupiters. ArXiv Astrophysics e-prints. Marley, M. S., G´omez, P., Podolak, M., 1995. Monte Carlo interior models for Uranus and Neptune. JGR 100, 23349–23354. Marley, M. S., McKay, C. P., 1999. Thermal Structure of Uranus’ Atmosphere. Icarus 138, 268–286. Mayer, L., Quinn, T., Wadsley, J., Stadel, J., 2004. The Evolution of Gravitationally Unstable Protoplanetary Disks: Fragmentation and Possible Giant Planet Formation. ApJ 609, 1045– 1064. Mayor, M., Queloz, D., 1995. A Jupiter-Mass Companion to a Solar-Type Star. Nature 378, 355–+. Militzer, B., Ceperley, D. M., Kress, J. D., Johnson, J. D., Collins, L. A., Mazevet, S., 2001. Calculation of a Deuterium Double Shock Hugoniot from Ab Initio Simulations. Physical Review Letters 87 (26), A265502+. Moorhead, A. V., Adams, F. C., 2005. Giant planet migration through the action of disk torques and planet planet scattering. Icarus 178, 517–539. Ness, N. F., Acuna, M. H., Behannon, K. W., Burlaga, L. F., Connerney, J. E. P., Lepping, R. P., 1986. Magnetic fields at Uranus. Science 233, 85–89. Ness, N. F., Acuna, M. H., Burlaga, L. F., Connerney, J. E. P., Lepping, R. P., 1989. Magnetic fields at Neptune. Science 246, 1473–1478. Owen, T., Mahaffy, P., Niemann, H. B., Atreya, S., Donahue, T., Bar-Nun, A., de Pater, I., 1999. A low-temperature origin for the planetesimals that formed Jupiter. Nature 402, 269–270. Pearl, J. C., Conrath, B. J., 1991. The albedo, effective temperature, and energy balance of Neptune, as determined from Voyager data. JGR 96, 18921–+. Pfaffenzeller, O., Hohl, D., Ballone, P., 1995. Miscibility of Hydrogen and Helium under Astro- physical Conditions. Physical Review Letters 74, 2599–2602. Podolak, M., Hubbard, W. B., Stevenson, D. J., 1991. Model of Uranus’ interior and magnetic field, pp. 29–61. Uranus, UofA Press. Podolak, M., Podolak, J. I., Marley, M. S., 2000. Further investigations of random models of Uranus and Neptune. Plan. Space Sci. 48, 143–151. Podolak, M., Weizman, A., Marley, M., 1995. Comparative models of Uranus and Neptune. Plan. Space Sci. 43, 1517–1522. Rages, K., Hammel, H. B., Lockwood, G. W., 2002. A Prominent Apparition of Neptune’s South Polar Feature. Icarus 159, 262–265. 43 Ribas, I., 2006. Masses and Radii of Low-Mass Stars: Theory Versus Observations. Ap&SS 304, 89–92. Ross, M., 1998. Linear-mixing model for shock-compressed liquid deuterium. Phys. Rev. B 58, 669–677. Ross, M., Yang, L. H., 2001. Effect of chainlike structures on shock-compressed liquid deuterium. Phys. Rev. B 64 (13), 134210–+. Roulston, M. S., Stevenson, D. J., 1995. Prediction of neon depletion in Jupiter’s atmosphere. In: EOS, Volume 76, pp. 343. Salpeter, E. E., 1973. On Convection and Gravitational Layering in Jupiter and in Stars of Low Mass. ApJL 181, L83+. Sanchez-Lavega, A., Lecacheux, J., Gomez, J. M., Colas, F., Laques, P., Noll, K., Gilmore, D., Miyazaki, I., Parker, D., 1996. Large-scale storms in Saturn’s atmosphere during 1994. Sci- ence 271, 631–634. Santos, N. C., Israelian, G., Mayor, M., 2004. Spectroscopic [Fe/H] for 98 extra-solar planet-host stars. Exploring the probability of planet formation. A&A 415, 1153–1166. Sato, B., Fischer, D. A., Henry, G. W., Laughlin, G., Butler, R. P., Marcy, G. W., Vogt, S. S., Bodenheimer, P., Ida, S., Toyota, E., Wolf, A., Valenti, J. A., Boyd, L. J., Johnson, J. A., Wright, J. T., Ammons, M., Robinson, S., Strader, J., McCarthy, C., Tah, K. L., Minniti, D., 2005. The N2K Consortium. II. A Transiting Hot Saturn around HD 149026 with a Large Dense Core. ApJ 633, 465–473. Saumon, D., Chabrier, G., van Horn, H. M., 1995. An Equation of State for Low-Mass Stars and Giant Planets. ApJS 99, 713–+. Saumon, D., Guillot, T., 2004. Shock Compression of Deuterium and the Interiors of Jupiter and Saturn. ApJ 609, 1170–1180. Saumon, D., Hubbard, W. B., Burrows, A., Guillot, T., Lunine, J. I., Chabrier, G., 1996. A Theory of Extrasolar Giant Planets. ApJ 460, 993–+. Seager, S., Hui, L., 2002. Constraining the Rotation Rate of Transiting Extrasolar Planets by Oblateness Measurements. ApJ 574, 1004–1010. Seager, S., Sasselov, D. D., 2000. Theoretical Transmission Spectra during Extrasolar Giant Planet Transits. ApJ 537, 916–921. Seiff, A., Kirk, D. B., Knight, T. C. D., Young, R. E., Mihalov, J. D., Young, L. A., Milos, F. S., Schubert, G., Blanchard, R. C., Atkinson, D., 1998. Thermal structure of Jupiter’s atmosphere near the edge of a 5-µm hot spot in the north equatorial belt. JGR 103, 22857–22890. Showman, A. P., Guillot, T., 2002. Atmospheric circulation and tides of “51 Pegasus b-like” planets. A&A 385, 166–180. Showman, A. P., Ingersoll, A. P., 1998. Interpretation of Galileo Probe Data and Implications for Jupiter’s Dry Downdrafts. Icarus 132, 205–220. 44 Simon-Miller, A. A., Gierasch, P. J., Beebe, R. F., Conrath, B., Flasar, F. M., Achterberg, R. K., the Cassini CIRS Team, 2002. New Observational Results Concerning Jupiter’s Great Red Spot. Icarus 158, 249–266. Stevenson, D. J., 1982. Interiors of the giant planets. Annual Review of Earth and Planetary Sciences 10, 257–295. Stevenson, D. J., 1983. Planetary magnetic fields. Reports of Progress in Physics 46, 555–557. Stevenson, D. J., 1985. Cosmochemistry and structure of the giant planets and their satellites. Icarus 62, 4–15. Stevenson, D. J., Salpeter, E. E., 1977a. The dynamics and helium distribution in hydrogen-helium fluid planets. ApJS 35, 239–261. Stevenson, D. J., Salpeter, E. E., 1977b. The phase diagram and transport properties for hydrogen- helium fluid planets. ApJS 35, 221–237. Sudarsky, D., Burrows, A., Hubeny, I., 2003. Theoretical Spectra and Atmospheres of Extrasolar Giant Planets. ApJ 588, 1121–1148. Trafton, L. M., 1967. Model atmospheres of the major planets. ApJ 147, 765–781. Tsiganis, K., Gomes, R., Morbidelli, A., Levison, H. F., 2005. Origin of the orbital architecture of the giant planets of the Solar System. Nature 435, 459–461. Tyler, G. L., Sweetnam, D. N., Anderson, J. D., Borutzki, S. E., Campbell, J. K., Kursinski, E. R., Levy, G. S., Lindal, G. F., Lyons, J. R., Wood, G. E., 1989. Voyager radio science observations of Neptune and Triton. Science 246, 1466–1473. Valencia, D., O’Connell, R. J., Sasselov, D., 2006. Internal structure of massive terrestrial planets. Icarus 181, 545–554. Vidal-Madjar, A., D´esert, J.-M., Lecavelier des Etangs, A., H´ebrard, G., Ballester, G. E., Ehrenre- ich, D., Ferlet, R., McConnell, J. C., Mayor, M., Parkinson, C. D., 2004. Detection of Oxygen and Carbon in the Hydrodynamically Escaping Atmosphere of the Extrasolar Planet HD 209458b. ApJL 604, L69–L72. Vidal-Madjar, A., Lecavelier des Etangs, A., D´esert, J.-M., Ballester, G. E., Ferlet, R., H´ebrard, G., Mayor, M., 2003. An extended upper atmosphere around the extrasolar planet HD209458b. Nature 422, 143–146. von Zahn, U., Hunten, D. M., Lehmacher, G., 1998. Helium in Jupiter’s atmosphere: Results from the Galileo probe helium interferometer experiment. JGR 103, 22815–22830. Warwick, J. W., Evans, D. R., Peltzer, G. R., Peltzer, R. G., Romig, J. H., Sawyer, C. B., Riddle, A. C., Schweitzer, A. E., Desch, M. D., Kaiser, M. L., 1989. Voyager planetary radio astronomy at Neptune. Science 246, 1498–1501. Warwick, J. W., Evans, D. R., Romig, J. H., Sawyer, C. B., Desch, M. D., Kaiser, M. L., Alexander, J. K., Gulkis, S., Poynter, R. L., 1986. Voyager 2 radio observations of Uranus. Science 233, 102–106. 45 Weir, S. T., Mitchell, A. C., Nellis, W. J., 1996. Metallization of Fluid Molecular Hydrogen at 140 GPa (1.4 Mbar). Physical Review Letters 76, 1860–1863. Winn, J. N., Holman, M. J., 2005. Obliquity Tides on Hot Jupiters. ApJL 628, L159–L162. Wong, M. H., Mahaffy, P. R., Atreya, S. K., Niemann, H. B., Owen, T. C., 2004. Updated Galileo probe mass spectrometer measurements of carbon, oxygen, nitrogen, and sulfur on Jupiter. Icarus 171, 153–170. Yelle, R. V., 2006. Corrigendum to ‘Aeronomy of extra-solar giant planets at small orbital distances’ [Icarus 170 (2004) 167 179]. Icarus 183, 508–508. Zharkov, V. N., Trubitsyn, V. P., 1974. Internal constitution and the figures of the giant planets. Physics of the Earth and Planetary Interiors 8, 105–107. Zharkov, V. N., Trubitsyn, V. P., 1978. Physics of planetary interiors. Astronomy and Astrophysics Series, Tucson: Pachart, 1978. 46
1001.2006
2
1001
2010-01-15T21:56:08
The TAOS Project: Upper Bounds on the Population of Small KBOs and Tests of Models of Formation and Evolution of the Outer Solar System
[ "astro-ph.EP", "astro-ph.IM" ]
We have analyzed the first 3.75 years of data from TAOS, the Taiwanese American Occultation Survey. TAOS monitors bright stars to search for occultations by Kuiper Belt Objects (KBOs). This dataset comprises 5e5 star-hours of multi-telescope photometric data taken at 4 or 5 Hz. No events consistent with KBO occultations were found in this dataset. We compute the number of events expected for the Kuiper Belt formation and evolution models of Pan & Sari (2005), Kenyon & Bromley (2004), Benavidez & Campo Bagatin (2009), and Fraser (2009). A comparison with the upper limits we derive from our data constrains the parameter space of these models. This is the first detailed comparison of models of the KBO size distribution with data from an occultation survey. Our results suggest that the KBO population is comprised of objects with low internal strength and that planetary migration played a role in the shaping of the size distribution.
astro-ph.EP
astro-ph
Draft version October 29, 2018 Preprint typeset using LATEX style emulateapj v. 03/07/07 THE TAOS PROJECT: UPPER BOUNDS ON THE POPULATION OF SMALL KBOS AND TESTS OF MODELS OF FORMATION AND EVOLUTION OF THE OUTER SOLAR SYSTEM F. B. Bianco1, 2, 3, 4, Z.-W. Zhang5,6, M. J. Lehner5,3,4, S. Mondal7, S.-K. King5, J. Giammarco8,9, M. J. Holman4, N. K. Coehlo10, J.-H. Wang5,6, C. Alcock4, T. Axelrod11, Y.-I. Byun12, W. P. Chen6, K. H. Cook13, R. Dave14, I. de Pater15, D.-W. Kim12,14, T. Lee5, H.-C. Lin5, J. J. Lissauer16, S. L. Marshall17, P. Protopapas14,4, J. A. Rice10, M. E. Schwamb18, S.-Y. Wang5 and C.-Y. Wen5 0 1 0 2 n a J 5 1 . ] P E h p - o r t s a [ 2 v 6 0 0 2 . 1 0 0 1 : v i X r a Draft version October 29, 2018 ABSTRACT We have analyzed the first 3.75 years of data from TAOS, the Taiwanese American Occultation Survey. TAOS monitors bright stars to search for occultations by Kuiper Belt Objects (KBOs). This dataset comprises 5 × 105 star-hours of multi-telescope photometric data taken at 4 or 5 Hz. No events consistent with KBO occultations were found in this dataset. We compute the number of events expected for the Kuiper Belt formation and evolution models of Pan & Sari (2005), Kenyon & Bromley (2004), Benavidez & Campo Bagatin (2009), and Fraser (2009). A comparison with the upper limits we derive from our data constrains the parameter space of these models. This is the first detailed comparison of models of the KBO size distribution with data from an occultation survey. Our results suggest that the KBO population is comprised of objects with low internal strength and that planetary migration played a role in the shaping of the size distribution. Subject headings: Kuiper Belt, occultations, Solar System: formation 1. INTRODUCTION The Kuiper Belt has been shaped by accretion and disruption processes throughout the history of the Solar System. With small orbital eccentricities, the relative velocities of the objects in the early Kuiper Belt were sufficiently low to allow accretion processes to form kilo- meter and much larger objects. Later, the velocity dis- Electronic address: [email protected] 1 Department of Physics, University of California Santa Barbara, Mail Code 9530, Santa Barbara CA 93106-9530 2 Las Cumbres Observatory Global Telescope Network, Inc. 6740 Cortona Dr. Suite 102, Santa Barbara, CA 93117 3 Department of Physics and Astronomy, University of Pennsyl- vania, 209 South 33rd Street, Philadelphia, PA 19104 4 Harvard-Smithsonian Center for Astrophysics, 60 Garden Street, Cambridge, MA 02138 5 Institute of Astronomy and Astrophysics, Academia Sinica. P.O. Box 23-141, Taipei 10617, Taiwan 6 Institute of Astronomy, National Central University, 300 Jhongda Rd, Jhongli 32054, Taiwan 7 Aryabhatta Research Institute of Observational Sciences (ARIES), Manora Peak, Nainital-263 129, INDIA 8 Department of Astronomy and Physics, Eastern University 1300 Eagle Road Saint Davids, PA 19087 9 Department of Physics, Villanova University, 800 Lancaster Avenue, Villanova, PA 19085 10 Department of Statistics, University of California Berkeley, 367 Evans Hall, Berkeley, CA 94720 11 Steward Observatory, 933 North Cherry Avenue, Room N204 Tucson AZ 85721 12 Department of Astronomy, Yonsei University, 134 Shinchon, Seoul 120-749, Korea 13 Institute for Geophysics and Planetary Physics, Lawrence Liv- ermore National Laboratory, Livermore, CA 94550 14 Initiative in Innovative Computing at Harvard, 60 Oxford St., Cambridge MA 02138 15 Department of Astronomy, University of California Berkeley, 601 Campbell Hall, Berkeley CA 94720 16 Space Science and Astrobiology Division 245-3, NASA Ames Research Center, Moffett Field, CA, 94035 17 Kavli Institute for Particle Astrophysics and Cosmology, 2575 Sand Hill Road, MS 29, Menlo Park, CA 94025 18 Division of Geological and Planetary Sciences, California In- stitute of Technology, 1201 E. California Blvd., Pasadena, CA 91125 persion increased, possibly as the KBO population was stirred up by the gravitational effects of the larger plan- ets and planetoids. Only large objects were then able to continue growing through impacts, whereas collisions among smaller bodies resulted in disruption. The de- tails of these processes depend on the internal strength of the KBOs and on the orbital and dynamical evolution of the gas giant planets. The size distribution of KBOs, therefore, contains information on the internal structure and composition of the KBOs -- and hence information on the location and epoch in which they formed -- and on planetary migration (Kenyon et al. 2008, and refer- ences therein). Direct observations have detected KBOs as faint as magnitude R ∼ 28.2 (Bernstein et al. 2004), which corresponds to a diameter of about 27 km as- suming a 4% albedo. The large end side of the KBO size distribution can therefore be characterized through its brightness distribution. The latter is well described by a power law Σ(< R) = 10α(R−R0), with an in- dex α = 0.6 and R0 = 23 (Fraser & Kavelaars 2009 Fuentes & Holman 2008) for objects brighter than about R = 25, or D ∼ 100 km. This is the region of the size spectrum which reflects the early history of agglomera- tion. Kenyon & Windhorst (2001) pointed out that the intensity of the infrared Zodiacal Background sets limits on the extrapolation of a straight power law to smaller sizes. The relatively shallow size distribution of Jupiter Family Comets (JFCs, Tancredi et al. 2006), which are believed to originate in the Kuiper Belt, and the cra- tering of Triton observed by Voyager 2 (Stern 1996), all point to a flatter distribution for small KBOs1. In 2004 evidence surfaced that a break in the power law oc- curs at a diameter larger than 10 km: Bernstein et al. (2004) conducted deep Hubble Space Telescope obser- vations with the Advanced Camera for Surveys which 1 The relationship between the cratering of Triton and the Kuiper Belt size distribution is questioned by Schenk & Zahnle (2007). 2 led to the discovery of only 3 new objects fainter than R = 26, about 4% of the number expected from a sin- gle power law distribution extrapolated to 10 km. While this work remains the state of the art for deep direct sur- veys of the Outer Solar System, recent campaigns have observed many more faint objects down to magnitude R = 27, which with the assumption of a 4% albedo corre- sponds to about 40 km in diameter2 (Fraser & Kavelaars 2008, Fuentes & Holman 2008, Fuentes et al. 2009, and Fraser & Kavelaars 2009). These recent data allowed them to locate a break in the power law size distribu- tion at diameters 30 . D . 120 km. The region of the size spectrum between tens of kilo- meters and meters in diameter is particularly interest- ing as models predict here the occurrence of transi- tions between different regimes where the binding en- ergy of KBOs is dominated either by gravity or internal strength. These transitions would leave a signature in the size distribution (Pan & Sari 2005, Kenyon & Bromley 2004, Benavidez & Campo Bagatin 2009, and references therein). Occultation surveys allow us to reach farther then the current limits of direct observations, and into this region of interest. These surveys monitor back- ground stars in order to detect the chance alignment of a KBO with a target star, which would generate a variation in the observed flux of the star. At distances in the Outer Solar System (tens to thousands of AU) the signature left in a lightcurve by the transits of D ∼ 1 km objects is dominated by diffraction. This technique requires high frequency photometric time series as the time scale for an occultation by an Outer Solar System Object is a frac- tion of a second (Roques & Moncuquet 2000, Nihei et al. 2007, Bickerton et al. 2009). A few such surveys have been attempted in the past several years and have recently started reporting results: e.g., Roques et al. (2006), Chang et al. (2007), Bickerton et al. (2008), Liu et al. (2008), Zhang et al. (2008) -- hereinafter Z08 -- , Bianco et al. (2009), and Wang et al. (2009). None of these surveys have claimed detections in the Kuiper Belt; upper limits have thus been placed on the number den- sity of KBOs in the sky. Bickerton et al. (2008) set an upper limit to the sky density of KBOs of ΣN (D ≥ 1 km) ≤ 2.8 × 109 deg−2 using the 40 Hz data from their own survey as well as the 45 Hz data from Roques et al. (2006) and the X-ray data from Chang et al. (2007). Bianco et al. (2009) car- ried out a 30 Hz survey with Megacam at the MMT setting a more stringent limit of ΣN (D ≥ 1 km) ≤ 2.0 × 108 deg−2 and a limit of ΣN (D ≥ 0.7 km) ≤ 4.8 × 108 deg−2. Wang et al. (2009) reported prelimi- nary analysis of videomode engineering data taken with the Pan-STARRS system. Recently Schlichting et al. (2009) reported the detec- tion of a candidate occultation event consistent with a D ∼ 1 km KBO in the analysis of archival guiding data from HST, and an estimate of the sky density of KBOs of ΣN (D ≥ 0.5 km) = 2.1+4.8 −1.2 × 107 deg−2. The Taiwanese American Occultation Survey (TAOS) has been operating since 2005 with two, three, and now 2 The magnitude of KBOs is converted into diameter by assum- ing a nominal 4% albedo throughout the paper, note however that Fraser & Kavelaars 2008 and Fraser & Kavelaars 2009 assumed an albedo of 6% in their work. Dataset parameters (3 -- telescope data) TABLE 1 Z08 this work Start Date End Date Light-curve sets Exposure (star -- hours) Tripletsa 2005 February 7 2006 December 31 110,554 152,787 2.6 × 109 2005 February 7 2008 August 2 366,083 500,339 9.0 × 109 a Multi -- telescope measurements. four telescopes simultaneously taking stellar photometry at 5 Hz3. The analysis of the first two years of TAOS reported no detections (Z08) and an upper limit was derived to the slope of the small size end of the size spectrum. The TAOS system is described in detail in Lehner et al. (2009). Using 50 cm aperture robotic tele- scopes in simultaneous observations and observing with the relatively low cadence compared to the aforemen- tioned occultation surveys, TAOS was designed to ad- dress the km-size region of the KBO size spectrum. We will show here that the marginal sensitivity to sub-km objects is more than compensated by the very large ex- posure of our star targets. Here we consider the first 3.75 years of TAOS data, a significantly larger dataset than the one explored in Z08. With these data we are able to constrain Kuiper Belt formation and evolution models. In Section 2 we describe the new dataset. In Section 3 we briefly describe our detection algorithms, as well as our efficiency analysis. We also discuss our recovery efficiency and discuss the most productive strategies for TAOS and the other occultation surveys, and, in Sec- tion 3.4, we derive the effective coverage of our survey. In Section 4 we derive model -- independent limits to the number of objects in the Kuiper Belt, and we compare our results with those of similar surveys. In Section 5 we briefly describe models for the formation and evolution of the Kuiper Belts and we then derive and discuss con- straints to these models. In Section 6 we compare our upper limits to the estimates on the number of KBOs set by dynamical simulations for JFC progenitor popu- lations. Finally we summarize and discuss our findings in Section 7. 2. 3.75 YEARS OF TAOS DATA TAOS is a dedicated survey that observes at a cadence of 5 Hz. The primary scientific goal of the survey is to estimate or set constraints on the number of KBOs in the region of the size spectrum that is currently too small to be observed directly: D . 10 km. Here we present an expanded analysis of three- telescope TAOS data4. These data consist of photo- metric measurements of target star fields collected syn- chronously with all three telescopes. The dataset ana- lyzed here was collected between January 2005 and Au- gust 2008. In a previous analysis of a subset of these data, Z08 reported an upper limit to the size distribu- tion of KBOs under the assumption of a single power law for small KBOs. If one models the size distribution 3 A small subset of early data was collected at 4 Hz cadence, comprising about 5% of the data analyzed in this work. 4 The fourth telescope, TAOS C, became operational in August 2008. The results presented in this paper are based on analysis of all of the three-telescope data collected to this point. 3 Fig. 1. -- Distribution of magnitudes for the TAOS target stars, (bin size 0.16 mag, left). SNR for the TAOS target stars, averaged over the duration of a run and over the three telescopes (bin size 0.73, right). A few targets at greater SNR and brighter magnitude, amounting to < 5% of the data, are not shown. for objects smaller than D = 28 km, the smallest direct observation (Bernstein et al. 2004), as a single power law dN /dD ∝ D−q, where N is the surface density of ob- jects, the slope of the distribution is limited to q ≤ 4.6. Throughout the remainder of this paper, a data run refers to a set of data collected in an uninterrupted ob- servation of any field. For a single star in the field a set of three lightcurves belonging to one data run will be referred to as a lightcurve set, and each three-telescope measurement, at a single time point, will be referred to as Fig. 2. -- SNR versus TAOS instrumental magnitude MTAOS; only a random sample of 1% of all stars is shown for clarity. a triplet. A star -- hour refers to an hour of high-cadence, multi-telescope observations on a single target star. The data set described in this paper amounts to amount to 5.0 × 105 star -- hours, while the data set used in Z08 comprises 1.5 × 105 star -- hours. The details of this dataset, and of the dataset published in Z08, are sum- marized in Table 1. Over 90% of our data is collected within 5◦ of the ecliptic plane in order to maximize the rate of occultations. TAOS uses the zipper -- mode technique to read out the CCD cameras at high frequency. This method, described in Lehner et al. (2009), enables high speed observations across the 3✷◦ field of view of the TAOS telescopes, but it artificially increases the crowding of the field and the background. In zipper -- mode readout each star in the field is represented in a subsection of the output image -- which we call rowblock and which comprises 76 rows for our 5 Hz data -- so that the field of view is entirely im- aged in each rowblock. Note that the images of different stars in a rowblock, however, do not necessarily belong to the same epoch. The zipper -- mode readout boosts the sky background by a factor of 27 at a 5 Hz readout rate. This limits the sensitivity of TAOS to stars as faint as MTAOS = 13.5, for which a signal -- to -- noise ratio (SNR) of ∼ 7 can be achieved in a dark night. The magnitude and SNR distributions for the target stars in our survey are shown in Figure 1. On the left panel, the x−axis is the TAOS instrumental magnitude MTAOS, which is defined by a regression on the USNO-B magnitudes to be similar to RUSNO. The correlation between instru- mental magnitude and SNR is shown in Figure 2. The scatter in the relationship between SNR and MTAOS is due to both changes in the sky background and in the weather conditions, and to different degrees of crowding in the fields. In Figure 3, left, we show the number of star -- hours at different angles from opposition. The top 4 Fig. 3. -- Left: distribution of angles from opposition for the TAOS targets. The top axis shows the relative velocity of a KBO at 43 AU, given the position of the field. The bin size is 10.5◦. Right: distribution of ecliptic latitude for the TAOS target fields (center of the field is assumed), bin size ∼ 2.5◦ scale indicates the velocity of a KBO at the center of a field at this elongation. We cover a large range of oppo- sition angles; our field selection algorithm favors ecliptic fields near zenith. Most angles are positive because the weather at the site tends to improve after midnight. The right panel of Figure 3 shows the distribution of ecliptic latitude of our data. The effects of the angle from op- position on our efficiency and event rate, as well as the efficiency as a function of magnitude and crowding are discussed further in Section 3.3. 3. ANALYSIS The first step in the analysis is the photometric reduc- tion of the zipper mode images in the data set. A custom aperture photometry package (Zhang et al. 2009) is used to measure the brightness of each star at each epoch, and the resulting series of flux measurements are then assem- bled into lightcurves for subsequent analysis, which is described in the following subsections. In Section 3.1 we describe our detection algorithm and the rejection of false positives. We then describe the efficiency tests: in Sec- tion 3.2 we describe in detail how we identify the angular size of our target stars to simulate occultations correctly. In Section 3.3 we show how we simulate and implant oc- cultation events in our lightcurves, and test the behavior of our efficiency as a function of various parameters rel- ative to the occultations and to the observing strategy. Finally, we can derive the effective coverage of our survey (Section 3.4). 3.1. Event Detection and False Positive Rejection 1 The Fresnel scale is defined as F = (λ∆/2) 2 where λ is the wavelength of observation, and ∆ the distance to the occulter (Roques et al. 1987; Born & Wolf 1980). For optical observations at the distance of the Kuiper Belt (about 43 AU) the Fresnel scale is F ≈ 1.4 km. Occul- tation events will therefore exhibit significant diffraction effects. Occultations are manifested in the lightcurve of an observed star as an alternation of bright and dark features, typically with an overall suppression of the flux. Theoretical occultation lightcurves are shown in Figure 5. The signature of an occultation by a KBO of sub-kilometer size has a duration of about 0.2 sec- ond at opposition, and about a second near quadrature. A typical KBO occultation is then expected to result in the suppression of the flux for one or a few consecu- tive points in a TAOS lightcurve. In order to ascertain the extra-terrestrial origin of a dip in a lightcurve TAOS observes simultaneously with multiple telescopes. This allows us to rule out, on the basis of simple parallax con- siderations, atmospheric scintillation phenomena which might mimic an occultation event and which could be a source of false positives in occultation surveys, as well as any non-atmospheric phenomena such as birds, air- planes, etc. In order to detect occultations we need to identify brief flux changes in a star simultaneously observed by all tele- scopes. The statistical significance of a simultaneous low point in our lightcurves can be assessed rigorously, and the probability of a low measurement being drawn out of pure noise decreases with the number of telescopes ob- serving the target, provided that the measurements for the telescopes are independent. The lightcurves are high- pass filtered to remove trends due to weather patterns and changes in atmospheric transparency. High-pass fil- tering the lightcurves preserves the information on time- scales relevant to occultation phenomena (one or a few points in a time series). The implementation of the filter is described in Z08. The filter produces a time series in which h(t), the measurement taken at time t, represents the deviation from the local mean of the lightcurve in units of local standard deviation. To detect events we rank -- order the photometric mea- 5 Fig. 4. -- Angular size distribution for a typical TAOS field (field 120, RA: 13◦.7, Dec: −10◦.7) derived from the 2MASS K − J colors (a). The curves show the theoretical behavior of the angular size for A0, F0, G0, K0 dwarf stars (thin lines) and G0, K0, K5 giants (thick lines). The size of the Fresnel scale at 43 AU is shown at 0.08 mas (dashed line). Best fit to the angular size distribution (b): x -axis is the USNO-B B − R color and y is the angular size derived using Equations 1 & 2, converting all stars to apparent magnitude R = 12. Implanted angular sizes, derived from USNO-B the B − R color for our simulation (c) and angular size of the star for which events are recovered (d); a random subset of 1% of our data is plotted in the bottom panels. i rB i rD i /N 3 i , rB i , rD surement in each of our lightcurves, from the lowest to the highest flux, independently for each telescope (la- beled A, B and D). The i-th point in a lightcurve will be associated to rank rT for telescope T. We then consider i the rank triplets (rA i ). The probability distri- bution of the quantity zi = − ln{rA p }, with Np the number of points in the lightcurve set, can be de- termined combinatorially. Knowing this, under the null hypothesis that there is no event in the triplet i, we can compute the probability for a random variable Z aris- ing from this distribution P (Z > zi) = ξ. We set a threshold such that we expect fewer than 0.27 events in our dataset that are due to random fluctuations. For the dataset discussed in this paper we accept as events all data points that produce a rank product less likely than ξ = 3.0 × 10−11 to be drawn from a random distribution. Note that events generated by large KBOs, or for ob- servations near quadrature, would affect more than one point in the lightcurve (Figure 5), and our rank -- based search algorithm is most efficient when the dip in the lightcurve is isolated. Therefore, in addition to search- ing for single -- point events, we also bin our lightcurves by 2, re-rank them and repeat the statistical tests de- scribed above. The probability of each data point is as- sessed for both unbinned and binned lightcurves. Each lightcurve is binned twice, with two different starting points. This increases the detectability of occultations by large KBOs and by KBOs transiting with low rela- tive velocity. For a detailed discussion of our statistical analysis see Lehner et al. (2010). For a set of lightcurves of a given star, the ranks in the three telescopes should not be correlated for the statis- tical analysis described above to be valid. We have de- veloped a series of statistical tests to identify data runs where significant correlations (typically due to fast mov- ing cirrus clouds) are found in the lightcurves. In such data runs the ranks are not independently distributed and thus we can not accurately determine the statisti- cal significance of any candidate events. Any data run where the independence of the measurements after filter- ing cannot be rigorously established is removed from our dataset. A complete description of these tests is beyond the scope of this paper, but they are discussed in detail in Lehner et al. (2010) For the next step we relax considerably the ulti- mate significance requirement described above, and se- lect as provisional candidates those triplets that have ξ ≤ 1.0 × 10−6. Note that the significance ξ refers to the probability that the point would be drawn from a ran- dom distribution, therefore the lower the value of ξ the higher the statistical significance of the event. Nearly 150,000 provisional candidates are found. We use all these measurements to identify and remove spurious re- gions of the lightcurves, and hence identify and reject false positive events which arise from sources other than random chance. The constraints described below allows us to recognize regions of lightcurves with atypical noise and contamination by transiting objects (satellites or me- teors, which turn out to be the major source of false alarms), and to identify high-frequency fluctuations in the raw data that are not removed by the high-pass fil- ter. These are the steps of our false positive rejection process: • Contiguity: Contiguous candidates within a lightcurve and candidates that are within three time-stamps of each other are removed. Only the one rank triplet that has the highest significance in a series of contiguous or proximate points is 6 Fig. 5. -- Steps of the generation of a simulated occultation event. Top, the left and right panels both show the same point source lightcurve for a 3 km KBO at 43 AU occulting an F0V star. Second row: finite source lightcurves for the same occultation parameters for a V = 11 star, corresponding to an angular size of 0.015 mas, for a zero impact parameter (left), and at an impact parameter of 2 km (right). Row three: the lightcurves in row two are integrated over intervals of 105 ms for the occultation above at opposition (left) and at 50◦ from opposition (right). The lightcurves are sampled at 5 Hz, with no time offset (left) and with a time offset of 50 ms (right). considered as a candidate. This removes double -- counted events: events caused by large KBOs or KBOs moving at low relative velocity would affect more than one contiguous point. Furthermore this removes events that are double -- counted because they appear significant in both the binned and un- binned lightcurves. This eliminates about 40% of the candidates. • Simultaneity: Candidates that appear in the lightcurves of more than one star simultaneously at the same time-stamp or within three rowblocks are considered to be false positives. We expect si- multaneous count drops in time-domain to be pri- marily due to inaccurate aperture positioning in the photometry. In the rowblock domain simulta- neous count drops might be due to inaccurate back- ground determination or non-occultation events al- tering the baseline of the lightcurve at, or around, the candidate event, or by fast moving cirrus clouds or other phenomena which induce high frequency fluctuations in multiple lightcurves which are not removed by the high-pass filter. This cut removes about 60% of the remaining candidates. • Number of telescopes: At this point we require all of our remaining candidates to have been observed by all three telescopes. Although we are only consid- ering three-telescope runs in our analysis, for some targets the lightcurve might not be extracted in the photometry phase for all telescopes. Small differ- ences in the field of view and in the field distortion might make one star target not visible to all tele- scopes if it is at the edge of the field or if the crowd- ing induced by the zipper -- mode readout caused overlap of the target with other stars (Lehner et al. 2009). About 35% of the remaining candidates are thus removed. Note that this cut cannot be applied earlier as simultaneous events might appear in only one 3 -- telescope lightcurve, but also in 2 -- telescope lightcurves, and we want to be able to recognize and remove these events. At this point there are still over 20,000 candidates left. Note that we do not count the star hours discarded by this cut in our total exposure of 5 × 105 star -- hours. • Significance threshold: We finally constrain ξ such as to expect fewer than 0.27 false positives in our dataset. This constraint depends on the size of the dataset: for the 9 × 109 triplets remaining ξ < 3.0 × 10−11 allows < 0.27 false positives due to random noise. Only 228 candidates remain. The remaining candidates require visual inspection: first of the lightcurves, and for any remaining candi- dates, of the images. Most of the events are caused by the passing of bright objects, such as artificial satellites, meteorites or asteroids, that generate a variation in the background or baseline of the lightcurve responsible for causing artificially low counts in the neighborhood of the object. Many, but not all, of these false positives are removed by the simultaneity cut described above. Note that in our observing mode bright stars generate a bright streak across the length of our images (see Zhang et al. 2009), as flux is collected during the shutterless row shift- ing. In the presence of a bright object overlapping with a star -- streak generated by the zipper -- mode readout, the brightness of the streak is overestimated, thus too much flux is subtracted from the rowblock column causing an artificial flux drop in the star time series. In many in- stances the foreground object will also appear inside the star aperture artificially boosting its brightness. This flux drop will then be associated with a very high flux measurement following or preceding the event epoch, a signature that allows us to remove these false positives by inspecting the lightcurve. We also inspect the centroid position of the aperture. If the aperture position has moved significantly at the time-stamp of the candidate the candidate is rejected. Of the remaining candidates, 90% are rejected by visual inspection of the time series. Finally we inspect the images of the remaining 23 can- didates: they also were all associated with bright moving objects overlapping star streaks. No candidate events were left in our dataset at the conclusion of this process. 3.2. Determination of the Stellar Angular Size The shape of a lightcurve during an occulta- tion event, and hence the detection efficiency, is strongly dependent on the angular size of the tar- get star (Roques & Moncuquet 2000, Nihei et al. 2007, Bickerton et al. 2009). Our fields contain a variety of stellar types and a large range of angular sizes (Fig- ure 4a). Here we describe the method we use to estimate the angular sizes of our target stars in order to account for this effect when estimating our detection efficiencies (see Section 3.3). Angular sizes have been related to the position of a star in the color-color or color-magnitude diagrams (e.g., van Belle 1999, Nordgren et al. 2002). We follow the work of Nordgren et al. (2002) and calculate the angu- lar size of our star targets using the 2MASS J and K color (Cutri et al. 2003) to invert the set of equations: FK = (3.942 ± 0.006) − (0.095 ± 0.007)(J − K) FK = 4.2207 − 0.1K − 0.5 log θ (1) (2) where FK is the surface brightness of a star in K-band, which is related to its J − K color, as well as to its unreddened apparent K magnitude and angular size θ. The relationship between the surface brightness and the color of a star (Equation 1) is calibrated using angu- lar sizes measured directly by long baseline interferome- try (Nordgren et al. 2002). Not all of our target stars, however, are identified 2MASS objects, while in the photometry phase we have identified all of our targets with USNO-B objects. We therefore devised a method that relies on USNO-B R and B magnitude to calculate the angular sizes of our targets. We first derive the angular size of a subset of targets identified with 2MASS objects using the above equations, and scale it to obtain the angular size the targets would have if their apparent magnitude were R = 12. We then considered the USNO-B B − R color for all of these targets and calculated a regression on these points. This generates a formula that allows us to go from the USNO- B color of any of our targets to 2MASS colors and thus predict angular sizes according to Equations 1 & 2, for an apparent magnitude R = 12. To calculate the true 5. The angular size we rescale from R = 12 to RUSNO angular sizes of a subset of TAOS targets, rescaled to R = 12, is plotted as derived from Equation 1 & 2 versus the USNO-B B − R color (Figure 4b). Our regression on the data is plotted as well (solid line). The scatter in the determination of the angular size via the method described above is large, as can be seen in Figure 4b. This is due to scatter in the USNO-B color 5 We do not use our instrumental magnitude for rescaling for consistency with what is used in the color determination. 7 TABLE 2 Distribution of synthetic events diameter (km) implantations recoveries 30.0 8.0 3.0 2.0 1.3 1.0 0.7 0.5 231 385 1078 2003 4393 13255 36222 447764 75 84 73 89 73 66 40 9 (≈ 0.3 mag, Monet et al. 2003), to the (much smaller) scatter in the J and K magnitudes, and to the scatter in the empirical determination of the relationship between θ and J − K in Equation 1 & 2. We have not used any interstellar reddening corrections, and the angular size estimation of an unknown reddened star from the near- IR relationship would be relatively less affected compared to that in visual bands. Reddening is typically small for our targets though, since we are only considering objects brighter than R ∼ 13.5. The distribution of angular sizes is well reproduced. Figure 4a shows the distribution of angular sizes for a typical TAOS field, calculated via Equations 1 & 2, and Figure 4c shows the distribution of angular sizes in our efficiency simulation obtained via the USNO-B color. The distributions do overlap. About 2% of our simulated angular sizes fall in the region θ > 0.15 mas and RTAOS > 11, where there are no observed objects. These objects have poor USNO-B color determination. Figure 4d shows the region of the θ − RTAOS space where simulated events are recovered. There are few recoveries in the region θ > 0.15 mas and RTAOS > 11, so these stars do not contribute the the expected event rate. 3.3. Detection Efficiency It is necessary to assess the efficiency of our recov- ery algorithm in order to derive the number density of KBOs from the number of events in our survey. In order to measure our recovery efficiency we implant our data with synthetic occultations. The data are then repro- cessed in the same way we did to search for true events. By implanting into the actual lightcurves we do not make any assumption regarding the nature of the noise in our data. Note that our detection algorithm, described in Section 3.1, is not affected by the spectral characteristics of the noise, as long as the distribution of flux measure- ments in a lightcurve is stationary (Lehner et al. 2010). Our occultation simulator is based on the work described in Nihei et al. (2007). We first generate diffraction lightcurves for KBOs oc- culting point sources. We integrate the diffraction pat- tern over the disk of our target star. Keeping the stellar type fixed, the angular size is modulated by changing the apparent magnitude of the star and we can use the point source lightcurve to integrate the occultation sig- nature over the star disk. A point source lightcurve for a D = 3 km KBO at ∆ = 43 AU is shown in Figure 5, top, and the finite source lightcurve for a magnitude V = 11 star is shown in the second panel, left. Note the smooth- ing of the diffraction features. We modify the lightcurve to account for a finite impact parameter b by using the finite source lightcurve at impact parameter b = 0 as 8 Fig. 6. -- Implanted occultations recovered by our pipeline. On the left the lightcurves refer to star targets in a field observed near opposition and on the right in a field observed at large angle from opposition. deviation at that epoch from the local mean of the lightcurve in units of standard deviation as measured by telescope TAOS D. The value of hA(0) and hB(0), for TAOS A and B, are indicated by horizontal arrows labeled A and B, respectively. Each event is described in Table 3 Parameter of implanted events in Figure 6 TABLE 3 D (km) b (km) vrel (km/s) SNR θ⋆ (mas) hA(0) hB(0) hD(0) (a) (b) (c) (d) (e) (f) (g) (h) 0.7 1.0 3.0 8.0 0.7 1.0 3.0 8.0 0.50 0.50 1.19 2.7 0.22 0.94 1.18 2.93 25.4 14.9 25.4 25.4 8.2 7.7 3.1 8.1 39.8 29.4 12.8 9.4 10.2 10.8 12.9 15.7 0.03 0.03 0.03 0.006 0.03 0.03 0.006 0.04 -4.0 -5.5 -5.4 -6.1 -5.8 -10.0 -2.5 -11.8 -6.1 -6.6 -7.5 -12.3 -3.3 -8.7 -4.2 -13.7 -5.2 -7.4 -8.5 -7.9 -6.3 -8.6 -8.4 -11.7 an input and calculating the intensity of the occultation signal at the new distance of each point form the cen- ter of the diffraction pattern by interpolating points of the finite source lightcurve. A lightcurve for an F0V, V =11 star and a 3 km KBO occulting at an impact pa- rameter b = 2 km is plotted on the right hand side of the second panel of Figure 5. Finally, after calcu- lating the relative velocity of the KBO as a function of distance and angle from opposition as per Liang et al. (2004), and Nihei et al. (2007), we smooth the lightcurve to account for finite exposure intervals, and we sample the finite exposure lightcurve at the appropriate sam- pling rate. In this step we can account for dead-time in the sampling interval, which for TAOS is 47.5% at 5 Hz. We also allow an offset in time between the center of the finite sampled lightcurve and the integration bin. In the bottom row of Figure 5 the lightcurves for the event in row two are integrated over 105 ms second intervals and sampled at 5 Hz, the typical sampling rate of TAOS, for an event at opposition and with no time offset (left), and for an event at 50◦ from opposition with an offset of 50 ms between the center of the sampling interval and the center of the occultation (right). In order to sample properly the space of diameters to which the survey is sensitive we implant synthetic oc- cultations by objects of diameter D = 0.5, 0.7, 1.0, 1.3, 2.0, 3.0, 8.0, and 30.0 km. For a 30 km diameter KBO the event falls in the geometric regime, diffraction effects are therefore no longer significant and our efficiency sta- bilizes. Because our sensitivity decreases with decreas- ing diameter we implant progressively more objects at smaller diameters. The number of implantations at each size is designed to allow us to obtain a good sampling at all sizes. In Table 2 we report the number of objects implanted for each size in one of our efficiency runs, and the number of recoveries6. For objects within the Kuiper Belt (about 30 to 60 AU), the differences induced by dif- ferent distances are negligible in the occultation features as observed by TAOS. We therefore set the distance to ∆ = 43 AU. Every occultation event is implanted at a random epoch in the lightcurve set and at a random impact parameter between 0 and H/2, where we set H, 6 Four runs are conducted to improve statistical accuracy in the determination of our efficiency. 9 Fig. 7. -- Recovery efficiency for 3 km KBOs. Panel (a): efficiency as function of SNR. Efficiency versus crowdedness of the field in panel (b), defined as the number of targets brighter than RUSNO = 13.5. The efficiency is weighted by the crowdedness. Panel (c): efficiency versus magnitude MTAOS . In panel (e) the efficiency as a function of magnitude is weighted by the number of targets at that magnitude. Panel (d): efficiency versus relative velocity of the KBO targets. In panel (f) the efficiency versus relative velocity is weighted by the relative velocity. All error bars are calculated in a Poissonian fashion from the square root of the number of recoveries. a measure of the cross section of the event, to the size of the Airy ring and the projected size of the star in accordance to Nihei et al. (2007). In order to implant the synthetic occultations into our data we modulate the lightcurve by subtracting (adding) the amount of flux suppressed (augmented) by the occul- tation at each data-point as done in Z08. This approach slightly overestimates the noise due to Poisson statistics where the flux is suppressed, giving us a conservative estimate of our efficiency. We implant exactly one occul- tation in each lightcurve in our dataset. We then process our implanted lightcurves as we previ- ously did to search for events, namely we filter, rank the lightcurves, and evaluate the significance ξ of each point in each lightcurve. We then remove the false positives as described in Section 3.1. Our efficiency decreases rapidly with the KBO diameter: from nearly 33% at D = 30 km to 2 × 10−5 at D = 0.5 km. Note that at D = 30 km we ignore diffraction effects and the occultations are mod- eled to suppress the flux completely for several consec- utive points, depending on the relative velocity, but our efficiency is still significantly less than 100%. Some of our lightcurves are too noisy to allow detections. A set of synthetic events recovered by our pipeline in shown in Figure 6, and the parameters of each plotted event are given in Table 3. Various parameters affect our recovery efficiency. Our efficiency for the recovery of D = 3.0 km occulting KBOs is plotted in Figure 7. The efficiency as a function of SNR is plotted in Fig- ure 7a -- a few targets at SNR > 100 are left out of 10 Fig. 8. -- Effective solid angle for the first 2 years of TAOS data (Z08, empty squares) and for the current 3.75 year dataset (solid line and filled squares). the plot. The behavior of our efficiency as a function of crowdedness, where the crowdedness is defined as the number of targets in that TAOS field brighter than RUSNO = 13.5, is plotted in Figure 7b. The efficiency is here plotted multiplied by the number of targets in the field, to give a better idea of the implication of this parameter for the event recovery. The largest number of detections is achieved for more crowded fields. The efficiency decreases with magnitude (Figure 7c) Fig. 9. -- Model-independent upper limits from the TAOS survey (solid dots and solid line). Each point represents the upper limit to the number of KBO of that size or larger given the TAOS effec- tive coverage. Similarly derived upper limits from Bickerton et al. (2008) at 1 km (BKW), Wang et al. (2009) at 0.5 km (PS) and Bianco et al. (2009) at 1 km and 0.7 km (MMT) are plotted as empty circles. The upper limit from Bernstein et al. (2004, di- rect survey) is also plotted (HST). The dashed lines represent the 95% upper and lower limits from S09. The black square (FGS) represents the best fit density for the entire S09 survey, with 1σ error bars. The estimates of the number of objects in the Clas- sical Belt (CB), of Plutinos and Scattered Disk objects (SD) are plotted, as derived by Levison & Duncan (1997), Morbidelli (1997) and Volk & Malhotra (2008) respectively, assuming each family is the unique precursor of JFCs. by about a factor of five between magnitude 9 and 13. The dominant effect here is the decrease in SNR, though a competing effect occurs since lower magnitudes are associated with larger angular sizes, and our efficiency decreases with increasing angular size (Figure 4). Fur- thermore there are many more dim than bright stars in the sky: Figure 7e shows the efficiency as a function of MTAOS multiplied by the number of TAOS targets at that magnitude. The highest number of detections hap- pen for stars with MTAOS ∼ 12.5. Our efficiency as a function of the relative velocity of the KBO is plotted in Figure 7d. Observing at a pointing where the relative velocity of the KBOs is higher boosts the event rate of the survey. Our efficiency, however, is larger for smaller transiting velocities, particularly for small KBOs for which the time-line of the event is shorter than one of our data-points at opposition. Ultimately, the effective sky coverage of our survey depends linearly on both the efficiency and the velocity (see section 3.4). The efficiency multiplied by the relative velocity vrel is plotted against vrel in Figure 7f. Pointing near opposi- tion increases the effective coverage of our survey, and thus it increases our event rate, for 3 km KBOs. The survey strategy can be optimized at different sizes tak- ing into account the size dependent efficiency as well as the expected KBO size distribution. 3.4. Effective Sky Coverage and Upper Limits We calculate the effective sky coverage of our survey, Ωe, as: Ωe(D) = 1 w(D) X∗ H(D, θ∗) ∆ vrel ∆ E∗, (3) where E∗ is the exposure of the star target (the dura- tion of the lightcurve set), ∆ the distance to the oc- culter, H the cross section of the event (Section 3.3), w(D) the weight factor for that diameter, i.e. the frac- tion of lightcurves implanted with occultations by KBOs of diameter D, and the sum is carried out only over the lightcurves where events are recovered. The effective coverage of our survey, which takes into account our efficiency, is plotted in Figure 8, for both the dataset published in Z08 (empty squares) and for the current work (filled squares). The solid line is a spline fit to the points. 4. MODEL INDEPENDENT LIMITS ON THE SIZE DISTRIBUTION OF KBOS We can use Ωe to calculate model-independent upper limits: at each size for which our efficiency calculation was conducted we calculate the number density of KBOs from the TAOS dataset as a single-point upper limit. These limits are shown as filled dots. We can then inter- polate these upper limits with a spline fit, obtaining the solid line in Figure 9. At each diameter D this represents the maximum surface density of KBOs of diameter & D. Figure 9 shows model-independent upper limits re- ported by the occultation surveys of Bickerton et al. (2008), Bianco et al. (2009) and Wang et al. (2009). The improvement is of nearly an order of magnitude at 700 m and over an order of magnitude at 1 km from the results of Bianco et al. (2009) and at 500 m from Wang et al. (2009). As discussed in Sections 1 and 3, the TAOS sen- sitivity is limited to relatively bright stars (V . 13.5), 11 The best fit S09 model leads to an expected number of 2.1 detections in our TAOS data. This is a reasonable consistency. We can rule out slopes steeper than q = 4.0 for this model of the size distribution at 95% c.l. Figure 9 also shows the estimates of JFC progenitor populations. Our constraints on these families are dis- cussed in Section6. 5. OUTER SOLAR SYSTEM COLLISIONAL MODELS The collisional and dynamical evolution of the Solar System shaped the size distribution of the Kuiper Belt. The belt was originally populated by very small dust grains, with small orbital eccentricities (e ≤ 0.01) such as those we observe in circumstellar disks around other stars (Moro-Martin et al. 2007). Initially these small objects merge and grow (Kenyon & Luu 1999a): 1 km KBOs in the Kuiper belt are thus formed. As their gravitational cross section grows larger than their geo- metric cross-section, gravitational focusing speeds up the growth rate of the largest bodies. This phase is referred to as runaway growth, and objects as large as hundreds of kilometers can form. One such population, shaped primarily through agglomeration processes is predicted by theory to have a power law distribution in diameter dN/dD ∝ D−qL with power qL ≈ 4.5 (Kenyon & Luu 1999b). Direct observations of large KBOs confirm the power law behavior in this regime, the gravitationally -- dominated region of the size spectrum, with a best fit of qL = 4.8 (Fraser & Kavelaars 2009, Fuentes & Holman 2008 and reference therein). The size distribution of these large objects, for which gravity dominates the in- ternal strength, is remarkably insensitive to parameters such as Neptune stirring or the internal tensile strength of the KBOs. Meanwhile, very large objects in the planetary re- gion of the Solar System are also forming into plan- ets, that are believed to undergo significant migra- tions (Tsiganis et al. 2005, and references therein). The orbits of the planetesimals are then stirred up via gravita- tional interaction to velocities such that further impacts will result in the disruption of the smaller objects: this is the catastrophic collisions phase (Davis & Farinella 1997; Kenyon & Luu 1999a; Morbidelli et al. 2008); the time scale to reach this phase is estimated to be be- tween 10 Myr and 1 Gyr (Kenyon & Bromley 2001). For very small objects (probably tens of meters and smaller) the collisionally evolved population transitions to a regime where the KBO binding energy is dominated by internal strength, rather than gravity. Here the col- lisional cascade will generate a size distribution which follows a power law with index qS = 3.5 (Dohnanyi 1969; Kenyon & Luu 1999a), also in a fashion that is largely independent on the details of the evolution of the protoplanets. Note that the study of collisions be- tween icy bodies is still in its infancy, and future work in this field will permit assessing the behavior of colliding small strength-less or loosely bound particles (Leinhardt 2008). future work on coupling collisional and dynamical evolution codes, recently pioneered by Charnoz & Morbidelli (2007), should provide further in- sight in the behavior of the size distribution. Similarly, We will refer to the region in between these two regimes as the intermediate region. The extent of, and behavior of the size distribution in the intermediate region are Fig. 10. -- Expected number of events in 3.75 years of TAOS data from the size distribution presented in S09. Slopes of q = 3.9 (best fit to the HST/FGS data), q = 3.6 and 4.2 (±1σ from the best fit) are used. and TAOS has a relatively low efficiency at detecting D < 3 km KBOs. The large star exposure of TAOS, however, provides a dataset that well compensates for the lower sensitivity. Limits from the recent HST/FGS occultation sur- vey (Schlichting et al. 2009, hereinafter S09) are included in Figure 9. S09 reported the detection of one pos- sible event at a relatively high ecliptic latitude. The event is consistent with an occultation by a KBO with D = 1.0 km. The dashed lines in Figure 9 represent the 95% c.l. upper and lower limits from S09. The TAOS upper limits are more constraining than the correspond- ing limits from S09 for D & 0.6 km. The black square (FGS) represents the best fit density for the entire S09 survey, with 1σ error bars. Note that this is not a model- independent estimate, as it is derived assuming a straight power law distribution for small KBOs. Note also that TAOS has a greater sky coverage at D ∼ 1 km, where the S09 event was detected, by a factor ∼ 6. This detection reported by S09 is, however, not statistically inconsistent with our upper limit. Figure 10 shows the expected number of events for the TAOS survey if one assumes the size distributions de- rived in S09. S09 estimated a cumulative size distribu- tion ΣN (D > 0.5 km) = 2.1 × 107 deg−2 from their de- tection and non-detections, and fit a power-law to their data, anchored at ΣN (D > 90 km) = 5.4 deg−2, to yield a differential slope of q = 3.9 ± 0.3. Figure 10 shows the expected number of events for the TAOS survey for these distributions. We calculate the number of events expected in our survey as: Nexp = Z dN dD Ωe dD. With no detections our survey can rule out distributions that predict Nexp ≥ 3 at the 95% confidence level (c.l.). 12 Fig. 11. -- PS05 model and expected event rate for TAOS. Left. Triangles are the data from Fraser & Kavelaars (2009). Empty circles are the data from Bernstein et al. (2004). We use these data from direct imaging surveys to set the location of the large end size distribution. The model is parametrized with two slopes, qL = 5 and qI = 3. The positions of the break are plotted at Db =80, and 20 km: limiting values for the PS05 models. Right. Expected yield of events in 3.75 years of TAOS data. The horizontal dashed line represents the highest number of events allowed given no detections in the TAOS data (≤ 3 events at the 95% c.l.). Any model above this limit is ruled out by our survey. instead very sensitive to the formation and evolution pa- rameters, and observational information on this region can be compared to evolution models. In general, for a weaker KBO population the transition to the qS = 3.5 power law behavior will occur earlier, reducing the size of the transition region. A strong KBO population will display an extended intermediate region, generally show- ing here oscillations around a mean power law of slope which also depends on the details of the population and its evolution. In this section, we present four models, from literature, of the KBO size distribution dN/dD, and a simple parametric model. On the basis of these models we calculate the number Nexp of events expected to be detected by the TAOS survey. Since no events were found in our survey, any model which predicts Nexp ≥ 3 is ruled out by TAOS at the 95% c.l. We caution the reader that the predictions presented here depend on both the shape of the size distribution and on scaling parameters. All of our models are scaled so that the differential size distribution is consistent with the direct observations at D = 200 km, where the direct surveys are most constraining: dN/dD(D = 200 km) ∼ 0.04 deg−2km−1 (4) . For each model we will point out the cumulative number of KBOs predicted at D = 100 km, ΣN (D > 100 km), with this scaling choice. Strengths and weak- nesses of all of these models are also discussed in Fraser (2009). 5.1. Pan & Sari (2005) is constant through the collisional processes. They as- sume for most of their model calculation that the inter- nal strength of the objects is negligible (gravity domi- nated objects). This assumption is motivated by studies of comets and asteroids (PS05 and references therein). The transition to the fully strength-dominated regime, where the size distribution follows Dohnanyi (1969) with a power law with slope qS = 3.5, occurs at D ≤ 300 m. This region of the size spectrum is entirely below the sensitivity of TAOS. PS05 derive an analytical double power law size distri- bution for objects D ≥ 300 m: dN/dD ∝ d−qL f or D > Db, dN/dD ∝ d−qI f or D < Db. (5) This model is shown in Figure 11, left. The slope q has value qL = 5 for large objects and qI = 3 for objects in the intermediate region. PS05 are thus able to calculate self-consistently the location of the break in the power Db, which represents the size of the largest KBOs that experienced catastrophic collisions, as a function of time. The location of the break moves toward larger objects as the size distribution evolves. Model distributions with break points Db = 80, and 20 km, limiting values for PS05, and ΣN (D ≥ 100 km) ∼ 32 deg−2 are shown in Figure 11. The corresponding predicted number of events for the TAOS data analyzed in this paper is plotted in the right panel7. structure of the size distribution in the intermediate region is however generally more complicated. Between Pan & Sari (2005, hereinafter PS05) derived a fully an- alytical model for the size distribution of KBOs by as- suming the population is in a steady state and the mass 7 Note that a simple extension of the large-size power law distribution, which represents the naive expectation before Bernstein et al. (2004) appeared, would lead to thousands of de- tections. This was the initial design target of the TAOS project. 13 the evolution in the gravitationally-dominated regime. The models display a variety of behaviors for smaller objects. In all simulations a small dip in the 10-40 km region is predicted (cf Figures 10 and 11 in KB04). This is roughly consistent with the results of Bernstein et al. (2004): here collisions destroy weak KBOs and models with Neptune stirring or weakly bound KBOs produce a more significant dip. This feature is followed by an excess with respect to the nominal power law for 2-15 km KBOs. The amplitude of the excess varies substantially, between a factor of a few and a factor of a few tens, depending on the internal strength of the objects and on the details of the effect of Neptune stirring. The size distribution remains sensitive to the details of the models down to about D = 50 m, where once again a power law behavior begins, with power 3 < q < 5 in the strength-dominated regime. Figures 12 shows our representation of the models in KB04, as they are presented in Figure 10 and 11 of KB04. The size distribution is here shown scaled by the slope of the large size region (D−4.5). The shaded region repre- sents the range of the KB04 models and the plotted lines are some of our models, covering a large region of this model space. Focusing on the region near the transition between primordial and collisionally evolved population, an excess near 5 km and a depletion near 25 km are both visible and well represented in our models. We model the large size distribution as a power law with qL = 4.5 and the intermediate region with qI = 2. We model the excess near D = 5 km with a Gaussian, so that: dNex dD = dN dD (cid:18)1 + Iex exp − (D − µex)2 2.0 σe (cid:19) . (6) We fix the width of the Gaussian excess to σe = 3.5 km, we set the location to µex = 5.5 km or µex = 1.6 km. The intensity of the excess is determined by Iex; in order to fairly represent all results from the KB04 simulations we consider models with an excess of Iex = 0, 10 and 100. Note that as the break diameter moves towards large sizes the models naturally simulate the small dip near 20 km (Figures 12). Our results are not sensitive to the presence of this small depletion. With our scaling we obtain a cumulative surface density of KBOs N (D ≥ 100 km) ∼ 25 deg−2. Figure 13 shows the models with µex = 5.5 km (left) and the corresponding expected number of events for the TAOS dataset (right) for the limiting break positions D = 60 km and D = 10 km for Iex = 0, 10, and 100. For a given excess intensity, we can constrain the loca- tion of the break: in absence of the excess break diam- eters smaller than Db = 16.5 km are ruled out; break diameters Db < 32.6 km are ruled out for Iex = 10 and Db < 75.3 km are ruled out for an Iex = 100 excess. When moving the excess towards smaller sizes (µex = 1.6 km), break locations Db < 28.3 km are ruled out for Iex = 10, and Db < 63.3 km for Iex = 100. Since the KB04 simulations show that models with weaker KBOs and Neptune stirring produce a location of the break at smaller sizes, our result strongly favors mod- els that incorporate the effects of Neptune stirring and weaker KBOs, where the bulk strength Qb . 103 erg g−1 5.3. Benavidez & Campo Bagatin (2009) Fig. 12. -- Our modeling of the KB04 results. The range of results of the simulations of KB04 is represented by the shaded region (see Figure 10 and 11 in KB04). A few of our models are plotted as solid lines. All models are normalized by the slope of the large KBO size distribution D−4.5. Db and the second break, which is the region that TAOS can probe, a realistic size distribution is expected to have an oscillatory behavior which in the PS05 models pre- serves an average slope of qI = −3 (PS05). These waves however, as they are described in PS05, are small in am- plitude and affect our prediction to less than 1%. On the basis of our no-detection result, break diameters of Db < 51.3 km are excluded at 95% c.l. This is consistent with the data from direct observations and with the au- thors' interpretation of the model: in absence of stirring by Neptune, the location of the break is consistent with a KBO population comprised of objects with little internal strength. Note that direct surveys also suggest that for one such distribution the location of the break should be at a large diameters (see Fraser & Kavelaars (2009) and Figure 11, left). 5.2. Kenyon & Bromley (2004) Kenyon & Bromley (2004, hereinafter KB04) devel- oped numerical models of the collisional evolution of KBOs, including the effects of the internal strength and gravitational binding of KBOs, as well as the initial mass in the belt and a model of stirring by Neptune: if in its migration Neptune reaches its current location early its stirring effect will influence the shaping of the size distribution of the Kuiper Belt. The disruption energy, defined as the energy necessary to remove 50% of the combined mass of the colliding bodies, is modeled as Qd = Qbrβb + ρ Qgrβg , where Qb and βb describe the internal binding energy and Qg and βg the gravitational energy, with ρ the density and r the radius of the object. In the KB04 simulations Qb is varied between 10 and 108 erg g−1, ρ Qg between 10−4 and 104 and βg between 0.5 and 2.0, while βb is set to 0, as this parameter has little effect on the simulation results. All KB04 models generally agree in the shape and slope of the size distribution for large KBOs (D = 80 km and larger) generating a cumulative size distribution which follows N (D & 80 km) ∝ D3.5, equivalent to a power law differential size distribution with power qL ≈ 4.5, in good agreement with data and theoretical predictions for 14 Fig. 13. -- KB04 models. The left panel shows the differential size distribution, parametrized as two slopes (qL = 4.5 and qI = 2) and on the right side are the corresponding event rates for the TAOS survey. Symbols are the same as in Figure 11. For limiting break positions of Db = 60 and 10 km the distribution is plotted for a model with no Gaussian excess, and models with Gaussian excess of intensity Iex = 10 and Iexp = 100, centered at µex = 5.5 km. Fig. 14. -- Our rendering of BCB09 models: the models differ in the prescription for fragmentation and all models are parametrized as a series of three slopes. Symbols are the same as in Figure 11. The first break point is fixed at 100 km. The slope for the smallest size objects is set to 3.7. The location of he second break and intermediate slope are 5 km and 1.0, 3.6 km and 2.0, 4.6 km and 2.5 and 0.36 km and 3.0. Right: corresponding number of events in our TAOS survey. Recent simulations by Benavidez & Campo Bagatin (2009, hereinafter BCB09) divide the Kuiper Belt into three dynamical families -- the CB, the Plutinos (Reso- nant Population) and the SD -- and follow the collisional evolution of each, while taking into account the physics of the fragmentation of icy and rocky bodies at the typical relative velocities of KBOs. This suite of models ignores the effects of Neptune stirring. The models incorpo- rate four scaling laws for fragmentation: a simple scaling driven by gravitational self-compression, two scaling laws which include both self-compression and the effects of strain-rate, as described by Farinella et al. (1982), with different diameter dependency (D−0.25 and D−0.5) and one that follows the modeling of Benz & Asphaug (1999) 15 Fig. 15. -- Left: Our modeling of the F09 models. The models are reproduced on the basis of Figure 2 in F09, and divots and excesses are generated with Gaussians. Symbols are the same as in Figure 11. The slope on the large hand side (D & 65 km) is qL = 4.8. Right: corresponding event rates for our survey. for icy bodies. They then vary the material strength of the objects between 105 and 107erg cm−3. In BCB09 the size distribution of large objects is set to a power law with slope qL ≈ 4.0. This is set as an initial condition to their simulations, and since the large objects are not undergoing many collisions, this shape is preserved. A first break is seen around 100 km. The size distribution then departs from power law be- havior for all parameter choices, with two to four orders of magnitudes fewer KBOs in the ∼ 1 km region than the nominal power law would predict. The size distribution then follows again a power law, with slope qS = 3.5, in the strength-dominated regime. The BCB09 models are shown in Figure 14, left. We simplified the size distri- bution behavior in the intermediate region with a single power law, though typically the behavior is oscillatory. Our scaling leads to N (D ≥ 100 km) ∼ 21 deg−2 for the BCB09 models. These models are all allowed by the TAOS data, partly because of the location of the initial break at Db = 100 and the slope of the large end size dis- tribution. The break location is at the large end of what is allowed by direct observations (indeed outside of the Fraser & Kavelaars (2008) allowed range of Db ∈ [50, 95 km] with assumption of a 6% albedo). The lo- cation of this first break does not evolve in the BCB09 simulations from what is set as an initial condition. Sim- ilarly, the choice of a slope at the large end of the size distribution of qL = 4.0 is slightly shallower than the cur- rent best fit value from Fraser & Kavelaars (2009) and Fuentes & Holman (2008). As the parameters relative to the large size end of the distribution are more firmly pinned down by direct observations, these models and future occultation data may place stronger constraints on the details of the shape of the size distribution below the first break and thus the details of the fragmenta- tion mechanisms, providing information on the internal structure of the KBOs. 5.4. Fraser (2009) Fraser (2009, hereinafter F09) considers the collisional evolution of the Kuiper Belt size distribution after the epoch of accretion. Starting with initial conditions that reproduce the observed large end size distribution (qL = 4.8), which will not further evolve in the colli- sional simulations, and Db1 = 2 km, the population is collisionally evolved over the age of the Solar Sys- tem. A depletion, or divot, forms at D ∼ 10 × Db1, or D ∼ 20 km. The size of this depletion changes when changing parameters relative to the internal strength of the KBOs or the impact velocity, as well as choices for the initial intermediate slope. An excess is also evident at smaller sizes: D ∼ 4 km. We model the F09 size distributions as they appear in Figure 2 of F09. Our size distributions are created as a sequence of 3 slopes: qL = 4.8, 1 ≤ qI ≤ 3, and 2.5 ≤ qs ≤ 4.5, and we model the divots and excesses with Gaussians (Figures 15, left). Here we scale the differential size distribution so that dN/dD(D = 200 km) ∼ 0.047 deg−1km−1, to repro- duce the number of objects at D = 50 km8. Our survey cannot constrain these models: the pronounced divot at D ∼ 20 km causes a very low event rate in our survey (Figure 15, right). 5.5. Generic 3 -- Regime Model: Constraints on the Intermediate Region of the Size Spectrum A simple, generic 3 -- regime model allows us to describe separately the primordial region, the intermediate re- gion and the fully collisional, strength-dominated regime. Knowing that both the large (gravitationally-dominated) 8 N (D = 50 km) ∼ 15, Fraser, private communication. 16 and the small (strength-dominated) regions of the size spectrum are relatively insensitive to the details of the in- ternal structure and evolution of the Kuiper Belt, it is the intermediate region that contains the most information about the physical details of the KBOs. We grossly sim- plify the expected structure in the intermediate regime and describe it with a power law. The parameters of one such models would then be three slopes qL, qI, and qS, and two break locations Db1 and Db2. We set the slope of the large end of the size spec- trum and the location of the first break to the best fit to the data from direct observations: qL = 4.8 and Db1 = 75 km. We model the intermediate region as a plateau, or with a shallow slope: qI ∈ [0, 3], and the small size end of the spectrum as a power law with slope 3.5 for the strength-dominated, collisionally evolved popula- tion. Figure 16 shows our 3-regime model (left) and the limit we can set to the intermediate slope, qI, and sec- ond break location, Db2, phase space (right). Any pair of values qI and Db2 that fall above the solid line are excluded. 6. THE JUPITER FAMILY COMETS PROGENITOR POPULATION The Jupiter Family Comets (JFCs) are believed to originate in the Outer Solar System. In this scenario the giant planets generate gravitational perturbations that affect the orbits of the Outer Solar System bod- ies, injecting them into the planetary region, where they are captured by Jupiter. The orbital inclination of the JFCs suggests that their precursor population has a disk-like distribution, favoring thus the Kuiper Belt over the Oort Cloud as a reservoir (Volk & Malhotra 2008, and references therein). The Classical Belt (CB), the Plutinos and the Scattered Disk (SD) have been con- sidered as precursors in various studies. The dynam- ical characteristics of each population determines the efficiency of the injection process and the number of objects in each progenitor family can thus be derived on the basis of the density of JFCs, which is observa- tionally constrained (see Tancredi et al. 2006, and refer- ences therein). Furthermore the size distribution of JFCs should reflect the size distribution of the progenitor pop- ulation. Bernstein et al. (2004) and Bianco et al. (2009) have pointed out that a better determination of the size distribution of the Kuiper Belt would help understand the origin of the JFCs. We assume the JFC precursors are in the size range 1 − 10 km as this is observed to be the typical size of JFCs (Lowry et al. 2008). We considered the esti- mates on the KBO populations (CB and Plutinos) and SD derived from dynamical simulations under the as- sumption that each population is the unique progeni- tor of the JFCs, and we compared them to the upper limits derived from our survey (Figure 9). We use the estimate of Levison & Duncan (1997) for a population of cometary precursors entirely in the CB, and that of Morbidelli (1997) for Plutino progenitors. These are converted into a surface density by assuming for each population a projected sky area of 104 deg2, as was done by Bernstein et al. (2004). We consider the results of Volk & Malhotra (2008) for a progenitor population in the SD . We calculate the minimum surface density of SD objects expected in the region of sky typically ob- served by the TAOS survey. For this we use information on the fraction of time the objects spend between 30 and 50 AU and within 3◦ of the ecliptic plane as provided by Volk & Malhotra (2008). These estimates on the number of objects are shown in Figure 9 as horizontal lines. Our results rule out a precursor family composed uniquely of CB objects with D & 3 km. Occultation surveys are the only surveys that at present can probe this region of the size spectrum, and our preliminary re- sult shows that future occultation surveys will be able to derive useful constraints on the origin of the JFCs. 7. CONCLUSIONS We presented an analysis of 3.75 years of TAOS data, comprising 5 × 105 star -- hours of lightcurves sampled at 4 or 5 Hz observed with three telescopes simultaneously. We searched for occultations of our target stars by KBOs in order to constrain the size distribution of KBOs, par- ticularly in the 500 m to 10 km region, which is currently out of reach of direct observation surveys. More than 90% of the TAOS data is collected within 5◦ of ecliptic latitude in order to maximize the occultation rate by ob- jects in the Kuiper Belt. Occultations near opposition lead to a higher event rate for TAOS, even after taking into account the increased recovery efficiency for small objects where the angle from opposition is larger and the relative velocity of the KBOs is lower. We found no occultation events in our data. This al- lowed us to set upper limits to the number density of KBOs that are stringent enough to be compared use- fully with models for the formation and evolution of the Kuiper Belt. We considered four theoretical mod- els, PS05, KB04, BCB09 and F09, all of which describe the present size distribution of the Kuiper Belt, and we set constraints on these models. This is the first detailed comparison of occultation data with specific model re- sults. Our result, particularly when compared with PS05 and KB04, suggests that the Kuiper Belt is populated by fragile bodies, and that the effect of the migration of Neptune played an important role in its formation. None of the BCB09 or F09 models can be ruled out. Using a generic model, where the size distribution is described by three consecutive power laws, and fixing the slope on the large end size and the location of the first break to the best fit from direct observations (qL = 4.8 km, Db1 = 75 km), and the slope of the small side to qs = 4.0 we can constrain the intermediate slope qI and the location of the second break Db2, as shown in Figure 16. As direct surveys are currently not sensitive to KBOs smaller than D ∼ 28 km occultation surveys provide the only probe of this region of the size spectrum, and the large TAOS dataset allowed us to set the first constraints to the location of the second break. We also considered the Jupiter Family Comets. As- suming the JFCs are injected into their present or- bit from one of the Kuiper Belt populations, Classical Belt, Plutinos, or from the Scattered Disk, we compared the upper limit derived from our survey to the esti- mates of the number of objects derived using the num- ber of JFCs by Levison & Duncan (1997) for a popula- tion of cometary precursors entirely in the CB, that of Morbidelli (1997) for Plutinos and of Volk & Malhotra (2008) for a progenitor population in the SD. We can 17 Fig. 16. -- Three-slope model described in Section 5.5 (left) for the minimum second break location Db2 allowed by slopes qI = 0 and 3. Slope -- break location phase space for the three-slopes model (right). The shaded region is excluded by our data. rule out the a unique precursor family composed of CB objects D & 3 km. This preliminary result confirms that occultation surveys can help understanding the origin of JFCs. A recent analysis of the HST guiding data (S09) re- ports the detection of an occultation by a D ∼ 1 km KBO. This object is detected at high ecliptic latiutde (b ∼ 14◦). The surface density derived in S09, assuming a straight power law size distribution for small KBOs, is within the upper limit set by TAOS, although a 1σ increment over the best fit surface density is ruled out to better than 95% c.l. TAOS has operated for over four years observing con- tinuously with three, and now four, 50 cm aperture tele- scopes. TAOS is only marginally sensitive to sub-km KBO occultations, but we were able to prove that the low sensitivity at sub-km sizes is more than compen- sated for by the vast exposure of which TAOS is capable. Improvements in our selection criteria and the analysis of four telescope data should increase our sensitivity to smaller objects. The authors wish to thank Scott Kenyon, for insight- ful conversations. Work at the CfA was supported in part by the NSF under grant AST-0501681 and by NASA under grant NNG04G113G. Work at NCU was supported by the grant NSC 96-2112-M-008-024-MY3. Work at ASIAA was supported in part by the the- matic research program AS-88-TP-A02. Work at Yon- sei was supported by National Research Foundation of Korea through Grant 2009-0075376. Space Science In- stitute. The work of N. Coehlo was supported in part by NSF grant DMS-0636667. Work at LLNL was per- formed in part under USDOE Contract W-7405-Eng-48 and Contract DE-AC52-07NA27344. Work at SLAC was performed under USDOE contract DE-AC02-76SF00515. Work at NASA Ames was supported by NASA's Plane- tary Geology & Geophysics Program. REFERENCES Asphaug, E. & Benz, W. 1996, Icarus, 121, 225 Benavidez, P. G. & Campo Bagatin, A. 2009, Planet. Space Sci., 57, 201 Benz, W. & Asphaug, E. 1999, Icarus, 142, 5 Bernstein, G. M. et al. 2004, AJ, 128, 1364 Bertin, E. & Arnouts, S. 1996, A&AS, 117, 393 Bianco, F. B. et al. 2009, AJ, 138, 568 Bickerton, S. J., Kavelaars, J. J., & Welch, D. L. 2008, AJ, 135, 1039 Bickerton, S. J., Welch, D. L., & Kavelaars, J. J. 2009, AJ, 137, 4270 Born, M. & Wolf, E. 1980, Principles of optics. Electromagnetic light theory of propagation, (Oxford: Pergamon Press, 1980, 6th corrected ed.) interference and diffraction of Chang, H.-K., et al. 2007, MNRAS, 378, 1287 Charnoz, S. & Morbidelli, A. 2007, Icarus, 188, 468 Cutri, R. M. et al. 2003, 2MASS All Sky Catalog of point sources. (IPAC, Pasadena) Davis, D. R. & Farinella, P. 1997, Icarus, 125, 50 Dohnanyi, J. W. 1969, J. Geophys. Res., 74, 2531 Elliot, J. L. et al. 2005, AJ, 129, 1117 Farinella, P., Paolicchi, P., & Zappala, V. 1982, Icarus, 52, 409 Fraser, W. C. & Kavelaars, J. J. 2008, Icarus, 198, 452 -- . 2009, AJ, 137, 72 Fraser, W. C. et al. 2008, Icarus, 195, 827 Fraser, W. 2009, ApJ, 706, 119 Fuentes, C. I., George, M. R., & Holman, M. J. 2009, ApJ, 696, 91 Fuentes, C. I. & Holman, M. J. 2008, AJ, 136, 83 Kenyon, S. J. & Bromley, B. C. 2001, AJ, 121, 538 -- . 2004, AJ, 128, 1916 Kenyon, S. J. 2008, Formation and Collisional Evolution of Kuiper Belt Objects, ed. M. A. Barucci, H. Boehnhardt, D. P. Cruikshank, & A. Morbidelli, 293 -- 313 Kenyon, S. J. & Luu, J. X. 1999a, AJ, 118, 1101 -- . 1999b, ApJ, 526, 465 Kenyon, S. J. & Windhorst, R. A. 2001, ApJ, 547, L69 18 Lehner, M. J. et al. 2009, PASP, 121, 138 -- . 2010, in preparation Leinhardt, Z. M. 2008, Physical Effects of Collisions in the Kuiper Belt (Barucci, M. A. and Boehnhardt, H. and Cruikshank, D. P. and Morbidelli, A.), 195 -- 211 Levison, H. F. & Duncan, M. J. 1997, Icarus, 127, 13 Liang, C. L. et al. 2004, Statistical Science, vol. 19, 265-274 Liu, C.-Y. et al. 2008, MNRAS, 388, L44 Lowry, S. 2008, Kuiper Belt Objects in the Planetary Region: The Jupiter-Family Comets (The Solar System Beyond Neptune), 397 -- 410 Mink, D. 2006, in ASP Conf. Ser. 351: Astronomical Data Analysis Software and Systems XV, ed. C. Gabriel, C. Arviset, D. Ponz, & S. Enrique, 204 -- + Monet, D. G. et al. 2003, AJ, 125, 984 Morbidelli, A. 1997, Icarus, 127, 1 Morbidelli, A., Levison, H. F., & Gomes, R. 2008, The Dynamical Structure of the Kuiper Belt and Its Primordial Origin, ed. M. A. Barucci, H. Boehnhardt, D. P. Cruikshank, & A. Morbidelli, 275 -- 292 Moro-Martin, A. et al., 2008, Extra Solar Kuiper Belt Dust Disks (The Solar System Beyond Neptune), 465 -- 480 Nihei, T. C. et al. 2007, AJ, 134, 1596 Nordgren, T. E. et al., P. 2002, AJ, 123, 3380 Pan, M. & Sari, R. 2005, Icarus, 173, 342 Rice, J. A. 2006, Mathematical Statistics and Data Analysis, Second Edition (Duxbury Press) Roques, F. & Moncuquet, M. 2000, Icarus, 147, 530 Roques, F., Moncuquet, M., & Sicardy, B. 1987, AJ, 93, 1549 Roques, F. et al. 2006, AJ, 132, 819 Schenk, P. M. & Zahnle, K. 2007, Icarus, 192, 135 Schlichting, H. E. et al. 2005, Nature, 462, 895 Stern, S. A. 1996, AJ, 112, 1203 Tancredi, G. et al. 2006, Icarus, 182, 527 Tsiganis, K. et al. 2005, Nature, 435, 459 van Belle, G. T. 1999, PASP, 111, 1515 Volk, K. & Malhotra, R. 2008, ApJ, 687, 714 Wang, J.-H.,et al. 2009, ArXiv e-prints 0910.5598 Zhang, Z.-W. et al. 2008, ApJ, 685, L157 Zhang, Z. W. et al. 2009, PASP, 121, 1429
1205.6890
1
1205
2012-05-31T05:16:10
Searching for young Jupiter analogs around AP Col: L-band high-contrast imaging of the closest pre-main sequence star
[ "astro-ph.EP", "astro-ph.GA" ]
The nearby M-dwarf AP Col was recently identified by Riedel et al. 2011 as a pre-main sequence star (age 12 - 50 Myr) situated only 8.4 pc from the Sun. The combination of its youth, distance, and intrinsically low luminosity make it an ideal target to search for extrasolar planets using direct imaging. We report deep adaptive optics observations of AP Col taken with VLT/NACO and Keck/NIRC2 in the L-band. Using aggressive speckle suppression and background subtraction techniques, we are able to rule out companions with mass m >= 0.5 - 1M_Jup for projected separations a>4.5 AU, and m >= 2 M_Jup for projected separations as small as 3 AU, assuming an age of 40 Myr using the COND theoretical evolutionary models. Using a different set of models the mass limits increase by a factor of ~2. The observations presented here are the deepest mass-sensitivity limits yet achieved within 20 AU on a star with direct imaging. While Doppler radial velocity surveys have shown that Jovian bodies with close-in orbits are rare around M-dwarfs, gravitational microlensing studies predict that ~17% of these stars host massive planets with orbital separations of 1-10 AU. Sensitive high-contrast imaging observations, like those presented here, will help to validate results from complementary detection techniques by determining the frequency of gas giant planets on wide orbits around M-dwarfs.
astro-ph.EP
astro-ph
To appear in ApJ Preprint typeset using LATEX style emulateapj v. 12/14/05 AP Col SEARCHING FOR YOUNG JUPITER ANALOGS AROUND AP COL: L-BAND HIGH-CONTRAST IMAGING OF THE CLOSEST PRE-MAIN-SEQUENCE STAR Sascha P. Quanz1,2, Justin R. Crepp3, Markus Janson4, Henning Avenhaus2, Michael R. Meyer2, and Lynne A. Hillenbrand3 To appear in ApJ ABSTRACT The nearby M-dwarf AP Col was recently identified by Riedel et al. 2011 as a pre-main-sequence star (age 12 – 50 Myr) situated only 8.4 pc from the Sun. The combination of its youth, distance, and intrinsically low luminosity make it an ideal target to search for extrasolar planets using direct imaging. We report deep adaptive optics observations of AP Col taken with VLT/NACO and Keck/NIRC2 in the L-band. Using aggressive speckle suppression and background subtraction techniques, we are able to rule out companions with mass m ≥ 0.5 – 1MJup for projected separations a > 4.5 AU, and m ≥ 2 MJup for projected separations as small as 3 AU, assuming an age of 40 Myr using the COND theoretical evolutionary models. Using a different set of models the mass limits increase by a factor of &2. The observations presented here are the deepest mass-sensitivity limits yet achieved within 20 AU on a star with direct imaging. While Doppler radial velocity surveys have shown that Jovian bodies with close-in orbits are rare around M-dwarfs, gravitational microlensing studies predict that 17+6 −9% of these stars host massive planets with orbital separations of 1-10 AU. Sensitive high-contrast imaging observations, like those presented here, will help to validate results from complementary detection techniques by determining the frequency of gas giant planets on wide orbits around M-dwarfs. Subject headings: stars: formation - planets and satellites: formation - planets and satellites: detection - stars: individual (AP Col) 1. INTRODUCTION The high-contrast imaging technique has provided some remarkable extrasolar planet discoveries over the past few years (e.g., Chauvin et al. 2005; Marois et al. 2008, 2010; Lagrange et al. 2010). Direct detection allows us not only to infer orbital separations and eventually determine orbits (Crepp et al. 2011), but also to study extrasolar planets in detail through multi-color photom- etry and spectroscopy (Mohanty et al. 2007; Quanz et al. 2010; Hinz et al. 2010; Janson et al. 2010; Bowler et al. 2010). Investigation of the small sample of directly im- aged planets known to date has already revealed unex- pected results, including the suggested presence of thick cloud layers (Skemer et al. 2011), enhanced metallicity, and indications of non-equilibrium chemistry (Barman et al. 2011). Such observations present a challenge to theoretical atmospheric models (Skemer et al. 2012). Dedicated imaging surveys have now searched host stars of all spectral-types, with an initial emphasis on FGK stars (Masciadri et al. 2005; Biller et al. 2007; Kasper et al. 2007; Lafreni`ere et al. 2007a; Chauvin et al. 2010; Heinze et al. 2010). The fact that the first hand- ful of directly imaged planets have been found orbiting A-stars provides an unambiguous clue that the efficiency of the planet formation process depends upon host star Electronic address: [email protected] 1 Based on observations collected at the European Organisation for Astronomical Research in the Southern Hemisphere, Chile, un- der program number 288.C-5005(A). 2 Institute for Astronomy, ETH Zurich, Wolfgang-Pauli-Strasse 27, 8093 Zurich, Switzerland 3 Department of Astrophysics, California Institute of Technol- ogy, 1200 E. California Blvd., Pasadena, CA 91125 4 Department of Astrophysical Sciences, Princeton University, Princeton, NJ, USA mass. Indeed, extrapolation of the Doppler radial veloc- ity planet population to separations accessible to high- contrast imaging instruments predicts that A-type stars are ideal targets, even though they are intrinsically bright (Crepp & Johnson 2011). Massive stars have a high gas giant planet occurrence rate, and a propensity to form more massive planets with long orbital periods (John- son et al. 2007). While the same line of reasoning sug- gests a relative paucity of giant planets orbiting M-stars (Kennedy & Kenyon 2008), initial results from gravita- tional microlensing surveys suggest a significantly higher frequency of planets located beyond the snow-line (Gould et al. 2010; Cassan et al. 2012) compared to rates of close- in planets inferred from Doppler measurements (Johnson et al. 2010; Bonfils et al. 2011). High-contrast imaging observations will soon be able to rectify any discrepancies found by indirect planet de- tection techniques. Planned instruments and new adap- tive optics (AO) systems at the VLT (Beuzit et al. 2006), Gemini South (Macintosh et al. 2007), Palomar (Hinkley et al. 2011), the LBT (Esposito et al. 2012), and Subaru (Tamura 2009) will generate unprecedented contrast lev- els, providing direct access to young planets with mass m ≈ 0.5MJup. It is imperative to reach this level of sensi- tivity because the "bottom-to-top" formation paradigm of core-accretion (Pollack et al. 1996) predicts an overall planet occurrence rate that rises steadily with decreasing planet mass. In other words, access to sub-Jovian-mass bodies significantly enhances the prospects for detecting an extrasolar planet around any type of star. Youth and distance are important parameters to con- sider when prioritizing high-contrast imaging targets. With an accurately measured distance of only 8.39±0.07 pc and estimated age between 12 – 50 Myr, the nearby 2 Quanz et al. M4.5 flare star AP Col (RA: 06h04m52s.16; DEC: −34◦33′36′′.0 (J2000, E2000)) has recently been iden- tified as the closest pre-main sequence star to the Sun (Riedel et al. 2011). Based on its apparent age and space motion, AP Col is most likely a member of the ∼40 Myr old Argus/IC 2391 Association (Riedel et al. 2011). Stel- lar companions have been ruled out by lucky imaging ob- servations prior to establishing its proximity and youth (Bergfors et al. 2010). Given its (extremely) convenient properties, AP Col is an ideal target to search for gas giant planets by means of direct imaging. In this paper, we present the results from two different sets of deep high-contrast imaging observations taken at the VLT and Keck. The observations were carried out in the L-band which, compared to shorter near-infrared (NIR) wavelengths, provides a higher Strehl ratio and more favorable star-planet flux ratio, while also main- taining manageable thermal background levels from the sky and instrument optics. Despite generating the deep- est direct imaging mass-sensitivity yet achieved for or- bital separations on the scale of the solar-system (a > 3 AU), we find no evidence for the existence of gas giant planets. We discuss the implications of these observa- tions to the extent that they support the notion that Jovian bodies appear to be rare around M-dwarfs. 2. OBSERVATIONS AND DATA REDUCTION 2.1. VLT/NACO Imaging data was acquired on December 4, 2011 UT, with the AO-fed, high spatial resolution camera NACO mounted on ESO's VLT/UT4 8.2-m telescope at Paranal (Lenzen et al. 2003; Rousset et al. 2003). All images were unocculted and taken with the L27 camera (plate scale ∼ 27.15 mas pixel−1) using the L′ filter (λc = 3.8 µm, ∆λ = 0.62 µm). We used pupil stabilization mode to enable angular differential imaging (ADI) (Marois et al. 2006). To enhance the signal-to-noise ratio (S/N) of potential companions, we chose to moderately saturate the stel- lar PSF core. Unsaturated images were also acquired to calibrate the photometry. Detector reads were recorded individually ("cube mode"). The integration time was set to 0.2s per read (unsaturated reads had an expo- sure time of 0.1s). We obtained 134 raw data cubes for AP Col in total. Each cube consisted of 64 reads. Be- tween cubes we moved the telescope by ≈9′′ on the sky executing a 5–point dither pattern to facilitate removal of background noise from the sky and instrument op- tics. Individual detector reads were checked for open loops and poor AO correction. We discarded one com- plete cube as well as several individual reads from other cubes. The remaining reads were median combined in stacks of 32, yielding 265 science images with an effective integration time of 32×0.2 s each. We measure the PSF FWHM in unsaturated images to be ∼4.3 px (≈ 0.12′′), which is comparable to the theoretical diffraction limit, Θ ≈ λc/D ≈ 0.10′′. Our basic data reduction steps (bad pixel correction, sky subtraction) are described in Quanz et al. (2010, 2011). We use the LOCI algorithm to subtract the stellar PSF (Lafreni`ere et al. 2007b). Following the same naming convention as in the original paper, we set pertinent reduction parameters to the following values: FWHM=4.5 px, Nδ=0.75, dr=5, NA=300. 2.2. Keck/NIRC2 data We also observed AP Col from Keck Observatory on January 7, 2012 UT. Using the 10-m Keck II telescope and AO system (Wizinowich et al. 2004), we acquired im- ages with the Near InfraRed Camera (NIRC2; PI: Keith Matthews) in the L′ filter (λc = 3.8µm, ∆λ = 0.7 µm). A total of 67 images were obtained, each consisting of 300 reads with 0.106s integration time per read. All im- ages were in the linear detector regime and recorded with the narrow camera setting which provides a plate scale of 9.963 mas pixel−1 (Ghez et al. 2008). Like the VLT data, we did not use a coronagraph. Vertical angle mode was used to allow for ADI operation. The AO system was running at a refresh rate of 438 Hz. The FWHM of the PSF was 9.1 px (≈ 0.09′′), i.e., close to the theoretical diffraction limit of Θ ≈ 0.08′′. Raw images were processed using standard techniques to remove dark current, flat-field, replace hot-pixel val- ues, and align the target star. In the same manner as described in the previous section with the VLT data, we used the LOCI algorithm on the 67 images to subtract the stellar PSF. The only difference was the value for the FWHM which was set to 9 px for the Keck data. Table 1 shows a summary of relevant observing parameters for both the VLT and the Keck datasets. 3. RESULTS We find no evidence for point sources in either the VLT or Keck data sets, in raw or processed images. Our checks for real companions did initially show potential candidates in the VLT data. However, they subsequently disappeared when combining only a fraction of the final images, and other bright speckles showed up instead. 3.1. Derivation of contrast curve A comparison of the 5-σ contrast levels obtained with NACO and NIRC2 is shown Figure 1. Contrast is defined as: ∆L′(r) = 2.5 · log10 (1) F∗ 5σ(r)/qπr2 ap where F∗ is the mean stellar flux per pixel in an aper- ture with radius rap = 3 pixels (VLT/NACO) or rap = 5 pixels (Keck/NIRC2), and σ is the standard deviation of the pixels within centro-symmetric annuli centered on the star, which are twice as wide as the aperture radius. The VLT/NACO observations provide slightly deeper contrast compared to Keck as a result of lower background levels, more field rotation, and lower airmass. We focus on these data for the remainder of the analysis. To derive our final detection limits, we insert fake plan- ets of known brightness into individual NACO frames. Two planets of the same brightness are inserted at the same radial distance (0.3′′, 0.4′′, 0.5′′, 0.6′′, 0.8′′, 1.0′′, 1.5′′, and 2.0′′), but at opposing sides from the central star. The signal-to-noise of the planets in the final LOCI processed image is computed via: (cid:16) S N (cid:17)Planet = FPlanet σ/qπr2 ap , (2) where FPlanet is the measured planet flux in an aperture with radius rap = 3 pixels and σ is the standard deviation Searching for young Jupiter Analogs around AP Col 3 of the pixels within an 7 pixel-wide arc between the two fake planets and centered at the same radial distance. By inserting two planets simultaneously and having two arcs for the noise estimate between the planets, we ob- tain 4 different values for (cid:0) S N (cid:1)Planet. To be conservative, we chose the combination yielding the lowest value as our reference. For a given radial separation we lowered the brightness of the planets in steps of 0.5 magnitudes until (cid:0) S N (cid:1)Planet < 5. The signal-to-noise of the next brightest planet was then used for calibration and the values ex- trapolated to a level that corresponds to (cid:0) S N (cid:1)Planet = 5. These values are shown in Figure 2. We emphasize that injecting fake planets of various brightness and retrieving them is the only way to ob- tain an accurate estimate for the achieved detection limit. LOCI removes flux from potential companions (Lafreni`ere et al. 2007b) as a function of separation from the star, LOCI input parameters, and brightness of the companions themselves (Pueyo et al. 2012). There are no analytical methods for predicting and correcting for flux losses. The limiting magnitude in the VLT/NACO data ap- proaches L′ limit ≈ 16.7 mag at a separation of ≈ 0.7", be- yond which the sensitivity becomes flat indicating that we have reached the floor set by thermal background noise (Fig. 2). Sensitivity is dominated by subtraction residuals closer to the star where the detection limits are governed more-so by residual scattered light. 3.2. Detection limits using COND models To convert our on-sky contrast into mass-sensitivity, we must first estimate the L′ magnitude of AP Col. Us- ing color transformations provided by the 2MASS team5, we estimate AP Col's K magnitude to be K = 6.88±0.05 mag, based on the apparent Ks value of Ks = 6.87 ± 0.02 mag (Cutri et al. 2003). A typical M4.5 star has an ap- proximate K − L′ color of 0.4 mag (Cox 2000). Thus, AP Col has an L′ magnitude of L′ ≈ 6.48 ± 0.05 mag. Figure 2 (left-hand panel) compares our detection lim- its with the predicted L′ magnitudes for 40 Myr objects having masses of 0.5, 1 and 2 MJup based on the COND models (Baraffe et al. 2003). We select this specific age because Riedel et al. (2011) derive a most likely age be- tween 12 – 50 Myr for AP Col, and membership with the ∼40 Myr old Argus/IC 2391 association seems plausible from AP Col's space motion. Our data are sufficiently sensitive to detect objects with masses between 0.5 – 1 MJup (or greater) for separations ≥4.5 AU. Inward of 4.5 AU, as close as ∼3 AU, objects with masses between 1 – 2 MJup (or greater) are detectable. Based on these model predictions, this is the first time that a young Jupiter analog could have been detected by means of direct imag- ing around another star. This is also demonstrated in Figure 3, where we have inserted two fake planets in the raw images at a projected separation of 5.2 AU and re- run the data reduction. The planets' brightness corre- sponds to a 1 MJup object with an age of 40 Myr based on the COND models. It clearly shows that we would have detected such an object in our data. To assess the likelihood of imaging a planet during a single epoch observation, the following relation can be 5 http://www.astro.caltech.edu/ jmc/2mass/v3/transformations/ used , Pdetect(a) = cos(arcsin(IWA/a)) (3) where Pdetect is the detection probability, a the semi- major axis of the planet in AU, and IWA is the inner working angle (in AU) where the contrast curves allows for the detection of this planet. This relation is only valid for circular orbits and assumes uniform sensitivity in azimuth, but takes into account all possible orbital inclinations. For eccentric orbits Pdetect would be higher for any given a, as the planet would spend more time close to apastron where it is easier to detect. In Figure 4 we plot the detection probability for a 1 MJup planet and a 2 MJup planet using IWA = 4.5 AU for the 1 MJup case and IWA = 3 AU for the 2 MJup case (see, lefthand plot in Figure 2). It is clear that we had a ∼50% chance to detect (at greater than 5-σ) a young Jupiter analog at 5.2 AU, assuming one is present. The probability rises quickly for more massive or more distant planets. 3.3. Comparison of theoretical models In the righthand plot of Figure 2, we have over-plotted in horizontal lines and shaded regions the magnitudes for a 1 MJup object with an age of either 10 or 50 Myr based on theoretical predictions from the COND models (Baraffe et al. 2003) or more recent models published by Spiegel & Burrows (2012). The COND models predict that even for an age of 50 Myr we would have been able to detect objects with masses between 0.5 – 1 MJup for sepa- rations ≥5 AU. The models by Spiegel & Burrows (2012), however, predict that planets are significantly fainter in L′ compared to the COND models, regardless of their formation history (hot vs. cold start models). In this case, only for young ages (.20–30 Myr) are our detec- tion limits sufficient to detect a 1 MJup object in the background limit. For an assumed age of 40 Myr only objects with masses ≥2 MJup would have been detected. 4. DISCUSSION Our detection limits are, to our knowledge, the deep- est yet achieved for a star targeted by direct imaging in terms of companion mass at a given physical separa- tion. Our results also illustrate an important challenge in interpreting (detections and) non-detections from high- contrast surveys: the systematic and strong dependence upon thermal evolutionary model predictions (see right panel in Figure 2). The Spiegel & Burrows (2012) mod- els consistently predict fainter L-band magnitudes for an 1 MJup object compared to the Baraffe et al. (2003) COND models, most likely as a result of different as- sumptions regarding atmospheric opacities (Spiegel 2012, private communication). As mentioned in the introduc- tion, presently available atmospheric models fail to self- consistently explain the colors and luminosity of young, low surface gravity objects (e.g., Mohanty et al. 2007; Skemer et al. 2011; Barman et al. 2011; Skemer et al. 2012). Nevertheless, our observations demonstrate that AP Col it is unlikely to host a gas giant at separations ≥5 AU (Figure 4). For comparison, Delorme et al. (2011) have observed 14 nearby, young M-stars, each resulting in non-detections. We note, however, that the highest detection probability of that survey was achieved at sep- arations &30 AU. Monte Carlo simulations of current and 4 Quanz et al. planned high-contrast surveys indeed indicate that low- mass stars are least likely to harbor a directly detectable planet (Crepp & Johnson 2011). These simulations are based on empirical results from radial velocity studies that find occurrence rates of .2% for giant planets with close-in orbits (Johnson et al. 2007; Cumming et al. 2008; Bonfils et al. 2011). However, microlensing results suggest that 17+6 −9% of low-mass stars host a gas giant planet with mass be- tween 0.3 – 10 MJup at separations between 0.5 and 10 AU (Cassan et al. 2012). Thus, the planet frequency seems to rise between the orbital separations probed by Doppler observations and those by microlensing. Com- bining the currently available results from RV, microlens- ing, and direct imaging, Quanz et al. (2012) have shown that the population of gas giant planets with semi- major axes between ∼0.03 – 30 AU can be described by dfPlanet = M αaβdM da, with M being the planets' mass, a their semi-major axis, and α = −1.31 (cf. Cum- ming et al. 2008) and β . 0.5 − 0.6. To further constrain the value for β and the orbital separation range where it is valid, additional direct imaging observations effectively probing the regions between ∼10 – 30 AU are required. 5. SUMMARY & CONCLUSIONS Our findings can be summarized as the following: • We have used two different telescopes to obtain deep, high-contrast, L-band observations of the nearest (8.4 pc) pre-main sequence star, AP Col, to search for planetary mass companions. Neither dataset revealed the presence of a candidate com- panion. • Our derived detection limits are the most sensitive yet obtained by direct imaging in terms of planet mass for a given physical separation. For an as- sumed age of 40 Myr, we could have identified ob- jects with masses & 2 MJup at separations between 3 – 4.5 AU, and & 0.5 – 1 MJup for separations ≥4.5 AU using the predicted brightness of planets according to the COND models. • Accounting for the unknown inclination and posi- tion of a potential planet, we had a ∼50% chance to directly image an 1 MJup object at 5.2 AU, pro- vided a young Jupiter analogue exists. • Our detection limits depend on the predicted L- band magnitudes for young, planetary mass ob- jects. Using two different sets of theoretical model predictions, the limits change significantly, demon- strating that: (a) care need be taken for the inter- pretation of direct imaging results, and (b) empir- ical model constraints are urgently required. Dedicated imaging surveys for planetary mass com- panions around M-stars are ongoing (e.g., Delorme et al. 2011). They will eventually provide stringent constraints on the overall frequency and semimajor axis distribu- tion of gas giant planets at separations &10 AU, com- plementing the results of RV and microlensing studies. Our results support the notion that future high-contrast imaging programs would maximize their yield by pref- erentially selecting massive (A, F-type) stars as targets (Crepp & Johnson 2011). We thank ESO for granting us Director's Discretionary Time for our VLT/NACO observations and are indebted to Lowell Tacconi-Garman and Julien Girard for their excellent support during the observing run. SPQ thanks D. Spiegel for useful discussions about theroretical mod- els of planetary atmopsheres. This research has made use of the SIMBAD database, operated at CDS, Strasbourg, France, and of NASA's Astrophysics Data System. Facilities: VLT:Yepun (NACO) Facilities: Keck (NIRC2) REFERENCES Baraffe, I., Chabrier, G., Barman, T. S., Allard, F., & Hauschildt, P. H. 2003, A&A, 402, 701 Barman, T. S., Macintosh, B., Konopacky, Q. M., & Marois, C. 2011, ApJ, 733, 65 Bergfors, C., Brandner, W., Janson, M., Daemgen, S., Geissler, K., Henning, T., Hippler, S., Hormuth, F., Joergens, V., & Kohler, R. 2010, A&A, 520, A54 Beuzit, J.-L., Feldt, M., Dohlen, K., Mouillet, D., Puget, P., Antichi, J., Baruffolo, A., Baudoz, P., Berton, A., Boccaletti, A., Carbillet, M., Charton, J., Claudi, R., Downing, M., Feautrier, P., Fedrigo, E., Fusco, T., Gratton, R., Hubin, N., Kasper, M., Langlois, M., Moutou, C., Mugnier, L., Pragt, J., Rabou, P., Saisse, M., Schmid, H. M., Stadler, E., Turrato, M., Udry, S., Waters, R., & Wildi, F. 2006, The Messenger, 125, 29 Biller, B. A., Close, L. M., Masciadri, E., Nielsen, E., Lenzen, R., Brandner, W., McCarthy, D., Hartung, M., Kellner, S., Mamajek, E., Henning, T., Miller, D., Kenworthy, M., & Kulesa, C. 2007, ApJS, 173, 143 Bonfils, X., Delfosse, X., Udry, S., Forveille, T., Mayor, M., Perrier, C., Bouchy, F., Gillon, M., Lovis, C., Pepe, F., Queloz, D., Santos, N. C., S´egransan, D., & Bertaux, J.-L. 2011, ArXiv e- prints Bowler, B. P., Liu, M. C., Dupuy, T. J., & Cushing, M. C. 2010, ApJ, 723, 850 Cassan, A., Kubas, D., Beaulieu, J.-P., Dominik, M., Horne, K., Greenhill, J., Wambsganss, J., Menzies, J., Williams, A., Jørgensen, U. G., Udalski, A., Bennett, D. P., Albrow, M. D., Batista, V., Brillant, S., Caldwell, J. A. R., Cole, A., Coutures, C., Cook, K. H., Dieters, S., Prester, D. D., Donatowicz, J., Fouqu´e, P., Hill, K., Kains, N., Kane, S., Marquette, J.-B., Martin, R., Pollard, K. R., Sahu, K. C., Vinter, C., Warren, D., Watson, B., Zub, M., Sumi, T., Szyma´nski, M. K., Kubiak, M., Poleski, R., Soszynski, I., Ulaczyk, K., Pietrzy´nski, G., & Wyrzykowski, L. 2012, Nature, 481, 167 Chauvin, G., Lagrange, A., Bonavita, M., Zuckerman, B., Dumas, C., Bessell, M. S., Beuzit, J., Bonnefoy, M., Desidera, S., Farihi, J., Lowrance, P., Mouillet, D., & Song, I. 2010, A&A, 509, A52+ Chauvin, G., Lagrange, A., Dumas, C., Zuckerman, B., Mouillet, D., Song, I., Beuzit, J., & Lowrance, P. 2005, A&A, 438, L25 Cox, A. N. 2000, Allen's astrophysical quantities, ed. Cox, A. N. Crepp, J. R. & Johnson, J. A. 2011, ApJ, 733, 126 Crepp, J. R., Johnson, J. A., Fischer, D. A., Howard, A. W., Marcy, G. W., Wright, J. T., Isaacson, H., Boyajian, T., von Braun, K., Hillenbrand, L. A., Hinkley, S., & Carpenter, J. M. 2011, ArXiv e-prints Cumming, A., Butler, R. P., Marcy, G. W., Vogt, S. S., Wright, J. T., & Fischer, D. A. 2008, PASP, 120, 531 Searching for young Jupiter Analogs around AP Col 5 Cutri, R. M., Skrutskie, M. F., van Dyk, S., Beichman, C. A., Carpenter, J. M., Chester, T., Cambresy, L., Evans, T., Fowler, J., Gizis, J., Howard, E., Huchra, J., Jarrett, T., Kopan, E. L., Kirkpatrick, J. D., Light, R. M., Marsh, K. A., McCallon, H., Schneider, S., Stiening, R., Sykes, M., Weinberg, M., Wheaton, W. A., Wheelock, S., & Zacarias, N. 2003, 2MASS All Sky Catalog of point sources., ed. Cutri, R. M., Skrutskie, M. F., van Dyk, S., Beichman, C. A., Carpenter, J. M., Chester, T., Cambresy, L., Evans, T., Fowler, J., Gizis, J., Howard, E., Huchra, J., Jarrett, T., Kopan, E. L., Kirkpatrick, J. D., Light, R. M., Marsh, K. A., McCallon, H., Schneider, S., Stiening, R., Sykes, M., Weinberg, M., Wheaton, W. A., Wheelock, S., & Zacarias, N. Delorme, P., Lagrange, A. M., Chauvin, G., Bonavita, M., Lacour, S., Bonnefoy, M., Ehrenreich, D., & Beust, H. 2011, ArXiv e- prints Esposito, S., Mesa, D., Skemer, A., Arcidiacono, C., Claudi, R. U., Desidera, S., Gratton, R., Mannucci, F., Marzari, F., Masciadri, E., Close, L., Hinz, P., Kulesa, C., McCarthy, D., Males, J., Agapito, G., Argomedo, J., Boutsia, K., Briguglio, R., Brusa, G., Busoni, L., Cresci, G., Fini, L., Fontana, A., Guerra, J. C., Hill, J. M., Miller, D., Paris, D., Puglisi, A., Quiros-Pacheco, F., Riccardi, A., Stefanini, P., Testa, V., Xompero, M., & Woodward, C. 2012, ArXiv e-prints Ghez, A. M., Salim, S., Weinberg, N. N., Lu, J. R., Do, T., Dunn, J. K., Matthews, K., Morris, M. R., Yelda, S., Becklin, E. E., Kremenek, T., Milosavljevic, M., & Naiman, J. 2008, ApJ, 689, 1044 Gould, A., Dong, S., Gaudi, B. S., Udalski, A., Bond, I. A., Greenhill, J., Street, R. A., Dominik, M., Sumi, T., Szyma´nski, M. K., Han, C., Allen, W., Bolt, G., Bos, M., Christie, G. W., DePoy, D. L., Drummond, J., Eastman, J. D., Gal- Yam, A., Higgins, D., Janczak, J., Kaspi, S., Koz lowski, S., Lee, C.-U., Mallia, F., Maury, A., Maoz, D., McCormick, J., Monard, L. A. G., Moorhouse, D., Morgan, N., Natusch, T., Ofek, E. O., Park, B.-G., Pogge, R. W., Polishook, D., Santallo, R., Shporer, A., Spector, O., Thornley, G., Yee, J. C., µFUN Collaboration, Kubiak, M., Pietrzy´nski, G., Soszy´nski, I., Szewczyk, O., Wyrzykowski, L., Ulaczyk, K., Poleski, R., OGLE Collaboration, Abe, F., Bennett, D. P., Botzler, C. S., Douchin, D., Freeman, M., Fukui, A., Furusawa, K., Hearnshaw, J. B., Hosaka, S., Itow, Y., Kamiya, K., Kilmartin, P. M., Korpela, A., Lin, W., Ling, C. H., Makita, S., Masuda, K., Matsubara, Y., Miyake, N., Muraki, Y., Nagaya, M., Nishimoto, K., Ohnishi, K., Okumura, T., Perrott, Y. C., Philpott, L., Rattenbury, N., Saito, T., Sako, T., Sullivan, D. J., Sweatman, W. L., Tristram, P. J., von Seggern, E., Yock, P. C. M., MOA Collaboration, Albrow, M., Batista, V., Beaulieu, J. P., Brillant, S., Caldwell, J., Calitz, J. J., Cassan, A., Cole, A., Cook, K., Coutures, C., Dieters, S., Dominis Prester, D., Donatowicz, J., Fouqu´e, P., Hill, K., Hoffman, M., Jablonski, F., Kane, S. R., Kains, N., Kubas, D., Marquette, J.-B., Martin, R., Martioli, E., Meintjes, P., Menzies, J., Pedretti, E., Pollard, K., Sahu, K. C., Vinter, C., Wambsganss, J., Watson, R., Williams, A., Zub, M., PLANET Collaboration, Allan, A., Bode, M. F., Bramich, D. M., Burgdorf, M. J., Clay, N., Fraser, S., Hawkins, E., Horne, K., Kerins, E., Lister, T. A., Mottram, C., Saunders, E. S., Snodgrass, C., Steele, I. A., Tsapras, Y., RoboNet Collaboration, Jørgensen, U. G., Anguita, T., Bozza, V., Calchi Novati, S., Harpsøe, K., Hinse, T. C., Hundertmark, M., Kjaergaard, P., Liebig, C., Mancini, L., Masi, G., Mathiasen, M., Rahvar, S., Ricci, D., Scarpetta, G., Southworth, J., Surdej, J., Thone, C. C., & MiNDSTEp Consortium. 2010, ApJ, 720, 1073 Heinze, A. N., Hinz, P. M., Kenworthy, M., Meyer, M., Sivanandam, S., & Miller, D. 2010, ApJ, 714, 1570 Hinkley, S., Oppenheimer, B. R., Zimmerman, N., Brenner, D., Parry, I. R., Crepp, J. R., Vasisht, G., Ligon, E., King, D., Soummer, R., Sivaramakrishnan, A., Beichman, C., Shao, M., Roberts, L. C., Bouchez, A., Dekany, R., Pueyo, L., Roberts, J. E., Lockhart, T., Zhai, C., Shelton, C., & Burruss, R. 2011, PASP, 123, 74 Hinz, P. M., Rodigas, T. J., Kenworthy, M. A., Sivanandam, S., Heinze, A. N., Mamajek, E. E., & Meyer, M. R. 2010, ApJ, 716, 417 Janson, M., Bergfors, C., Goto, M., Brandner, W., & Lafreni`ere, D. 2010, ApJ, 710, L35 Johnson, J. A., Aller, K. M., Howard, A. W., & Crepp, J. R. 2010, PASP, 122, 905 Johnson, J. A., Butler, R. P., Marcy, G. W., Fischer, D. A., Vogt, S. S., Wright, J. T., & Peek, K. M. G. 2007, ApJ, 670, 833 Kasper, M., Apai, D., Janson, M., & Brandner, W. 2007, A&A, 472, 321 Kennedy, G. M. & Kenyon, S. J. 2008, ApJ, 673, 502 Lafreni`ere, D., Doyon, R., Marois, C., Nadeau, D., Oppenheimer, B. R., Roche, P. F., Rigaut, F., Graham, J. R., Jayawardhana, R., Johnstone, D., Kalas, P. G., Macintosh, B., & Racine, R. 2007a, ApJ, 670, 1367 Lafreni`ere, D., Marois, C., Doyon, R., Nadeau, D., & Artigau, ´E. 2007b, ApJ, 660, 770 Lagrange, A.-M., Bonnefoy, M., Chauvin, G., Apai, D., Ehrenreich, D., Boccaletti, A., Gratadour, D., Rouan, D., Mouillet, D., Lacour, S., & Kasper, M. 2010, Science, science.1187187 Lenzen, R., Hartung, M., Brandner, W., Finger, G., Hubin, N. N., Lacombe, F., Lagrange, A., Lehnert, M. D., Moorwood, A. F. M., & Mouillet, D. 2003, in Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series, Vol. 4841, Society of Photo- Optical Instrumentation Engineers (SPIE) Conference Series, ed. M. Iye & A. F. M. Moorwood, 944–952 Macintosh, B., Graham, J., Palmer, D., Doyon, R., Gavel, D., Larkin, J., Oppenheimer, B., Saddlemyer, L., Wallace, J. K., Bauman, B., Erikson, D., Poyneer, L., Sivaramakrishnan, A., Soummer, R., & Veran, J.-P. 2007, Comptes Rendus Physique, 8, 365 Marois, C., Lafreni`ere, D., Doyon, R., Macintosh, B., & Nadeau, D. 2006, ApJ, 641, 556 Marois, C., Macintosh, B., Barman, T., Zuckerman, B., Song, I., Patience, J., Lafreni`ere, D., & Doyon, R. 2008, Science, 322, 1348 Marois, C., Zuckerman, B., Konopacky, Q. M., Macintosh, B., & Barman, T. 2010, Nature, 468, 1080 Masciadri, E., Mundt, R., Henning, T., Alvarez, C., & Barrado y Navascu´es, D. 2005, ApJ, 625, 1004 Mohanty, S., Jayawardhana, R., Hu´elamo, N., & Mamajek, E. 2007, ApJ, 657, 1064 Pollack, J. B., Hubickyj, O., Bodenheimer, P., Lissauer, J. J., Podolak, M., & Greenzweig, Y. 1996, Icarus, 124, 62 Pueyo, L., Crepp, J. R., Vasisht, G., Brenner, D., Oppenheimer, B. R., Zimmerman, N., Hinkley, S., Parry, I., Beichman, C., Hillenbrand, L., Roberts, L. C., Dekany, R., Shao, M., Burruss, R., Bouchez, A., Roberts, J., & Soummer, R. 2012, ApJS, 199, 6 Quanz, S. P., Kenworthy, M. A., Meyer, M. R., Girard, J. H. V., & Kasper, M. 2011, ApJ, 736, L32 Quanz, S. P., Lafreni`ere, D., Meyer, M. R., Reggiani, M. M., & Buenzli, E. 2012, A&A, 541, A133 Quanz, S. P., Meyer, M. R., Kenworthy, M. A., Girard, J. H. V., Kasper, M., Lagrange, A., Apai, D., Boccaletti, A., Bonnefoy, M., Chauvin, G., Hinz, P. M., & Lenzen, R. 2010, ApJ, 722, L49 Riedel, A. R., Murphy, S. J., Henry, T. J., Melis, C., Jao, W.-C., & Subasavage, J. P. 2011, AJ, 142, 104 Rousset, G., Lacombe, F., Puget, P., Hubin, N. N., Gendron, E., Fusco, T., Arsenault, R., Charton, J., Feautrier, P., Gigan, P., Kern, P. Y., Lagrange, A., Madec, P., Mouillet, D., Rabaud, D., Rabou, P., Stadler, E., & Zins, G. 2003, in Society of Photo- Optical Instrumentation Engineers (SPIE) Conference Series, Vol. 4839, Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series, ed. P. L. Wizinowich & D. Bonaccini, 140–149 Skemer, A. J., Close, L. M., Szucs, L., Apai, D., Pascucci, I., & Biller, B. A. 2011, ApJ, 732, 107 Skemer, A. J., Hinz, P. M., Esposito, S., Burrows, A., Leisenring, J., Skrutskie, M., Desidera, S., Mesa, D., Arcidiacono, C., Mannucci, F., Rodigas, T. J., Close, L., McCarthy, D., Kulesa, C., Agapito, G., Apai, D., Argomedo, J., Bailey, V., Boutsia, K., Briguglio, R., Brusa, G., Busoni, L., Claudi, R., Eisner, J., Fini, L., Follette, K. B., Garnavich, P., Gratton, R., Guerra, J. C., Hill, J. M., Hoffmann, W. F., Jones, T., Krejny, M., Males, J., Masciadri, E., Meyer, M. R., Miller, D. L., Morzinski, K., Nelson, M., Pinna, E., Puglisi, A., Quanz, S. P., Quiros-Pacheco, F., Riccardi, A., Stefanini, P., Vaitheeswaran, V., Wilson, J. C., & Xompero, M. 2012, ArXiv e-prints Spiegel, D. S. & Burrows, A. 2012, ApJ, 745, 174 Tamura, M. 2009, in American Institute of Physics Conference Series, Vol. 1158, American Institute of Physics Conference Series, ed. T. Usuda, M. Tamura, & M. Ishii, 11–16 6 Quanz et al. Wizinowich, P. L., Le Mignant, D., Bouchez, A., Chin, J., Contos, A., Hartman, S., Johansson, E., Lafon, R., Neyman, C., Stomski, P., Summers, D., & van Dam, M. A. 2004, in Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series, Vol. 5490, Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series, ed. D. Bonaccini Calia, B. L. Ellerbroek, & R. Ragazzoni, 1–11 Searching for young Jupiter Analogs around AP Col 7 Fig. 1.- Comparison between the 5–σ contrast curves obtained with Keck/NIRC2 and VLT/NACO. Fig. 2.- 5–σ detection limit in L′ magnitudes for our VLT/NACO data as a function of separation from AP Col (black crosses and curve). Left: Overplotted as horizontal dashed lines are predicted L′ magnitudes for 40 Myr old objects with masses of 0.5 , 1 and 2 MJup at the distance of AP Col based on the COND models (Baraffe et al. 2003). The red asterisk shows the location of a young Jupiter analog (i.e, 1 MJup at 5.2 AU projected separation). Right: Overplotted are predictions from theoretical models for a 1 MJup object with an age of 10 or 50 Myr. The blue, dashed lines are based on the COND models. The grey shaded areas are predictions from the hot and cold start models from Spiegel & Burrows (2012), where the spread in magnitude for a given model, i.e., the width of the grey shaded areas, shows the difference between atmospheres with and without hybrid clouds. All models assume solar metallicity. Summary of deep L′ imaging observations of AP Col TABLE 1 Parameter No. of detector reads × exp. time No. of data cubes / stacked images Parallactic angle start / end Airmass range Typical DIMM seeing PSF FWHMa hECimean b VLT/NACO 64×0.2 s 134 cubes -50.11◦ / +59.70◦ 1.02 – 1.05 0.7′′. . . 1.0′′ ∼0.12′′ (0.10′′) 19.1% Keck/NIRC2 300×0.106 s 67 images -16.24◦ / +5.59◦ 1.72 – 1.78 ∼0.6′′ ∼0.09′′ (0.08′′) - aMeasured FWHM and theoretical diffraction limited FWHM in parenthesis. bAverage value of the coherent energy of the NACO/PSF in data cubes. Calculated by the Real Time Computer of the AO system. 8 Quanz et al. Fig. 3.- Final VLT/NACO image after a fake "Jupiter" has been inserted in the raw images at two different positions with a projected separation of 5.2 AU from the central star. The brightness of the "planets" corresponds to a 1 MJup object with an age of 40 Myr based on the COND models. Both "planets" are clearly detected with with a signal-to-noise of ∼9.6 and ∼12.1, respectively. The innermost regions that are dominated by subtraction residuals have been masked out. Fig. 4.- Detection probability of a 1 MJup object (black, solid line) and a 2 MJup object (red, dashed line) for our single epoch VLT/NACO observation assuming the detection limits from Figure 2 (left). The vertical, dashed line indicates the position of Jupiter in our Solar System at 5.2 AU.
1809.01001
1
1809
2018-09-04T14:09:54
High-Contrast study of the candidate planets and protoplanetary disk around HD~100546
[ "astro-ph.EP" ]
The nearby Herbig Be star HD100546 is known to be a laboratory for the study of protoplanets and their relation with the circumstellar disk that is carved by at least 2 gaps. We observed the HD100546 environment with high contrast imaging exploiting several different observing modes of SPHERE, including datasets with/without coronagraphs, dual band imaging, integral field spectroscopy and polarimetry. The picture emerging from these different data sets is complex. Flux-conservative algorithms images clearly show the disk up to 200au. More aggressive algorithms reveal several rings and warped arms overlapping the main disk. The bright parts of this ring lie at considerable height over the disk mid-plane at about 30au. Our images demonstrate that the brightest wings close to the star in the near side of the disk are a unique structure, corresponding to the outer edge of the intermediate disk at ~40au. Modeling of the scattered light from the disk with a geometrical algorithm reveals that a moderately thin structure can well reproduce the light distribution in the flux-conservative images. We suggest that the gap between 44 and 113 au span between the 1:2 and 3:2 resonance orbits of a massive body located at ~70au that might coincide with the candidate planet HD100546b detected with previous thermal IR observations. In this picture, the two wings can be the near side of a ring formed by disk material brought out of the disk at the 1:2 resonance with the same massive object. While we find no clear evidence confirming detection of the planet candidate HD100546c in our data, we find a diffuse emission close to the expected position of HD100546b. This source can be described as an extremely reddened substellar object surrounded by a dust cloud or its circumplanetary disk. Its astrometry is broadly consistent with a circular orbital motion on the disk plane.
astro-ph.EP
astro-ph
Astronomy & Astrophysics manuscript no. hd100546Sissa September 5, 2018 c(cid:13)ESO 2018 High-Contrast study of the candidate planets and protoplanetary disk around HD 100546(cid:63) E. Sissa1, 2, R. Gratton1, A. Garufi3, 4, E. Rigliaco1, A. Zurlo5, 6 , D. Mesa1, 7, M. Langlois7, 8, J. de Boer9, S. Desidera1, C. Ginski9, A.-M. Lagrange10, A.-L. Maire11, A. Vigan6, M. Dima1, J. Antichi1, 12, A. Baruffolo1, A. Bazzon3, M. Benisty10, J.-L. Beuzit10, B. Biller11, 13, A. Boccaletti14, M. Bonavita1, 13, M. Bonnefoy10, W. Brandner11, P. Bruno15, E. Buenzli3, E. Cascone16, G. Chauvin10, 17, A. Cheetham18, R.U. Claudi1, M. Cudel10, V. De Caprio16, C. Dominik9, D. Fantinel1, G. Farisato1, M. Feldt11, C. Fontanive13, R. Galicher14, E. Giro1, 19, J. Hagelberg10, S. Incorvaia20, M. Janson11, 21, M. Kasper10, 22, M. Keppler11, T. Kopytova11, E. Lagadec23, J. Lannier10, C. Lazzoni1, 2, H. LeCoroller6, L. Lessio1, R. Ligi6, F. Marzari1, F. Menard10, M.R. Meyer3, 24, D. Mouillet10, S. Peretti18, C. Perrot14, P. J. Potiron23, D. Rouan14, B. Salasnich1, G. Salter6, M. Samland11, T. Schmidt14, S. Scuderi15, F. Wildi18 (Affiliations can be found after the references) Received / Accepted ABSTRACT The nearby Herbig Be star HD 100546 is known to be a laboratory for the study of protoplanets and their relation with the circumstellar disk, that is carved by at least two gaps. We observed the HD 100546 environment with high contrast imaging exploiting several different observing modes of SPHERE, including data sets with/without coronagraphs, dual band imaging, integral field spectroscopy and polarimetry. The picture emerging from these different data sets is complex. Flux-conservative algorithms images clearly show the disk up to 200 au. More aggressive algorithms reveal several rings and warped arms that are seen overlapping the main disk. The bright parts of this ring are found out to lie at considerable height over the disk mid-plane at about 30 au. Our images demonstrate that the brightest wings close to the star in the near side of the disk are a unique structure, corresponding to the outer edge of the intermediate disk at ∼ 40 au. Modelling of the scattered light from the disk with a geometrical algorithm reveals that a moderately thin structure (H/r=0.18 at 40 au) can well reproduce the light distribution in the flux-conservative images. We suggest that the gap between 44 au and 113 au span between the 1:2 and 3:2 resonance orbits of a massive body located at ∼ 70 au, that might coincide with the candidate planet HD 100546b detected with previous thermal IR observations. In this picture, the two wings can be the near side of a ring formed by disk material brought out of the disk at the 1:2 resonance with the same massive object. While we find no clear evidence confirming detection of the planet candidate HD 100546c in our data, we find a diffuse emission close to the expected position of HD 100546b. This source can be described as an extremely reddened substellar object surrounded by a dust cloud or its circumplanetary disk. Its astrometry is broadly consistent with a circular orbital motion on the disk plane, a result that could be confirmed with new observations. Further observations at various wavelengths are required to fully understand the complex phenomenology of HD 100546. Key words. star: individual: HD 100546 - techniques: high angular resolution, polarimetric - Planets and satellites: detection - protoplanetary disks 1. Introduction The number of known exoplanets around main-sequence stars is increasing rapidly. However, to date, the number of detected (forming) exoplanets around pre-main-sequence stars is still low. Possibly the best examples are LkCa 15 (Kraus & Ireland 2012; Sallum et al. 2015), HD 169142 (Biller et al. 2014; Reggiani et al. 2014), HD 100546 (Quanz et al. 2013; Currie et al. 2014; Quanz et al. 2015; Currie et al. 2015; Rameau et al. 2017) and PDS 70 (Keppler et al. 2018; Müller et al. 2018). To understand the formation pro- Send offprint requests to: E. Sissa, e-mail: [email protected] (cid:63) Based on data collected at the European Southern Observa- tory, Chile (ESO Programs 095.C-0298, 096.C-0241, 096.C-0248, 097.C-0523, 097.C-0865 and 098.C-0209). cess of planets, we need to study the initial condition and evolution of circumstellar disks, and how the disk can be shaped by ongoing planet formation. Recently developed high-contrast and high-angular res- olution imaging instruments such as SPHERE (Spectro- Polarimetric High-contrast Exoplanet REsearch; Beuzit et al. 2008), GPI (Gemini Planet Imager; Macintosh et al. 2014), SCExAO (Subaru Coronagraphic Extreme Adaptive Optics, Jovanovic et al. 2015) and FLAO (First Light AO system, Quirós-Pacheco et al. 2010) provide the excellent capability to directly obtain images of protoplanetary disks around nearby young stars, up to the inner tens of au, in scattered light and thermal emission. This makes sometime possible to observe planets in their birthplace investigat- ing their interaction with the disk. Imaging protoplanetary disks around young stars allows us to detect spiral struc- Article number, page 1 of 19 8 1 0 2 p e S 4 . ] P E h p - o r t s a [ 1 v 1 0 0 1 0 . 9 0 8 1 : v i X r a A&A proofs: manuscript no. hd100546Sissa tures and gaps which might be produced by gravitational perturbation of forming planets, as demonstrated by e.g. Grady et al. (2001); Thalmann et al. (2010); Garufi et al. (2013); Pinilla et al. (2015); Dong et al. (2016). HD 100546 is a nearby (d=109 ± 4 pc, Gaia Collabora- tion et al. 2016) well studied Herbig Be star with spectral type B9Vne (Levenhagen & Leister 2006) which harbours a large disk. Several studies have been conducted in the past years on this disk, from spectroscopic and photometric anal- ysis to direct imaging. The first evidence of the presence of the disk around HD 100546 was obtained by Hu et al. (1989) measuring the IR excess in the SED. The mid-IR excess in the SED of this source requires a thickening of the disk at ∼ 0.1(cid:48)(cid:48), that can be explained by a proto-Jupiter carving a gap (Bouwman et al. 2003). This idea is also supported by far-ultraviolet long-slit spectroscopy with HST/STIS, that detected a central cavity up to 0.13(cid:48)(cid:48) (Grady et al. 2005), by UVES observations of [OI] emission region (Acke & van den Ancker 2006), using spectro-astrometry with CO roto- vibrational emission by van der Plas et al. (2009) and by AMBER/VLTI observations in the K-band (Benisty et al. 2010). The disk was first imaged in scattered light in the J and K bands with the adaptive optics system ADONIS coupled with a pre-focal optics coronagraph (Pantin et al. 2000). The disk was detected up to 2(cid:48)(cid:48) from the star, with a density peak at ∼ 40 au. Many successive works revealed the complexity of its geometry by means of both scattered light images (e.g., Augereau et al. 2001; Ardila et al. 2007; Quanz et al. 2011; Garufi et al. 2016; Follette et al. 2017) and also, (sub-)millimeter images (e.g., Walsh et al. 2014; Pineda et al. 2014; Wright et al. 2015). The radial pro- file of the disk brightness becomes less steep at separations < 2.7(cid:48)(cid:48) in both HST/NICMOS2 H-band (Augereau et al. 2001) and ADONIS Ks (Grady et al. 2001) band. However this change is not visible in the optical. Moreover, the semi- minor axis brightness profile is asymmetric in the H-band (Augereau et al. 2001); this can be related to an optically thick circumstellar disk inside 0.8(cid:48)(cid:48) and an optically thin disk at larger separations. Thanks to the higher angular resolution reachable with HST/STIS coronagraphic images, the first disk structures were detected: spiral dark filaments appear mostly in the SW at separations greater then 2.3(cid:48)(cid:48)(Grady et al. 2001). Following observations revealed additional spiral structures both at short and at far separations, possibly related to forming planets (Ardila et al. 2007; Boccaletti et al. 2013; Currie et al. 2015; Garufi et al. 2016). The first Polarimet- ric Differential Imaging (PDI) detection of the disk (Quanz et al. 2011) in H and Ks NACO filters resolved the disk between 0.1(cid:48)(cid:48) and 1.4(cid:48)(cid:48), locating the disk inner rim at 0.15(cid:48)(cid:48). They noted an asymmetry along the brightness profile that they interpreted as the presence of two dust populations. Deeper observations by Avenhaus et al. (2014) reached the innermost region (∼ 0.03(cid:48)(cid:48)) and found a spiral arm in the far-side of the disk. The disk is found out to be strongly flared in the outer regions (> 80-100 au) and it is expected to generate shadowing effect on the forward-scattering side of the disk. The disk appears faint and red, due to the combination of particles sizes and disk geometry as demon- strated by Mulders et al. (2013) and later confirmed by Stolker et al. (2016) comparing models to PDI data. In addition to these data from the disk, several stud- ies suggest that HD 100456 may host at least two planets Article number, page 2 of 19 (Quanz et al. 2013; Brittain et al. 2014; Quanz et al. 2015; Currie et al. 2015) located at ∼ 55 au (hereafter CCb) and ∼ 13 au (hereafter CCc) from the central star. This makes the system a powerful laboratory to study planet formation and planet-disk interaction. Quanz et al. (2015) inferred the temperature and emit- ting radius of CCb in the L(cid:48) and M bands and concluded that what they imaged could be a warm circumplanetary disk rather than the planet itself. An extended emission at near-IR wavelengths, possibly linked with this candidate planet was also detected by Currie et al. (2015) centred at 0.469± 0.012(cid:48)(cid:48), P A=7.04◦±1.39◦. Its infrared colors are ex- tremely red compared to models, indicating that it could still be embedded by accreting material. The existence of CCc was proposed by Brittain et al. (2013), who ascribed the variability of the CO ro- vibrational lines to a compact source of emission at ∼15 au from the central star. Currie et al. (2015) showed that this object is detected in GPI H-band images. However there is a debate on this result, because detection strongly depends on the reduction method used (see e.g. Garufi et al. 2016; Follette et al. 2017; Currie et al. 2017) and on its nature, since it can be interpreted with alternative equally plau- sible origins like disk hot spot or indirect signature of the presence of the planet. First observations of HD 100546 with VLT/SPHERE were presented in Garufi et al. (2016), where it was found that the disk around HD 100546 is truncated at about 11 au and the cavity is consistent with being intrinsically circular. More recent works by Follette et al. (2017); Rameau et al. (2017); Currie et al. (2017), all based on GPI data sets, debate on the presence of both the planets and the impact of the reduction methods on them. HD 100546 is located in the Sco-Cen complex. In this analysis we consider a stellar mass of M∗=2.4M(cid:12) (Brittain et al. 2014), and apparent magnitudes of J=6.42, H=5.96, and K=5.42 (Cutri et al. 2003), as well as L(cid:48)=4.52 and M(cid:48)=4.13 as in Quanz et al. (2015). For the disk, we assume an inclination i=42◦ (Ardila et al. 2007; Pineda et al. 2014) and P A=146◦ (Pineda et al. 2014). In this paper, we present the results of a two year ob- servational campaign on this star, based on observations in direct imaging with VLT/SPHERE, the new high-contrast and high-resolution instrument dedicated to exoplanet and disk imaging, that includes both pupil tracking and polari- metric observations. SPHERE includes a powerful extreme adaptive optics system (Sphere AO for eXoplanet Observa- tion, SAXO, Fusco et al. 2006), providing a currently un- matched Strehl Ratio of up to 92% in the H-band for bright (R < 9) sources, various coronagraphs (see Martinez et al. 2009; Carbillet et al. 2011), an infrared differential imaging camera (IRDIS, Dohlen et al. 2008), an infrared integral field spectrograph (IFS, Claudi et al. 2008) and a visible differential polarimeter and imager (ZIMPOL, Thalmann et al. 2008). The paper is structured as follows. The observing strat- egy, sky conditions and the reduction of these data set are described in Sect. 2. The results for the disk and planets are given in Sect. 3 and 4 respectively. Conclusions are given in Sec. 5. E. Sissa et al.: High-Contrast study of the candidate planets and protoplanetary disk around HD 100546 were obtained combining the root mean square scatter of a 1λ/D diameter spot along a 1λ/D wide annulus, at dif- ferent separation from the star as described in Mesa et al. (2015). These contrast values were then corrected for the low number statistics according to Mawet et al. (2014). We noticed that these raw image contrast are comparable to those obtained with the two GPI data sets presented in Currie et al. (2017). We found that the highest quality coronagraphic ob- servations are those from May 2015 and May 2016 while the observations obtained in January and March 2016 are of poor quality due to the atmospheric conditions or low system performances (calibration of the deformable mirror voltages was not optimal in January 2016). For the non- coronagraphic data sets, the lack of a coronagraph implies much stronger diffraction patterns and the need of using the neutral density filter, common to IFS and IRDIS, to avoid saturation close to the center. This strongly reduces the signal; for this reason the sensitivity of these images is far from optimal at separations larger than 0.1(cid:48)(cid:48) but they allow access to the closest separations with unprecedented spatial resolution. 2.2. IRDIFS and IRDIFS_EXT data reduction We performed the basic data reduction of IRDIS and IFS (bad pixel removal, flat fielding, image alignment, sky sub- traction) with version 0.15.0 of SPHERE Data Reduction and Handling (DRH) pipeline (Pavlov et al. 2008). Fur- ther elaboration of the images (deconvolution for lenslet- to-lenslet cross talk, refinement of the wavelength calibra- tion, correction for distortion, fine centering of the images, frame selection) was performed at the SPHERE Data Cen- ter (DC) in Grenoble1 (Delorme et al. 2017). Additional details on the adopted procedures are described in Zurlo et al. (2014), Mesa et al. (2015) and Maire et al. (2016). We then applied various algorithms for differential imaging such as classical ADI (cADI, Marois et al. 2006), template- LOCI (TLOCI, Marois et al. 2014), and Principal Compo- nent Analysis (PCA, Soummer et al. 2012; Amara & Quanz 2012). These procedures are available at the SPHERE DC through the SpeCal software (Galicher et al. 2018). For IFS, the cADI is based on a median combination of the images, while the PCA is evaluated on the whole image at the same time; no exclusion zones are considered. For IRDIS, no spatial filtering was applied to the data beforehand. We performed several reduction and, verified that both disk and wide companions (see Appendix A) are detected using the alternate methods. We also confirm that these detections are robust across a range of algorithm pa- rameter space (e.g. varying numbers of principal compo- nents from 2 to 5). In classical ADI, SpeCal averages the cube of frames over the angular dimension for each spec- tral channel to remove the speckle contamination. After a rotation is applied, the frames are averaged using mean or median combination to produce one final image per spec- tral channel. The median-combination is applied for disk images because it is less sensitive to uncorrected hot/bad pixels. When estimating the candidate companions pho- tometry the mean is used instead of the median to pre- serve linearity. PCA is evaluated on the whole image at the same time; no exclusion zones are considered and the 1 http://sphere.osug.fr/spip.php?rubrique16&lang=en Article number, page 3 of 19 Fig. 1. 5σ raw contrast of the IFS images at all epochs but March 2016. 2. Observations and data reduction 2.1. Observations We observed HD 100546 at different epochs as part of the SHINE (SpHere INfrared survey for Exoplanets) Guaran- teed Time Observations (GTO) program using IRDIS and IFS simultaneously: the IRDIFS mode (with IFS operating between 0.95 and 1.35 µm and IRDIS working in in dual imaging mode in the H2H3 filter pair at 1.59 and 1.67 µm, Vigan et al. 2010 ) and the IRDIFS_EXT mode (with IFS operating at Y-H wavelengths 0.95-1.65 µm and IRDIS in the K1K2 band filters at 2.11 and 2.25 µm). These observa- tions were all carried-out with the N_ALC_YJH_S apodized Lyot coronagraph (inner working angle, IWA ∼ 0.1(cid:48)(cid:48), Boc- caletti et al. 2008) except for two sequences which were acquired without coronagraph in order to study the central region (at less than 0.1(cid:48)(cid:48) from the star). In this case, to avoid heavy saturation of the star, a neutral density filter with average transmission ∼1/100 was used. All data were taken in pupil stabilized mode in order to perform Angular Differential Imaging (ADI, see e.g. Marois et al. 2006) for subtracting the stellar residuals. HD 100546 was also observed in Polarimetric Differen- tial Imaging (PDI, see e.g. Kuhn et al. 2001; Quanz et al. 2011), as part of the DISK GTO program, using IRDIS with the same coronagraphic mask as for the classical imaging. The source was observed for a total time of 21 minutes in the J band (1.26 µm) and 38 minutes in the K band (2.181 µm) during the night of March 31st, 2016, and May 25th, 2016, respectively. The orientation of the derotator was cho- sen in order to optimize the polarimetric efficiency of IRDIS (de Boer et al., in prep.). Several of the observations were carried-out under mod- erate and good weather conditions (with an average coher- ence time, τ0, longer than 2.0 milliseconds). Details on the full set of observations are presented in Table 1, τ0 and Strehl ratio were measured by SPARTA, the ESO standard real-time computer platform that controls the AO loop. In order to access the quality of the observational ma- terial, we plotted in Fig. 1 the contrast we achieved in the IFS raw images as a function of the separation for all the non polarimetric images but March 2016 data set, due to its very unstable sky conditions. The contrast plotted here 0.10.20.30.40.50.60.70.810-510-410-310-210-10.10.20.30.40.50.60.70.8Sep [au]10-510-410-310-210-15σ raw contrastMay 2015Jun 2015May 2016Jan 2016Apr 2016Feb 2017 Table 1. Observations summary A&A proofs: manuscript no. hd100546Sissa Date May 3 2015 May 29 2015 Jan 17 2016 Mar 26 2016 Mar 31 2016 Apr 16 2016 May 25 2016 May 31 2016 Feb 7 2017 Obs. mode IRDIFS_EXT IRDIFS IRDIFS_EXT IRDIFS_EXT IRDIS DPI J IRDIFS_EXT IRDIS DPI K IRDIFS_EXT IRDIFS_EXT texp [s] Rot [◦] 22.32 3840 36.37 6048 22.23 4096 4096 22.25 1280 4512 2304 4400 5280 28.95 28.38 29.42 - - 92.5 92.5 92.5 92.5 92.5 no 92.5 92.5 no rC [mas] ND Filter σ [(cid:48)(cid:48)] 0.75 0.70 1.54 2.58 1.98 0.66 0.75 0.75 0.78 τ0 [ms] 2.1 1.9 1.7 0.9 0.93 3.3 5.10 2.4 6.7 13.21 13.88 12.88 12.68 SR 5σ (cid:64)0.5(cid:48)(cid:48) 0.65 0.80 0.80 0.57 0.85 0.78 na 0.55 0.84 13.35 10.55 10.11 no no no no no 2.0 no no 2.0 Notes. Date, SPHERE observing mode, Total integration time (texp), Total field rotation (Rot), Coronagraph radius (rC), Neutral Density Filter, average DIMM seeing FWHM on source in V band (σ), the average coherence time τ0, the average Strehl ratio in H band (SR) coming from SPARTA data are presented. The deepest 5σ contrast limit reached at 0.5(cid:48)(cid:48)with IFS after spectral ADI is also given to show the relative quality of the data sets. number of PCA modes range from 2 to 5. There is no frame selection to minimize the self-subtraction of point- like sources when deriving the principal components. The principal components are calculated for each spectral chan- nel independently. Each frame is then projected onto the first 2 to 5 components to estimate the speckle contamina- tion. For TLOCI reduction, the minimum residual flux of a putative companion, because of self subtraction, is at least 15% of the candidate flux. The minimum radius where the speckle are calibrated and subtracted is 1.5 FWHM using annular sections. The maximum number of frames to es- timate the speckles is 80, no singular value decomposition cutoff was used. The speckle contamination is estimated by linear combination of frames that minimizes the residual en- ergy in the considered region using the bounded variables least-squares algorithm by Lawson & Hanson (1995). The contrast curves calculation assumes a flat spectra for the companion candidates. Although Specal Software can perform spectral differ- ential imaging (SDI, Racine et al. 1999), we use pro- cedures that treat each individual spectral channel sepa- rately. We also performed Reference Star Differential Imag- ing (RDI), subtracting to each datacube frame a reference PSF obtained from the star observed during the same or a close night, with the same set-up that is the most simi- lar, in terms of luminosity and noise model, to our target. HD 95086 is ideal for our purpose since the debris disk and close companion around this star are faint (Chauvin et al. 2018) and do not impact on the photometry of HD 100546. The RDI was performed using as reference the observations of HD 95086 taken on May 3rd, 2015. In addition, we used a set of customized data analysis procedures set up for performing the monochromatic PCA, taking into account the effect of the self-subtraction. 2.3. IRDIS PDI data reduction In IRDIS PDI observations, the stellar light is split into two beams with perpendicular polarization states. A half-wave plate allows to shift the orientation of the polarization four times by 22.5◦ in order to obtain a full set of polarimetric Stokes vectors. The data presented in this work are reduced following the double difference method (Kuhn et al. 2001) as described by Ginski et al. (2016). The resulting Q and U parameter are finally combined to obtain the polar Stokes Article number, page 4 of 19 vector Qφ and Uφ as from: Qφ = +Q cos (2φ) + U sin (2φ) (1) Uφ = −Q sin (2φ) + U cos (2φ) (2) where φ is the position angle of the location of interest (Schmid et al. 2006). With these definitions, positive Qφ values correspond to azimuthally polarized light, while neg- ative signal is radially polarized light. Uφ contains all signal with 45◦ offset from radial or azimuthal. 3. The Disk 3.1. Disk morphology Intensity Images Figure 2 shows the central part of the HD 100546 disk as obtained applying the RDI technique both for IFS and IRDIS. The light distribution appears elliptical, oriented in agreement with PA=146◦ and the ratio between major and minor axis is consistent with an inclination of 42◦ as found by Ardila et al. (2007) using HST/ACS observations of HD 100546. The South-East (SE) part in both images ap- pears brighter than the North-West (NW) part. The South- West (SW) part of the disk along the minor axis represents the near side to the observer and shows a strong depletion in the light distribution at a separation of > 0.2(cid:48)(cid:48). The two wings detected in the H2H3 and K1K2 bands by Garufi et al. (2016) are also distinguishable at IFS shorter wave- lengths, especially the northern one. Hereafter, we will keep Garufi et al. (2016) nomenclature for these two structures since they appear symmetric with respect to the minor axis, and do not have the same concavity, as expected in the case of multiple spirals disk. In these images we do not detect the spiral arm seen by Avenhaus et al. (2014), but in the IRDIS K1K2 data there is a thin spiral arm like structure, almost circular, at separation of 0.2(cid:48)(cid:48), from PA∼ 100◦ to PA∼ 4◦ that is barely visible also in the IFS images. No relevant features are detected at separations greater then ∼ 0.8(cid:48)(cid:48) in both these images, so that they are masked. The bright ring seen in IFS data at a separation of 0.65(cid:48)(cid:48) cor- responds to the SPHERE AO control radius2. The control 2 Within r ≈ 20 × λ/D (with 20 being half the number of deformable mirror actuators along a side) the SPHERE AO sys- E. Sissa et al.: High-Contrast study of the candidate planets and protoplanetary disk around HD 100546 Fig. 2. May 2015 IFS collapsed YH (left) and IRDIS K1K2 (right) results for HD100546 using RDI. Both images are r2-scaled to also enhance structures at larger radii. The grey circle in the left image indicate the SPHERE AO control radius at r ≈ 20 × λ/D. row, we report for radius at the wavelengths of IRDIS K1K2 bands is at a separation of 1.1(cid:48)(cid:48), therefore not visible in Fig 2. Figure 3 shows the results of the cADI analysis of the SHINE observations of HD 100546, where several disk structures are well visible. the first In the first time SPHERE/IFS images of the disk-bearing HD 100546. YJ collapsed image (Fig. 3a) shows the two wings described in Garufi et al. (2016) with a higher angular resolution that allows to detect an additional curved structure in the NW side, between 284◦ and 350◦, at separation of about 0.3(cid:48)(cid:48) and a southern spot. IFS also recovered the SE fainter arm. The IFS YH collapsed image (see Fig. 3b) reveals a single wider northern wing that barely reaches the expected po- sition of CCb. The prominent southern wing extends up to 0.7(cid:48)(cid:48) and PA=90◦. Two additional spiral arm-like features are also clearly visible. Referring to Follette et al. (2017), that provide a clear description of the innermost structure of HD 100546, these two arms may be identified with "S4" and "S6"; the southern one corresponds to the SE arm. In our image, an additional arm that lies above the southern wing, and starts from it, seems to extend in the East direc- tion up to ∼0.6(cid:48)(cid:48) and PA=90◦(E arm in Fig. 3b) and can be properly reproduced with a logarithmic spiral with a large pitch angle (∼ 50◦). There is also a hint of the "S2" spiral structure observed by Follette et al. (2017), E to the north- ern wing. No bright PSF-like knots are recovered along the arms at these wavelengths. We confirm the presence of a dark area just offset from the wings in the West direction, up to ∼ 0.3(cid:48)(cid:48), also detected in the RDI, where the disk shape is less altered thanks to a negligible self subtraction. tem efficiently suppresses the PSF down to the level of a residual stellar halo. At this radius, the AO-corrected image may show a circular artifact which is highlighted by the spatial filtering. This dark area is also present in the IFS YJ images where it looks less deep. In Fig. 3c the non-coronagraphic YH col- lapsed image clearly reveals the two bright wings and the SE arm and it appears clear that the distribution of light is uninterrupted between the wings, with only a small de- pletion along the minor axis that is likely an artifact of the ADI analysis: the two wings appear then as a unique struc- ture, symmetric to the minor axis, whose central part lies very close to the star. The darker area in the West is still visible. In the bottom row of Fig. 3 we present IRDIS TLOCI images. In Fig. 3d and 3e, we show the median H2H3 and K1K2 images, respectively. The two wings, the SE arm, the East arm and the dark area appear clearly visible in both images, but show some differences with respect to shorter wavelengths. A diffuse signal is visible on the top of the northern wing in H2H3 and also in K1K2, but is less lu- minous. This is located close to the expected position for CCb (Quanz et al. 2015; Currie et al. 2015). The East arm extends up to r=0.7(cid:48)(cid:48) and PA=64◦ in the H2H3 filter pair but only to r ∼0.4(cid:48)(cid:48) and PA∼ 90◦ in K1K2. In the H2H3 image, a second dark area is visible into the west side start- ing at ∼ 0.6(cid:48)(cid:48) till ∼ 0.9(cid:48)(cid:48). At larger separations, at least two spiral arms are visible in the south, ranging from ∼ 1.1(cid:48)(cid:48) to ∼ 2.3(cid:48)(cid:48). We notice that the SPHERE data were obtained under worst seeing conditions than GPI ones. The biggest differ- ence in the disk structures concerns the dimension of the "S4" spiral (Follette et al. 2017) which looks more extended in the GPI observations than in the SPHERE data. More- over, a few differences between GPI and SPHERE observa- tions can be noted on the extended structures around the location of the candidate companion CCb. We will discuss this in Sect. 4. Article number, page 5 of 19 NEsemi-major axissemi-minor axisWingsSpiral armNEsemi-major axissemi-minor axisWingsSpiral arm A&A proofs: manuscript no. hd100546Sissa Fig. 3. ADI images obtained for HD 100546. Top: IFS cADI images: (a) YJ from June 15; (b) YH from May15; (c) JH no coronagraphic image from Apr 16; Bottom: IRDIS TLOCI images (d) H2H3 from June 2015; (e) K1K2 from May 2015; (f) K1K2 no coronographic image from Feb 17. In all the images North is up, Est is left. Polarimetric Images The polarimetric images resulting from the data reduction described in Sect. 2.3 are shown in Fig. 4. We found that the outer disk (> 1(cid:48)(cid:48)) is well imaged in the short exposure in the J band. The K-band image is instead affected by thermal background in the detector, which prevents us to obtain similarly good image at larger radii as in the J-band. The Qφ image of the whole system is shown in Fig. 4a, whereas the respective Uφ is shown in Fig. 4b. Strong signal is detected from both images. In particular, the Uφ image shows positive values to North and South and negative val- ues to East and West. Four prominent stellar spikes are left from the data reduction in both images. A possible expla- nation for these artifacts is related to the bright (J=6.42 mag) central star polarization degree that makes the spikes not perfectly cancelled throughout the process of subtrac- tion of the beams with different polarization states. There is indeed a compact polarized inner ring (0.24-0.7 au, Panić et al. 2014, unresolved from the star in our images) that can contribute, even if the near-IR excess originating from that region is not particularly large for this source and this effect is not appreciable in other stars with similar sub-au disks. Article number, page 6 of 19 The circumstellar disk is clearly visible in Fig. 4a. To image a larger flux range, the images are shown in logarith- mic scale. In the image, a number of wrapped arms at radii 1.5(cid:48)(cid:48)-3(cid:48)(cid:48) from the central star are visible at all azimuthal an- gles. In particular, we recover the arms to South imaged by Boccaletti et al. (2013) and Garufi et al. (2016) and those to North visible from the Hubble Space Telescope (HST) image by Ardila et al. (2007). To highlight elusive disk features from Fig. 4a, we ap- plied an unsharp masking to the image similarly to what is described by Garufi et al. (2016). This technique consists in subtracting a smoothed version of the original image to the image itself. The result of the unsharp masking, obtained by smoothing Fig. 4a by ∼ 10 times the angular resolution of the observations, is shown in Fig. 4c. Looking at this im- age, it is possible that the majority of the structures visible from Fig. 4a are part of a unique arm which is wrapping for at least 540◦, as indicated by the green line in Fig. 4d and as already noticed by Garufi et al. (2017). Another bright arm visible to South seems to have a similar origin but to extend Eastward with a larger aperture angle (see the cyan line of Fig. 4d). knotcurved structureSE armE armSE wingNW wingCCbCCbCCbCCbCCbCCb E. Sissa et al.: High-Contrast study of the candidate planets and protoplanetary disk around HD 100546 Fig. 4. Polarized light images of HD100546 from SPHERE/IRDIS. (a): The Qφ image in the J band. (b): The Uφ image in the J band. (c): Unsharp masking of the Qφ image in the J band (see text). The inner box defines the region displayed in the bottom row. (d): Labeled version of (c). The inner inset circle is from (g). (e): Inner detail of the Qφ image in the J band. (f): Inner detail of the Qφ image in the K band. (g): Unsharp masking (see text) of (e). The grey circle indicates the IRDIS control radius at r ≈ 20 × λ/D (see text). (h): Unsharp masking (see text) of (f). Features visible from both (g) and (h) are labeled. The purple dot indicates the location of CCb from Quanz et al. (2013). In all images, the central star is in the middle of the grey circle, symbolizing the instrument coronagraphic mask. The logarithmic color stretch is relatively arbitrary and refers to positive values, except in (a) and (b) where it is the same. North is up, East is left. In the inner zone of the disk, inside ∼ 1(cid:48)(cid:48), the polarized flux is dramatically higher, even though two regions with reduced flux (appearing black in the image) are visible at P.A.∼ 160◦ and P.A.∼ 300◦ (Fig. 4a), similarly to what found by Quanz et al. (2011) and Avenhaus et al. (2014). Fig. 4e and of Fig. 4f show the zoomed Qφ image in J and K band of the innermost disk region. Two bright lobes are visible to the SE and to the NW, with the former be- ing brighter than the latter. Inward of these lobes, the disk cavity is marginally visible just outside of the instrument coronagraphic mask. Similarly to Fig. 4a, disk features are not easily recognizable. Therefore, we also apply an unsharp masking to these images, with smoothing by ∼ 6×FWHM, which results in Fig. 4g and Fig. 4h respectively. In both images, a number of possible features are visible. Among them, we only give credit to those that are persistent across unsharp masking procedures (i.e., by varying the smoothing factor) and, more importantly, across wavebands. Features that are radially distributed at r ≈ 20× λ/D (∼ 60 au) are masked out in Fig. 4g, because this region corresponds to the SPHERE control radius (see Fig. 4g). Features present in both the J and K band images are labeled in red in Fig. 4h. The morphology of the identified structures resem- bles the shape of the arms at larger radii and is consistent with the known disk geometry. The innermost of these fea- tures was also detected by Avenhaus et al. (2014) and Garufi et al. (2016), whereas the others were not. Interestingly, CCb as from Quanz et al. (2013) lies in correspondence of one of these possible features. From this data set, we cannot infer whether a spatial connection between the inner arms (in red) and the outer arm (in green) exists. 3.2. Comparison between ADI and PDI While the southern spiral arms of the outer disk (> 1(cid:48)(cid:48)) in the IRDIS H and K images are well detected also in the IRDIS PDI J (Fig. 3d-e and Fig. 4a), the central part of the PDI data do not show any axis-asymmetric structure. Using the unsharp masking technique on PDI data, in the inner- most region we found out three spiral arms (Fig. 4h) that have no counterpart in the intensity image. To compare ADI results with the PDI ones, we have then performed pseudo- ADI on the IRDIS DPI J-band images as follows: we use the parallactic angle values of the three different IRDIFS_EXT observations to simulate a simple cADI analysis based on the IRDIS DPI image. The results are given in Fig. 5: the simulation generates an image that is in better agreement with the IRDIS cADI images. On the other hand, PDI im- ages are directly comparable with RDI images. These two methods allow to better study the light distribution in the disk without self subtraction effects. 3.3. Disk spectrum In general, disk flux in the NIR is dominated by scattered light, while the brightness of a young planet is dominated by thermal emission. Therefore, to distinguish between a self shining planet and the disk we studied the total spectrum of the HD 100546 disk. Previous attempts to derive the spectrum of the disk of HD 100546 were done by Mulders et al. (2013) using HST data, and by Stolker et al. (2016) using data from NACO at VLT. Article number, page 7 of 19 0280.00190.0320.0630.0930.120.150.180.210.240.2250.380.50.630.760.8811.11.31.41240.390.530.680.830.981.11.31.41.61250.380.50.630.760.8811.11.31.411″QΦUΦ1″1″QΦ0280.00190.0320.0630.0930.120.150.180.210.240.21″Filtered QΦ1″(a)(b)(e)(c)(g)(d)Filtered QΦFiltered QΦJ band(f)(h)1″1″1″J bandK bandJ bandK bandFiltered QΦQΦb A&A proofs: manuscript no. hd100546Sissa Fig. 5. IRDIS PDI J band Qφ image (left) and the results of a simulated cADI analysis on this image. Overplotted in cyan the isophotes of IRDIS cADI H2 observations. the final contrast spectrum (all data sets considered have correlation coefficient > 0.98 with that of HD 100546). The relative flux spectrum is shown in Fig. 6. Since we are interested in the disk albedo, the flux calibration of the disk is based on broad-band photometry. The uncertainties on each wavelength (blue bars) were estimated as the stan- dard deviation of the mean of the spectra for individual pixels. Moreover we also show the intensity 1σ range over the population of individual pixels (light blue area). The re- gions between 1.08-1.15 µm and 1.35-1.48 µm are affected by strong water absorption due to the Earth atmosphere that cannot be perfectly recovered during the RDI process and were therefore removed. The difference in contrast magnitude is close to 0 be- tween H and K bands (K-H ∼ 0.03, in agreement with similar analysis presented by Avenhaus et al. (2014) ex- ploiting NACO H, Ks and L(cid:48) filters), and the spectrum appears featureless, with the disk reflecting about 8% the stellar light in the Y band and about 11% in K band. This is very similar to the spectrum obtained by Mulders et al. (2013). We tried to represent this spectral distribution as- suming light reflected by dust with a constant albedo. Mul- ders et al. (2013) suggested that in the near-infrared regime, light scattered off the HD 100546 outer disk surface is not only scattered stellar light, but the contribution of the light reprocessed by the inner disk (0.24-0.7 au, Panić et al. 2014) is not negligible. Therefore, we took into account a possible reddening due to the presence of the inner disk that can absorb stellar light. This was done following Cardelli et al. (1989). The free parameters are then reddening and the (total) disk reflectivity. The first is related via extinction (AV) and reddening (RV) to the slope of the spectrum, the latter is the combination of the albedo and the disk emit- ting area defined above3. The best solution gives AV=2.11 and albedo A=0.96 that is very high and not credible. An 3 Using the model derived in Sec. 3.4, the disk height at 40 au corresponds to 7.7◦ above the disk mid-plane. Fig. 6. Disk mean spectrum of HD 100546 along the wings. The blue dots and the shaded regions refers to the median value and its 3σ uncertainty. The green line represent the best fit obtained with light reflected by dust with constant albedo A = 0.65. In the bottom-right inset, the area selected for the spectrum extraction on the RDI images is delimited in blue. In order to derive the disk spectrum, we estimated the ratio between the total disk flux and the stellar one. We used the RDI datacube obtained with IFS in May 2015, where self subtraction is less aggressive. The target to ref- erence intensity ratio is measured in a specific region of the image, defined as the pixels of the wavelength-combined RDI image with a value above a determined threshold. This procedure yielded an elliptical region with deprojected ra- dius ∼ 40 au, without the central region that corresponds to the coronagraph, as shown in the small inset in Fig. 6. An adequately rescaled mask, in order to sample the same area of the disk, was also applied to the IRDIS RDI frames. Sev- eral reference stars were tested to confirm that the choice of the HD 100546 data set and the reference star do not affect Article number, page 8 of 19 -451526405983115158215292 E. Sissa et al.: High-Contrast study of the candidate planets and protoplanetary disk around HD 100546 alternative more palatable explanation is that the albedo of dust is not constant with wavelength. A similar expla- nation was already suggested by Mulders et al. (2013) and Stolker et al. (2016). A variation of the scattering efficiency from 0.43 at 1 µm up to 0.63 in the K-band may explain the observations. This points toward rather large particle sizes. The reddened spectrum we derive was also seen in HD 100546 HST data and was explained by the effect of forward scattering of micron-sized particles (Mulders et al. 2013). Further analysis based on the PDI data confirmed the presence of these particles, determining the phase func- tion (Stolker et al. 2016). 3.4. Disk geometrical model The disk structure of HD 100546 has been investigated in different wavelengths regimes. ALMA observations suggest that the millimeter-sized grains are located in two rings: a compact ring centered at 26 au with a width of 21 au, and an outer ring with a width of 75 ± 3 au centered at 190 ± 3 au (Walsh et al. 2014). The same two-rings structure was suggested by Panić et al. (2014), who used MIDI/VLTI data to probe micron-sized grains. They also came to the conclusion that there is an additional inner disc that extends no farther than 0.7 au from the star. The gap is about 10 au wide and free of detectable mid-infrared emission. Much larger dust, rocks, and planetesimals are not efficient H-band emitters, and our data do not exclude their presence inside this gap. Analysis of the SED of HD 100546 shows that the radial extent of the gas is ∼ 400 au (e.g., Benisty et al. 2010). The presence of an extended disk is clearly indicated also by scattered light, NIR and sub-millimeter imaging (Pantin et al. 2000; Ardila et al. 2007; Quanz et al. 2013; Avenhaus et al. 2014; Currie et al. 2014; Garufi et al. 2016; Follette et al. 2017). To constrain some of the HD 100546 disk properties we built a simple geometrical model based on analytic func- tions, that can describe the inner part of the system, be- tween 10 and 200 au. This model is similar to that described by Stolker et al. (2016). The basic assumption of this model is that the disk photometry can be described by light scat- tering on a single surface. In our model, the scattering func- tion for total intensity images is described by a two compo- nents Henyey-Greenstein (HG) function (Henyey & Green- stein 1941) with coefficients as defined in Milli et al. (2017) for HR 4796. For PDI data we used the function given by Stolker et al. (2016), cf r. Eq. (8). The light from the star is reflected by the optically thick disk surface, that lies above the disk mid plane as described by the power law: (cid:18) r (cid:19)b H = c r0 [au], (3) where r0 corresponds to 1 au and c is in au. Gaps present in the disk are modelled by decreasing the disk height H to zero au. This does not mean that the gaps are really empty, but simply that we cannot model emission from these re- gions with our approximation. Assuming the disk geome- try as found from the MIDI data, which trace micron-sized grains, and the ALMA data, which trace millimeter-sized grains, we have also assumed for the scattered light traced by SPHERE the same disk geometry, being aware that dif- ferent wavelengths probe of course different grain size and that different grain size may occupy different physical loca- tion in the disk. Therefore we set the rings range between ∼10 and ∼40 au (here after ring 1) and then from ∼150 to ∼230 au (hereafter ring 2). The thin inner ring (hereafter ring 0) extending between 0.24 and 0.7 au is behind the SPHERE coronagraph and therefore it is not included. The positions of the inner and the outer edge of the ring 1 are well constrained by the observed radial profiles: the inner edge must be at least partly behind the coronagraph, otherwise the disk wall will be clearly visible on the semi minor axis radial profile on our images, while the outer edge is visible on the semi major axis. The flaring parameter b is derived by the flux ratio between the peak of the near and the far side of the semi minor axis while the scale factor c impacts the shadowing effect of the disk. We first compared our model to RDI images; we did it for IFS and applied the same model also to IRDIS. We found out that a disk with b=1.08 and c=0.13 best repro- duces the HD 100546 radial profiles (Fig. 7), quite similar to the τ=1 disk surface for H and K band derived by Stolker et al. (2016). The disk rings 1 and 2 extend between ∼ 12 au and ∼ 44 au and between ∼ 110 au and 250 au respectively, consistent with previous results based on ALMA observa- tions (e.g. Walsh et al. 2014). The model for the outer disk is however fainter than observed, suggesting a larger flaring at large separations, though it should be considered that the RDI images are not very accurate at very high contrast levels. However, this model is also in quite good agreement with the IRDIS PDI images in broad J band as shown in Fig. 8. We notice that the inner edge of the ring 2 and the outer edge of the ring 1 might correspond to resonances 3:2 and 1:2 with a hypothetical massive object located at 70 au radius orbit, that is not far from the observed location of CCb (see next Section). The resulting RDI images are shown in Fig. 9. The resid- uals clearly show the presence of the two wings, that are even more evident when applying to this model the ADI method. To show this, we constructed a data cube (x, y, lambda, time) made of the sum of the disk model and of the reference image used in our RDI analysis, that is the data set of HD 95086. This datacube was then processed by the same ADI routine used for the original HD 100546 datacube. In this case, the two bright wings visible in all the images in Fig. 3 can be reproduced only by assuming a much thicker disk, with a value of H ∼ 30 au at r ∼ 40 au. This is because the bright wings visible in the ADI images, that correspond to the edge of the disk, are described by an ellipse whose center is far from the star. A similar result has been found for several disks, a clas- sical example being HD 97048 (Ginski et al. 2016). As ex- plained in their Fig. 5, disk images that may be described by off-center ellipses indicate the presence of material located well above the disk mid-plane. This was also demonstrated by recent hydrodynamical simulations performed by Dong et al. (2016) coupled with simple radiative transfer mod- els. These simulations demonstrate that a giant planet can open a gap in a disk creating a ring that, when seen at a intermediate viewing angle, and reinforced by the ADI process as explained above, appears as two pseudo-arms placed symmetrically around the minor axis winding in op- posite directions. They conclude that these simulations can describe HD 100546 as well. They also predicted the pres- ence of a "dark lane" parallel to these disk pseudo-arms Article number, page 9 of 19 A&A proofs: manuscript no. hd100546Sissa Fig. 7. Radial profile along the semi-major and the semi-minor axis of HD 100546 RDI images (red dots) compared with the best fit model (green line) obtained for IFS and IRDIS. Fig. 8. Radial profile along the semi-major and the semi-minor axis oh HD 100546 PDI images (continuum line) compared with the best fit model (dashed line) IRDIS. that is due to the self shadowing by the disk. In the case of HD 100546, the elliptical light distribution may indicate the presence of a ring of material at ∼40 au, close to the outer edge of the intermediate disk, but at an height of ∼ 30 au on the disk plane. This is about four times the value obtained with Eq. 2, using the best value we obtained for b and c. The origin of this ring is not well clear; we remind however that the edge of the intermediate disk can be explained by a 2:1 resonance with a hypothetical massive objects located at about 70 au. In order to agree with PDI and RDI data, we have to as- sume that this material is optically thin, contributing only marginally to the total emission from the disk. However, the higher (about a factor 25) sensitivity provided by ADI allows its detection. We note that a multi-rings configuration similar to that observed in HD 100546 is also seen around other objects. We mention in particular HD 141569A, where rings and outer spiral arms were observed with HST in the visible (Augereau et al. 1999; Mouillet et al. 2001; Clampin et al. Article number, page 10 of 19 -100-500501000.00010.00100.01000.10001.0000-100-50050100sep [AU]0.00010.00100.01000.10001.0000Rel. IntensityMajor Axis IFSsouth eastnorth west-100-500501000.00010.00100.01000.10001.0000-100-50050100sep [AU]0.00010.00100.01000.10001.0000Rel. IntensityMinor Axis IFSfar sidenear side-200-10001002000.00010.00100.01000.10001.0000-200-1000100200sep [AU]0.00010.00100.01000.10001.0000Rel. IntensityMajor Axis IRDISsouth eastnorth west-200-10001002000.00010.00100.01000.10001.0000-200-1000100200sep [AU]0.00010.00100.01000.10001.0000Rel. IntensityMinor Axis IRDISfar sidenear side-200-1000100200-4-3-2-10-200-1000100200Sep (au)-4-3-2-10Rel.IntensityMajor Axis-200-1000100200-4-3-2-10MINOR AXIS POL-200-1000100200Sep (au)-4-3-2-10Rel.IntensityMinor Axis E. Sissa et al.: High-Contrast study of the candidate planets and protoplanetary disk around HD 100546 Fig. 9. From left to right, IFS RDI image, model and residuals. Image are displayed with the same linear color bar from 6 × 10−3 dex to 1 dex. North is up, East is left. 2003), and in the near-IR with NICI (Biller et al. 2015; Ma- zoyer et al. 2016) and SPHERE (Perrot et al. 2016). More- over, the face-on disk surrounding TW Hya (van Boekel et al. 2017) presents three rings and three gaps within ∼2(cid:48)(cid:48) from the central star. These features were identified using optical and near-infrared scattered light surface brightness distribution, observed with SPHERE. Multiple rings were also observed in the inclined system RX J1615.3-3255 (de Boer et al. 2016) combining both visible polarimetric im- ages from ZIMPOL with IRDIS and IFS in scattered light. Finally, we notice that planet sculpting a cavity is not a unique explanation to the gaps. In addition some models with planets located outside of the gap can reproduce the gaps, or a single planet can open, under particular circum- stances, multiple gaps (Dong et al. 2017). 4. The candidate planets 4.1. Detection limits Two candidate companions around HD 100546 have been proposed over the years (Quanz et al. 2013; Brittain et al. 2014; Quanz et al. 2015; Currie et al. 2015). We investigated IFS and IRDIS images to find signatures of the already pro- posed candidate companions or new signatures. None of the IFS images shows evidence of new candidate companions above threshold while in the IRDIS images seven objects were identified as background stars (see Appendix A). To properly discuss our data, we should consider the detection limits for point sources on our data. The IRDIS 5σ detection limits for point sources were obtained through the SpeCal software (Galicher et al. 2018): it estimates the noise level σ as the azimuthal standard deviation of the flux in the reduced image inside FWHM/2 wide annuli at increasing separations, and rescales it as function of the stellar flux derived from the off-axis PSF images, taking into account the transmission of the neutral density filter used to avoid the star image saturation. We evaluated the IFS 5σ contrast limits in the PCA images following the method described in Mesa et al. (2015). In summary, the algorithm determines for each pixel the standard deviation of the flux in a 1.5 × 1.5λ/D box, centred on the pixel it- self and divides it by the proper stellar flux (as described above). For a given separation, the 5σ is then computed as five times the standard deviation of the values obtained for all the pixels at that separation. Finally both IRDIS and IFS 5σ values are corrected for the algorithm throughput and for the small number statistics (Mawet et al. 2014). To this purpose, fake companions, ten times more luminous than the noise residuals in the final reduced image, were injected in the pre-processed frames at various separations from the star and the datacube is reduced as before. This is then repeated at several different position angles and the throughput final value is the average of fake planets flux de- pletion at the same distance, corrected for the coronagraph attenuation. In Fig. 10 (Top) we show the deepest 5σ contrast curves obtained with IRDIS and IFS for HD 100546 in the H band. The TLOCI-based analysis shows that IRDIS could reach a 5σ contrast > 12 mag at separation > 0.5(cid:48)(cid:48) and even 15 mag at separations of 1.5(cid:48)(cid:48). in both H2 and H3 bands The IFS deepest contrast is reached when applying SDI and PCA. However, all these approaches, widely used for isolated point-like sources, likely underestimate the con- trast limits in the presence of a bright disk such as that of HD 100546, because the disk flux enters both in the esti- mation of the background and in the attenuation effect and therefore a less aggressive approach (e.g. monochromatic PCA with low number of modes) is more suitable. We con- sidered the HD 95086 data set mentioned above to show what will be the expected contrast limit in absence of the disk (red dash-dotted line in Fig. 10 In this Figure, we also show the H magnitudes of the two candidate companions around HD 100546 from Currie et al. (2015) using GPI: the impact of the disk on the contrast limit for IFS images is not negligible at separation of both CCc (∼ 2.2 mag) and CCb (∼ 1.5 mag). We note here that CCc with a contrast of 7.12 mag should be nominally visible in the IFS May 2015 and May 2016 data sets, where the contrast limits at the candidate separation are 7.3 mag and 7.8 mag, respec- tively. CCb (∆H = 13.44, r ∼ 0.47(cid:48)(cid:48)), instead, is below the detection limit in the H band. However it would be above the detection limit if we consider the limits applicable to a star without a luminous disk. 4.2. CCc No obvious point-like structure is visible at the expected position of CCc in any of our IFS and IRDIS data sets Article number, page 11 of 19 2.40e-074.01e-058.00e-051.20e-041.60e-042.00e-042.40e-042.80e-043.20e-043.60e-044.00e-040.3"0.3"0.3" A&A proofs: manuscript no. hd100546Sissa Fig. 11. Intensity profile as a function of phase of the two wings at different epochs. Phase 0 correspond to the near side of the semi-minor axis and negative values correspond to the SE wing. The dashed profile correspond to a point like source located at the expected position of CCc crossing the ring with the contrast upper limit given by Currie et al. (2015). April 2016 data set was scaled by an arbitrary factor of 1.6 to overlap other profiles. star on the disk plane as detected by Brittain et al. (2014) should move by ∼ 15◦ (one resolution element at that sep- aration) during our campaign. The feature is clearly visible only when using not aggressive reduction methods, such as monochromatic PCA with only one component. A key challenge in interpreting these data is the effect of processing. At CCc's angular separation, parallactic angle motion is small (1¯2λ/D for our data) and self-subtraction due to processing is severe. Processing could anneal the disk wings to look like a point source, (a shock, a con- vergence of spiral features, etc.). Alternatively, processing might preferentially anneal a point source, making it ap- pear radially elongated and indistinguishable from the disk wings. Additionally, the proximity of CCc to the corona- graph edge when a mask was used might affect both GPI and SPHERE results. With the coronagraph removed, the brightest part of both southern and northern wings shift closer to the minor axis (Fig. 11). The SPHERE coron- agraph we used, indeed, is 0.03" smaller than that used in GPI H-band observations and this technical difference could justify some different results obtained with the two instruments. Finally, we should consider that the observa- tion taken by Currie et al. (2015) is not simultaneous to IFS and IRDIS observations and the sky conditions were better. This leaves open the additional possibility that we could not recover the same feature detected by Currie et al. (2015) because the orbital motion moved it behind the coronagraph or the circumstellar disk, during this interval of time. Investigating the nature of the claimed CCc further re- quires a detailed forward-model of both any point source and the disk over multiple data sets, that takes into ac- count the impact of the coronagraphs, and is beyond the scope of this paper. Fig. 10. Top: contrast curves for HD 100546 obtained in the HIF S band for the two best IFS datasets applying SDI+PCA and for IRDIS H2H3 dataset applying TLOCI. Bottom: The IFS May 2015 contrast curve shown above is compared with the re- sult obtained with 2 modes monochromatic PCA (orange curve) and with the contrast curve in the HIF S band for HD 95086 ap- plying SDI+PCA (in red dashed-dotted line). The planets con- trast values measured in the H band obtained by Currie et al. 2015 with GPI are reported in both panels. in all epochs. Using a different reduction approach, Garufi et al. (2016) identified a bright knot in the SE arm (r ∼ 0.120 ± 0.016(cid:48)(cid:48), P A ∼ 165 ± 6◦) at least in H band for the May 2015 data. On one hand, one could try to asso- ciate CCc with this knot, as it may lie at/just inside the disk cavity and is counterclockwise from the position noted by Currie et al. (2015), as expected for an orbiting object predicted from Brittain et al. (2013). On the other hand, these results do not establish CCc as a companion since there are equally compelling alternatives. We detect bright, elongated emission at a similar location in all our IFS and IRDIS H2H3 data set, suggesting confusion between bright disk emission and that of any point source. Moreover, all our data sets have a light distribution along the two wings that is nearly symmetric around the disk semi-minor axis, with the NW side being more luminous, as already found by Follette et al. (2017). The emission could simply be a non-polarized hot spot in the disk. Furthermore, this struc- ture is detected at locations compatible with a stationary object with respect to the star, despite a planet orbiting the Article number, page 12 of 19 0.51.01.52.0141210860.51.01.52.0Sep [arcsec]14121086contrast [mag]050100150200250Sep [au]IRDIS H2 TLOCIIRDIS H3 TLOCIIFS HIFS PCA2015 May2016 May0.51.01.52.0141210860.51.01.52.0Sep [arcsec]14121086contrast [mag]050100150200250Sep [au]HIFS SDI+PCAHIFS monoPCAHIFS SDI+PCA Ref.-60-40-2002040600.00000.00050.00100.0015-60-40-200204060Degree along ring from minor axis0.00000.00050.00100.0015Rel. IntensityMay 2015May 2016Apr 2016Feb 2017 E. Sissa et al.: High-Contrast study of the candidate planets and protoplanetary disk around HD 100546 et al. (2013, 2015); Currie et al. (2014) and the two best K1K2 epochs from SPHERE. We adopted the disk plane inclination of 42◦ and position angle of 146◦ (Pineda et al. 2014), obtaining an orbital radius of 545±15 mas that corre- sponds to 59.5± 2 AU at the distance of HD 100546. Given the stellar mass of 2.4 M(cid:12) (see e.g. Quanz et al. 2013), the corresponding period is 299 yr. We over-imposed the predictions for a similar orbit on the run of separation and position angles shown in Fig. 12, dashed lines. The agree- ment between expectation and observations is fairly good, in view of the rather large error bars associated with all these data. We conclude that the motion of CCb is com- patible with a circular orbit on the plane of the disk, in the same direction as the disk rotation, although there is room for different orbital solutions and also for stationari- ness, given the uncertainties on the positions. If this orbital solution is confidently assessed, we could therefore predict that in Spring 2019 CCb will have moved enough to disen- tangle the motion from being a stationary object, or it will disappear behind a disk structure. For example, assuming a nominal co-planar, circular orbit at 59.5 au, CCb should appear at separation of ∼ 440 mas and PA ∼ 17◦ in May 2019, which is roughly three resolution elements away from its position in May 2015. However, the astrometric points hint about a slower motion that can be due to an eccentric and/or no-coplanar orbit. If so, the time baseline needed to confirm the CCb movement stretches out. This object has a contrast of 12.08±0.49 mag in K1 and 11.68±0.60 mag in K2, compatible with the NACO non de- tection by Boccaletti et al. (2013), while its H band upper limit is 13.75 ± 0.05 mag. On the other hand, the IRDIS H2H3 data set and the H part of the IFS images (simulta- neous to the K1K2 IRDIS data sets) show a diffuse emis- sion, located ∼ 70 mas NW with respect to the previous one in continuation with the Northern wing, as shown in Fig. 13. This emission, with median position r = 477 ± 12 mas, P A = 7.2 ± 1.5◦, is consistent with the detection in the H-band by Currie et al. (2015) and Rameau et al. (2017). This source is clearly distinct from that detected at longer wavelengths, and looks extended. However, given the angu- lar resolution of GPI and the different H-band filter used in GPI and SPHERE, the detection of Currie et al. (2015, 2017), intermediate between the H and K detection with SPHERE, can be interpreted as a combination between these two. Finally, only very weak diffuse emission was vis- ible at wavelengths shorter than 1.1 µm. Additional information on CCb cannot be retrieved by the polarimetric data: no emission is seen in the IRDIS Qφ images as shown in Figure 4(h), while in the PDI+ADI image the residuals due of the telescope spider are not neg- ligible and fall at the location of CCb, so it is impossible to tell whether or not a disk structure at that location could be interpreted as a point source after filtering by ADI. 4.4. Interpreting CCb as extended source Previous works suggested that CCb is surrounded by cir- cumplanetary disk (Quanz et al. 2015; Currie et al. 2015; Rameau et al. 2017). In Quanz et al. (2015) the presence of a spatially unresolved circumplanetary disk (r ∼ 1.4 au around a 2MJ object) is considered as an explanation of the discrepancy between the observed values of radius and ef- fective temperature with those obtained by models for very young gas giant planets. Article number, page 13 of 19 Fig. 12. Run of deprojected separation (left panel) and of posi- tion angle (right panel) of CCb with time. Data are from Quanz et al. (2013); Currie et al. (2014); Quanz et al. (2015) (blue dots) and this paper (open diamonds). Dashed lines are predictions for a circular orbit on the plane of the disk. Table 2. Astromety of CCb in the SPHERE K band at different epochs. Date May 3 2015 May 31 2016 r[mas] 454 ± 10 456 ± 10 PA [deg] 10.4 ± 1.5 12.5 ± 1.5 4.3. CCb Since the initial discovery of HD100546 b by Quanz et al. (2013), its re-detection has been reported from different groups using different instruments and in multiple wave- lengths (Quanz et al. 2013; Currie et al. 2014; Quanz et al. 2015). However, the debate on the nature of CCb was re- cently reactivated by the results obtained by Currie et al. (2015) and Rameau et al. (2017), both with GPI H-band observations. In both cases, a clear signal is detected at a location possibly compatible with CCb, but, on one hand Currie et al. (2015, 2017) identified this with CCb (a point source over-imposed to a flat disk component), on the other Rameau et al. (2017) interpreted it as stellar scattered light, being point-like or extended depending on the processing method used. In the following discussion we use a not aggressive monochromatic PCA approach with only one component over the whole IFS FoV simultaneously, and over the inner circle of 0.8(cid:48)(cid:48) for IRDIS one. Moreover, we exploit the very wide spectral range simultaneously offered by the SPHERE IRDIFS_EXT set up (from 0.95 up to 2.2 µm). All IRDIS K1K2 datas ets show a clear diffuse emission on top of the Northern wing (as already noticed by Garufi et al. 2016 for the 2015 data sets) that is compatible with the CCb detections in L(cid:48) and M band using NACO. In order to obtain a more precise characterization of this feature, we focused on the two best IRDIFS_EXT data sets and we found out that the position of this source varies a little between the two epochs, as shown in Table 2. From the disk velocity retrieved from spectroastrometric analysis of different molecules (see e.g., Acke & van den Ancker 2006; Panić et al. 2010; Brittain et al. 2009) and/or with ALMA observations (Walsh et al. 2014), it was derived that the disk rotates counter-clock wise, with the SW part being the closest to the observers. We discover that the feature we detect is moving counter-clockwise and its astrometry, combined with previous detections, is compatible with a Keplerian motion on the disk plane. This is shown in Fig. 12, where we present the run of the separation and posi- tion angle of this source with time, using data from Quanz 5500600065007000750080000.500.520.540.560.580.60550060006500700075008000JD - 24500000.500.520.540.560.580.60Sep ["]550060006500700075008000JD - 24500004648505254565860PA [deg] A&A proofs: manuscript no. hd100546Sissa Fig. 13. Comparison of HD 100546 images in different band filters. One component PCA images of IRDIS H2H3 taken on May 29th 2015 (left), K1K2 taken on May 3rd 2015 (middle) and their combination(right). Contours refer to IRDIS H2H3. The cross represents Currie et al. (2015) detection in GPI H band , the arrow in the central panel indicate the motion of CCb between Quanz et al. (2015) detection in NACO L(cid:48) band and our May 2016 K1K2 detection of CCb. The grey circle represent the resolution element of the images. Table 3. At different wavelengths, we report the star magnitude m∗, the contrast of CCb or it upper limit (contb) and the asso- ciated uncertainty (errcont). The values obtained by this work are not corrected for any dust extinction. λ [µm] 4.8 3.8 3.8 3.8 2.25 2.2 2.1 2.11 1.60 1.58 1.25 m∗ [mag] 4.13 4.2 4.52 contb [mag] 9.2 9.0 9.4 5.42 11.68 5.42 > 9.09 5.42 > 9.60 5.42 12.08 5.96 > 12.5 6.42 > 13.6 mb [mag] 13.33±0.16 13.2 ±0.4 13.92±0.1 13.06±0.51 17.10 ±0.60 > 14.51 > 15.02 17.50 ±0.49 19.40 ± 0.32a > 18.46 > 20.02 Reference 1 2 1 3 4 5 1 4 6 4 4 Notes. 1 Quanz et al. (2015), 2 Quanz et al. (2013), 3 Currie et al. (2014), 4 this work, 5 Boccaletti et al. (2013), 6 Currie et al. (2015). aThis value refers to the apparent magnitude of the source iden- tifyed by (Currie et al. 2015) and located at about 28 mas (∼ 0.7λ/D) from the expected position of the source detected at longer wavelengths. trast limits in the corresponding area of CCb, while in the H2H3 filter we only detect a very weak diffuse emission at its location; (b) what we see in the IRDIS data at positions corresponding to CCb could be simply a bright structure of the disk; (c) we are not subtracting the disk contribu- tion to the flux. Similar to what discussed above, the self subtraction on extended sources is not negligible even with this not aggressive reduction method and it is difficult to estimate because it depends on the specific distribution of light than cannot be easily modelled. This implies that the ADI technique alters the photometry of the extended ob- ject. What we can conclude is simply that the structure seen in our K1K2 images is compatible with an extended object whose apparent magnitude is brighter than what is expected for a point source at the same location and is com- Fig. 14. SED of CCb planet and disk combining previous results (black) with our upper limits (orange). The green line represent a black body of 932 K with an absorption AV = 28 mag (see text). The grey curve represent a grey contrast of 14 mag fitting the short wavelengths observations. In particular, only data cor- responding to black filled circles were considered for the relative astrometry. The black star symbol refers to apparent magnitude of a source located at about 28 mas (∼ 0.7λ/D) from the po- sition of the source detected at longer wavelengths, as given by Currie et al. (2015). At wavelengths shorter than 2 µm, we evaluated the 5σ limit to the magnitudes of a point source at the pre- sumed location of CCb. With IFS we obtain that it is fainter than the apparent magnitude 19.84 in the J band (1.20 − 1.30 µm), fainter than 19.80 mag in J broad band (1.15 − 1.35 µm) and fainter than 19.05 mag in H; uncer- tainties on these values came from the noise distribution estimation at the CCb location. Putting together all the results and the literature ones (see Tab. 3) we obtain the CCb SED plotted in Fig. 14. Some important arguments that one should take into ac- count interpreting this SED are: (a) CCb is below detection threshold in our IFS images, so we only estimated the con- Article number, page 14 of 19 1234522201816141212345λ [µm]222018161412App magnitude Prev. worksThis workT=798K, Av=0 mag E. Sissa et al.: High-Contrast study of the candidate planets and protoplanetary disk around HD 100546 patible, within the uncertainties, with the flux observed at longer wavelengths. Given the small difference of contrast between the J/H- band diffuse source and the strong difference with the L/M- band putative point-source, we may represent this emission as the sum of two different sources: a very red compact source and a more extended one with a flatter spectrum. It is clear in fact that this source is redder than the star and cannot be explained as stellar scattered light alone, since a flat contrast (grey line) that fits the short wave- length points largely fails to reproduce the L and M data. The green line represents the SED of a black body with Tef f ∼ 800 K and no absorption, that implies a radius of the emitting area R = 12.5 RJ. Taking into account the new determination of HD 100546 distance, this result is in fair agreement with Quanz et al. 2015 (Tef f = 932+193−202 K, R = 6.9+2.7−2.9 RJ). It is noteworthy that, given the small number of photometric points and limited spectral range, absorption and temperature are degenerate, and therefore many combination of temperature and reddening can well reproduce the SED. However, given its very young age, the object likely does not emit as a black-body, but the accret- ing disk shock contribution to the total luminosity is not negligible, and could also be dominant (Mordasini et al. 2017). Also the filter used are built to provide information on absorbing molecules and therefore can play a role in the flux estimation at given wavelengths. The luminosity of a forming and accreting planet in dif- ferent formation scenarios is evaluated in Mordasini et al. (2017). They analyzed the case of HD 100546 b in detail, starting from the physical parameters derived by Quanz et al. (2015), and could constrain the mass of this source only with large uncertainties, due to different mechanisms allowed. To disentangle the emitting sources and hence es- timate the CCb mass, we need to better characterize the SED, that can be possible with space observations with new facilities (like JWST) or with the ELT class instruments. Alternatively, a better spatial resolution will allow to re- solve the circumplanetary region and retrieve the mass of CCb through dynamical models. The nature of the extended emission detected in the H2H3 filter is not clear. It can be the either a source phys- ically bound to CCb, or a bright structure of the disk. Fur- ther deep observations and/or a longer (3-4 yr) temporal coverage can disentangle between these possibilities. If we assume that they are an unique extended source, we obtain that the radius is of the order of r=34 mas ((cid:39) 3.7 au), compatible with previous estimations. If this were a cir- cumplanetary disk, the corresponding Hill's radius (RHill) of the planet would be expected to be about 3 times larger, following models by e.g. Shabram & Boley (2013); Ayliffe & Bate (2009); Quillen & Trilling (1998). Therefore, the observed structure is compatible with a circumplanetary material around a massive planet that could be responsi- ble for carving the gap as demonstrated by hydrodynamical models (e.g. Pinilla et al. 2015; Dong et al. 2016). To justify the extended emission, we further consider a spherical cloud around CCb, located at a separation d=65± 5 au from the star, that reflects stellar light. If its optical depth is τ >> 1, then we expect that the total reflected light is c=Aπr2/(4πd2) where A is the albedo. If A=0.5 then the expected contrast is c=4.2×10−4, that corresponds to 8.6 mag, in agreement with our observations. This feature would be unresolved at L and M wavelength and therefore contributes to the flux of the compact source but it should be far less than the contribution due to the companion: its luminosity is compatible within the uncertainties with reflected stellar light from the J to the K band. We finally notice that, in this scenario we are not consid- ering the absorption due to the circumstellar disk material between the star and the circumplanetary disk. This de- pends on the thickness of the circumstellar disk and on the height of the planet over the circumstellar disk plane at the epoch of our observations. This effect, and the presence of shadows due to the circumstellar disk that may affect the amount of light incident on a circumplanetary disk, could be not negligible and possibly cause an irregular illumina- tion of the circumplanetary disk. Of course the truth can be a combination of all these factors. We conclude that the origin of this emitting area is still unclear. While its appearance and the SED are compatible with a highly reddened substellar object surrounded by a dust cloud, we cannot exclude other interpretations, such as the super imposition of two spiral arms, the northern wing and the small IRDIS North arm (see Fig. 3), or disk material flowing to a planet due to its perturbation induced on the disk. 5. Conclusion We observed HD 100546 with SPHERE using its subsys- tems IRDIS and IFS in direct imaging and in polarime- try. Our observations confirm the presence of a very struc- tured disk and reveal additional features. The different post processing techniques reveal different characteristics of this complicated disk. RDI and PDI images are dominated by the almost symmetric intermediate and outer disks, while more aggressive differential imaging technique tell a differ- ent story, featuring strongly de-centred rings and spirals. These two views can be reconciled in a picture as follows: -- The two bright wings, dominant structures in the IR at separations closer than 500 mas, are a unique structure. This is quite evident from the non coronagraphic im- ages. The presence of the coronagraph and the use of the ADI technique contributed to cancel out the light in the rings region closest to the star. -- The new PDI data confirms the presence of a unique arm warping for 540◦. In the innermost regions, three small spiral arms are detected in both J and K band. -- Modelling a geometrical representation of the disk cou- pled with an analytic scattering function, we obtain that the disk rings 1 and 2 extend between ∼ 15 au and ∼ 40 au and between ∼ 110 au and 250 au respectively, consistent with previous results. The inner edge of the ring 2 and the outer edge of the ring 1 correspond to resonances 3:2 and 1:2 with a 70 au radius orbit, sug- gesting the presence of a massive object located at that separation. -- We do not exclude the presence of additional spiral arms inside the disk rings. In particular, we confirm detec- tion of the two possible spiral arms East and South of HD 100546 previously identified by Follette et al. (2017). -- The spectrum of this disk does not show obvious evi- dence for segregation of dust of different size and is well explained by micron sized particles. Article number, page 15 of 19 For what concerns the planets, we have no clear evidences of the CCc detected by Currie et al. (2015); processing could both cause the disk wings to look like a point source or anneal a point source to be indistinguishable from disk emission. This does not exclude that the planet has moved behind the disk in the time between Currie et al. (2015) observations and the time SPHERE data were acquired. We identify a spatially diffuse source in K band broadly consistent with CCb. When combined with previous mea- surements, its photometry is consistent with a blackbody- like emitting source of ∼ 800 K, compatible with a highly reddened massive planet or brown dwarf surrounded by a dust cloud or its circumplanetary disk. Its astrometry might have revealed evidence for orbital motion, a result that can be confirmed with future observations. This object can in- deed be the disk perturber suggested by the disk modelling. However, other hypothesis are also possible, such as the overimposition of two spiral arms at the location of the L' and M' detections. Acknowledgements. The authors thank the anonymous referee for a very constructive referee report that improved the initial manuscript. The authors thank the ESO Paranal Staff for support for conduct- ing the observations. The authors thank Sascha Quanz, Adriana Pohl and Tomas Stolker for the very useful comments that improved a lot the quality of the paper. E.S., R.G., D.M., S.D. and R.U.C. ac- knowledge support from the "Progetti Premiali" funding scheme of the Italian Ministry of Education, University, and Research. E.R. is supported by the European Union's Horizon 2020 research and in- novation programme under the Marie Skłodowska-Curie grant agree- ment No 664931. This work has been supported by the project PRIN- INAF 2016 The Cradle of Life - GENESIS-SKA (General Conditions in Early Planetary Systems for the rise of life with SKA). The au- thors acknowledge financial support from the Programme National de Planétologie (PNP) and the Programme National de Physique Stel- laire (PNPS) of CNRS-INSU. This work has also been supported by a grant from the French Labex OSUG@2020 (Investissements d'avenir - ANR10 LABX56). The project is supported by CNRS, by the Agence Nationale de la Recherche (ANR-14-CE33-0018). This work is partly based on data products produced at the SPHERE Data Centre hosted at OSUG/IPAG, Grenoble. We thank P. Delorme and E. Lagadec (SPHERE Data Centre) for their efficient help during the data reduc- tion process. SPHERE is an instrument designed and built by a con- sortium consisting of IPAG (Grenoble, France), MPIA (Heidelberg, Germany), LAM (Marseille, France), LESIA (Paris, France), Labo- ratoire Lagrange (Nice, France), INAF Osservatorio Astronomico di Padova (Italy), Observatoire de Genève (Switzerland), ETH Zurich (Switzerland), NOVA (Netherlands), ONERA (France) and ASTRON (Netherlands) in collaboration with ESO. SPHERE was funded by ESO, with additional contributions from CNRS (France), MPIA (Ger- many), INAF (Italy), FINES (Switzerland) and NOVA (Netherlands). SPHERE also received funding from the European Commission Sixth and Seventh Framework Programmes as part of the Optical Infrared Coordination Network for Astronomy (OPTICON) under grant num- ber RII3-Ct-2004-001566 for FP6 (2004-2008), grant number 226604 for FP7 (2009-2012) and grant number 312430 for FP7 (2013-2016). References Acke, B., & van den Ancker, M. E. 2006, A&A, 449, 267 Amara, A., & Quanz, S. P. 2012, MNRAS, 427, 948 Ardila, D. R., Golimowski, D. A., Krist, J. E., et al. 2007, ApJ, 665, Augereau, J. C., Lagrange, A. M., Mouillet, D., & Ménard, F. 1999, 363 A&A, 350, L51 -- . 2001, A&A, 365, 78 Avenhaus, H., Quanz, S. P., Meyer, M. R., et al. 2014, ApJ, 790, 56 Ayliffe, B. A., & Bate, M. R. 2009, MNRAS, 397, 657 Benisty, M., Tatulli, E., Ménard, F., & Swain, M. R. 2010, A&A, 511, 512 A75 Beuzit, J.-L., Feldt, M., Dohlen, K., et al. 2008, in Proc. SPIE, Vol. 7014, Ground-based and Airborne Instrumentation for Astronomy II, 701418 A&A, 372, L61 2013, A&A, 549, A112 Article number, page 16 of 19 A&A proofs: manuscript no. hd100546Sissa Biller, B. A., Males, J., Rodigas, T., et al. 2014, ApJ, 792, L22 Biller, B. A., Liu, M. C., Rice, K., et al. 2015, MNRAS, 450, 4446 Boccaletti, A., Pantin, E., Lagrange, A.-M., et al. 2013, A&A, 560, A20 Boccaletti, A., Abe, L., Baudrand, J., et al. 2008, in Proc. SPIE, Vol. 7015, Adaptive Optics Systems, 70151B Bouwman, J., de Koter, A., Dominik, C., & Waters, L. B. F. M. 2003, Brittain, S. D., Carr, J. S., Najita, J. R., Quanz, S. P., & Meyer, M. R. A&A, 401, 577 2014, ApJ, 791, 136 Brittain, S. D., Najita, J. R., & Carr, J. S. 2009, ApJ, 702, 85 Brittain, S. D., Najita, J. R., Carr, J. S., et al. 2013, ApJ, 767, 159 Carbillet, M., Bendjoya, P., Abe, L., et al. 2011, Experimental As- Cardelli, J. A., Clayton, G. C., & Mathis, J. S. 1989, ApJ, 345, 245 Chauvin, G., Gratton, R., Bonnefoy, M., et al. 2018, ArXiv e-prints, tronomy, 30, 39 arXiv:1801.05850 Clampin, M., Krist, J. E., Ardila, D. R., et al. 2003, AJ, 126, 385 Claudi, R. U., Turatto, M., Gratton, R. G., et al. 2008, in Proc. SPIE, Vol. 7014, Ground-based and Airborne Instrumentation for Astron- omy II, 70143E Currie, T., Brittain, S., Grady, C. A., Kenyon, S. J., & Muto, T. 2017, Research Notes of the American Astronomical Society, 1, 40 Currie, T., Cloutier, R., Brittain, S., et al. 2015, ApJ, 814, L27 Currie, T., Muto, T., Kudo, T., et al. 2014, ApJ, 796, L30 Cutri, R. M., Skrutskie, M. F., van Dyk, S., et al. 2003, VizieR Online Data Catalog, 2246 de Boer, J., Salter, G., Benisty, M., et al. 2016, A&A, 595, A114 Delorme, P., Meunier, N., Albert, D., et al. 2017, in SF2A-2017: Pro- ceedings of the Annual meeting of the French Society of Astronomy and Astrophysics, ed. C. Reylé, P. Di Matteo, F. Herpin, E. La- gadec, A. Lançon, Z. Meliani, & F. Royer, 347 -- 361 Dohlen, K., Langlois, M., Saisse, M., et al. 2008, in Proc. SPIE, Vol. 7014, Ground-based and Airborne Instrumentation for Astronomy II, 70143L Dong, R., Fung, J., & Chiang, E. 2016, ApJ, 826, 75 Dong, R., Li, S., Chiang, E., & Li, H. 2017, ApJ, 843, 127 Follette, K. B., Rameau, J., Dong, R., et al. 2017, AJ, 153, 264 Fusco, T., Petit, C., Rousset, G., et al. 2006, in Proc. SPIE, Vol. 6272, Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series, 62720K Gaia Collaboration, Brown, A. G. A., Vallenari, A., et al. 2016, A&A, 595, A2 Galicher, R., Boccaletti, A., Mesa, D., et al. 2018, A&A, 615, A92 Garufi, A., Quanz, S. P., Avenhaus, H., et al. 2013, A&A, 560, A105 Garufi, A., Quanz, S. P., Schmid, H. M., et al. 2016, A&A, 588, A8 Garufi, A., Meeus, G., Benisty, M., et al. 2017, A&A, 603, A21 Ginski, C., Stolker, T., Pinilla, P., et al. 2016, A&A, 595, A112 Grady, C. A., Woodgate, B., Heap, S. R., et al. 2005, ApJ, 620, 470 Grady, C. A., Polomski, E. F., Henning, T., et al. 2001, AJ, 122, 3396 Henyey, L. G., & Greenstein, J. L. 1941, ApJ, 93, 70 Hu, J. Y., The, P. S., & de Winter, D. 1989, A&A, 208, 213 Jovanovic, N., Martinache, F., Guyon, O., et al. 2015, PASP, 127, 890 Keppler, M., Benisty, M., Müller, A., et al. 2018, ArXiv e-prints, arXiv:1806.11568 Kraus, A. L., & Ireland, M. J. 2012, ApJ, 745, 5 Kuhn, J. R., Potter, D., & Parise, B. 2001, ApJ, 553, L189 Levenhagen, R. S., & Leister, N. V. 2006, MNRAS, 371, 252 Macintosh, B., Graham, J. R., Ingraham, P., et al. 2014, Proceedings of the National Academy of Science, 111, 12661 Maire, A.-L., Langlois, M., Dohlen, K., et al. 2016, in Proc. SPIE, Vol. 9908, Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series, 990834 Marois, C., Correia, C., Véran, J.-P., & Currie, T. 2014, in IAU Sym- posium, Vol. 299, Exploring the Formation and Evolution of Plan- etary Systems, ed. M. Booth, B. C. Matthews, & J. R. Graham, 48 -- 49 Marois, C., Lafrenière, D., Doyon, R., Macintosh, B., & Nadeau, D. 2006, ApJ, 641, 556 Martinez, P., Dorrer, C., Aller Carpentier, E., et al. 2009, A&A, 495, Mawet, D., Milli, J., Wahhaj, Z., et al. 2014, ApJ, 792, 97 Mazoyer, J., Boccaletti, A., Choquet, É., et al. 2016, ApJ, 818, 150 Mesa, D., Gratton, R., Zurlo, A., et al. 2015, A&A, 576, A121 Milli, J., Vigan, A., Mouillet, D., et al. 2017, A&A, 599, A108 Mordasini, C., Marleau, G.-D., & Mollière, P. 2017, A&A, 608, A72 Mouillet, D., Lagrange, A. M., Augereau, J. C., & Ménard, F. 2001, Mulders, G. D., Min, M., Dominik, C., Debes, J. H., & Schneider, G. E. Sissa et al.: High-Contrast study of the candidate planets and protoplanetary disk around HD 100546 Müller, A., Keppler, M., Henning, T., et al. 2018, ArXiv e-prints, arXiv:1806.11567 Panić, O., Ratzka, T., Mulders, G. D., et al. 2014, A&A, 562, A101 Panić, O., van Dishoeck, E. F., Hogerheijde, M. R., et al. 2010, A&A, 519, A110 Pantin, E., Waelkens, C., & Lagage, P. O. 2000, A&A, 361, L9 Pavlov, A., Möller-Nilsson, O., Feldt, M., et al. 2008, in Proc. SPIE, Vol. 7019, Advanced Software and Control for Astronomy II, 701939 Perrot, C., Boccaletti, A., Pantin, E., et al. 2016, A&A, 590, L7 Pineda, J. E., Quanz, S. P., Meru, F., et al. 2014, ApJ, 788, L34 Pinilla, P., Birnstiel, T., & Walsh, C. 2015, A&A, 580, A105 Quanz, S. P., Amara, A., Meyer, M. R., et al. 2015, ApJ, 807, 64 -- . 2013, ApJ, 766, L1 Quanz, S. P., Schmid, H. M., Geissler, K., et al. 2011, ApJ, 738, 23 Quillen, A. C., & Trilling, D. E. 1998, ApJ, 508, 707 Quirós-Pacheco, F., Busoni, L., Agapito, G., et al. 2010, in Proc. SPIE, Vol. 7736, Adaptive Optics Systems II, 77363H Racine, R., Walker, G. A. H., Nadeau, D., Doyon, R., & Marois, C. 1999, PASP, 111, 587 Rameau, J., Follette, K. B., Pueyo, L., et al. 2017, AJ, 153, 244 Reggiani, M., Quanz, S. P., Meyer, M. R., et al. 2014, ApJ, 792, L23 Sallum, S., Follette, K. B., Eisner, J. A., et al. 2015, Nature, 527, 342 Schmid, H. M., Joos, F., & Tschan, D. 2006, A&A, 452, 657 Shabram, M., & Boley, A. C. 2013, ApJ, 767, 63 Soummer, R., Pueyo, L., & Larkin, J. 2012, ApJ, 755, L28 Stolker, T., Dominik, C., Min, M., et al. 2016, A&A, 596, A70 Thalmann, C., Schmid, H. M., Boccaletti, A., et al. 2008, in Proc. SPIE, Vol. 7014, Ground-based and Airborne Instrumenta- tion for Astronomy II, 70143F Thalmann, C., Grady, C. A., Goto, M., et al. 2010, ApJ, 718, L87 van Boekel, R., Henning, T., Menu, J., et al. 2017, ApJ, 837, 132 van der Plas, G., van den Ancker, M. E., Acke, B., et al. 2009, A&A, 500, 1137 Vigan, A., Moutou, C., Langlois, M., et al. 2010, MNRAS, 407, 71 Walsh, C., Juhász, A., Pinilla, P., et al. 2014, ApJ, 791, L6 Wright, C. M., Maddison, S. T., Wilner, D. J., et al. 2015, MNRAS, 453, 414 Zurlo, A., Vigan, A., Mesa, D., et al. 2014, A&A, 572, A85 1 INAF-Osservatorio Astronomico di Padova, Vicolo dell'Osservatorio 5, I-35122, Padova, Italy 2 Dipartimento di Fisica e Astronomia - Universita' di Padova, Vicolo dell'Osservatorio 3, I-35122, Padova, Italy 3 Institute for Particle Physics and Astrophysics, ETH Zurich, Wolfgang-Pauli-Strasse 27, CH-8093 Zurich, Switzerland 4 Universidad Autonóma de Madrid, Dpto. Física Teórica, Módulo 15, Facultad de Ciencias, Campus de Cantoblanco, 28049, Madrid, Spain 5 Núcleo de Astronomía, Facultad de Ingeniería, Universidad Diego Portales, Av. Ejercito 441, Santiago, Chile 6 Aix-Marseille Université, CNRS, LAM (Laboratoire d'Astrophysique de Marseille) UMR 7326, 13388, Marseille, France 7 University of Atacama, Copayapu 485, Copiapo, Chile 8 CRAL, UMR 5574, CNRS, Université de Lyon, Ecole Nor- male Suprieure de Lyon, 46 Alle d'Italie, F-69364 Lyon Cedex 07, France 9 Leiden Observatory, Leiden University, PO Box 9513, 2300 RA Leiden, The Netherlands 10 Université Grenoble Alpes, CNRS, IPAG, 38000 Grenoble, 11 Max-Planck-Institut für Astronomie, Königstuhl 17, D-69117 France Heidelberg, Germany 50125 Firenze, Italy 12 INAF-Osservatorio Astrofisico di Arcetri, L.go E. Fermi 5, 13 Institute for Astronomy, University of Edinburgh, Blackford Hill, Edinburgh EH9 3HJ, UK 14 LESIA, Observatoire de Paris-Meudon, CNRS, Université Pierre et Marie Curie, Université Paris Diderot, 5 Place Jules Janssen, F-92195 Meudon, France 15 INAF-Osservatorio Astrofisico di Catania, Via S. Sofia 78, I-95123 Catania, Italy 16 INAF-Osservatorio Astronomico di Capodimonte, Via Moiariello,16 I-80131 Napoli, Italy 17 Unidad Mixta Internacional Franco-Chilena de Astronomia, CNRS/INSU UMI 3386 and Departamento de Astronomia„ Universidad de Chile, Casilla 36-D, Santiago, Chile 18 Observatoire de Genéve, University of Geneva, 51 Chemin des Maillettes, 1290, Versoix, Switzerland 19 INAF-Osservatorio Astronomico di Brera, via Emilio Bianchi 46, 23807, Merate (LC), Italy 20 INAF-Istituto di Astrofisica Spaziale e Fisica Cosmica di Mi- lano, Via E. Bassini 15, 20133 Milano, Italy 21 Department of Astronomy, Stockholm University, SE-106 91 Stockholm, Sweden 22 European Southern Observatory, Karl-Schwarzschild-Str. 2, D85748 Garching, Germany 23 Laboratoire Lagrange (UMR 7293), UNSA, CNRS, Obser- vatoire de la Côte d'Azur, Bd. de l'Observatoire, 06304 Nice Cedex 4, France 24 Department of Astronomy, University of Michigan, 311 West Hall, 1085 S. University Avenue, Ann Arbor, MI 48109, USA Appendix A: Background Objects in the IRDIS field of view As described in Sect. 3, seven points sources are detected in the IRDIS FoV. Their astrometry and photometry are listed in Table A.1 and Fig. A.1 unambiguously shows that all of them are background objects. Article number, page 17 of 19 A&A proofs: manuscript no. hd100546Sissa Table A.1. Separation in α and δ of the background objects in the IRDIS FoV of HD 100546 considering data from 4 May 2015, May 29th 2015 and May 31st 2016. Epochs with Italic shape values correspond to those observed points not represented in Figure A.1. date 5/4/15 5/30/15 6/1/16 5/4/15 5/30/15 6/1/16 5/4/15 5/30/15 6/1/16 5/4/15 5/30/15 6/1/16 5/4/15 5/30/15 6/1/16 5/4/15 5/30/15 6/1/16 5/4/15 5/30/15 6/1/16 cc_id ∆RA ∆RAerr [mas] 1 1 1 1 1 1 1 1 1 7 2 0 0 1 4 7 5 0 12 5 0 [mas] -2887 -2887 -2850 -759 -760 -718 2682 2686 2723 -504 -505 -490 3137 3137 3162 -2286 -2292 -2230 4574 4574 4645 0 0 0 1 1 1 2 2 2 3 3 3 4 4 4 5 5 5 6 6 6 ∆δ ∆δerr [mas] 1 1 2 1 1 1 2 1 1 9 2 0 0 2 4 8 5 0 14 5 0 [mas] 4263 4261 4253 -4371 -4383 -4378 -4743 -4756 -4751 -3151 -3167 -3211 5233 5227 5232 4042 4045 4044 2626 2617 2623 Article number, page 18 of 19 E. Sissa et al.: High-Contrast study of the candidate planets and protoplanetary disk around HD 100546 Fig. A.1. Top panel: proper motion in α and δ of the background objects in the IRDIS FoV of HD100546 considering data from 4 May 2015 (filled black), May 29th 2015 (open black), May 31st 2016 (filled red). Dotted points and errorbars represent the expected position for a background object. Middle panel: observed time variation of the separation compared with the expected one for a background object. Bottom panel: observed time variation of the position angle compared with the expected one for a background object. Article number, page 19 of 19 HIP_56379_cc000-2840-2860-2880-2900∆α (mas)42204240426042804300∆δ (mas)-0.20.00.20.40.60.81.0Time (years)51005120514051605180Separation (mas)-0.20.00.20.40.60.81.0Time (years)325.4325.6325.8326.0326.2326.4326.6Position Angle (o)HIP_56379_cc001-700-720-740-760-780∆α (mas)-4420-4400-4380-4360-4340∆δ (mas)-0.20.00.20.40.60.81.0Time (years)442044404460Separation (mas)-0.20.00.20.40.60.81.0Time (years)189.0189.5190.0Position Angle (o)HIP_56379_cc00227402720270026802660∆α (mas)-4780-4760-4740-4720∆δ (mas)-0.20.00.20.40.60.81.0Time (years)54205440546054805500Separation (mas)-0.20.00.20.40.60.81.0Time (years)149.8150.0150.2150.4150.6150.8Position Angle (o)HIP_56379_cc003-440-460-480-500-520∆α (mas)-3200-3180-3160-3140-3120∆δ (mas)-0.20.00.20.40.60.81.0Time (years)314031603180320032203240Separation (mas)-0.20.00.20.40.60.81.0Time (years)187.5188.0188.5189.0189.5190.0Position Angle (o)HIP_56379_cc00432003180316031403120∆α (mas)5200522052405260∆δ (mas)-0.20.00.20.40.60.81.0Time (years)6080610061206140Separation (mas)-0.20.00.20.40.60.81.0Time (years)30.630.831.031.231.431.6Position Angle (o)HIP_56379_cc005-2220-2240-2260-2280-2300∆α (mas)40004020404040604080∆δ (mas)-0.20.00.20.40.60.81.0Time (years)458046004620464046604680Separation (mas)-0.20.00.20.40.60.81.0Time (years)330.0330.5331.0331.5Position Angle (o)HIP_56379_cc00646404620460045804560∆α (mas)25802600262026402660∆δ (mas)-0.20.00.20.40.60.81.0Time (years)5200525053005350Separation (mas)-0.20.00.20.40.60.81.0Time (years)59.560.060.561.0Position Angle (o)
1610.04992
1
1610
2016-10-17T07:37:41
Generation of Highly Inclined Trans-Neptunian Objects by Planet Nine
[ "astro-ph.EP" ]
The trans-Neptunian region of the solar system exhibits an intricate dynamical structure, much of which can be explained by an instability-driven orbital history of the giant planets. However, the origins of a highly inclined, and in certain cases retrograde, population of trans-Neptunian objects remain elusive within the framework of this evolutionary picture. In this work, we show that the existence of a distant, Neptune-like planet that resides on an eccentric and mildly inclined orbit fully accounts for the anomalous component the trans-Neptunian orbital distribution. Adopting the same parameters for Planet Nine as those previously invoked to explain the clustering of distant Kuiper belt orbits in physical space, we carry out a series of numerical experiments which elucidate the physical process though which highly inclined Kuiper belt objects with semi-major axes smaller than 100 AU are generated. The identified dynamical pathway demonstrates that enigmatic members of the Kuiper belt such as Drac and Niku are derived from the extended scattered disk of the solar system.
astro-ph.EP
astro-ph
Draft version October 18, 2016 Preprint typeset using LATEX style emulateapj v. 5/2/11 6 1 0 2 t c O 7 1 . ] P E h p - o r t s a [ 1 v 2 9 9 4 0 . 0 1 6 1 : v i X r a GENERATION OF HIGHLY INCLINED TRANS-NEPTUNIAN OBJECTS BY PLANET NINE Division of Geological and Planetary Sciences, California Institute of Technology, Pasadena, CA 91125 Konstantin Batygin & Michael E. Brown Draft version October 18, 2016 ABSTRACT The trans-Neptunian region of the solar system exhibits an intricate dynamical structure, much of which can be explained by an instability-driven orbital history of the giant planets. However, the origins of a highly inclined, and in certain cases retrograde, population of trans-Neptunian objects remain elusive within the framework of this evolutionary picture. In this work, we show that the existence of a distant, Neptune-like planet that resides on an eccentric and mildly inclined orbit fully accounts for the anomalous component the trans-Neptunian orbital distribution. Adopting the same parameters for Planet Nine as those previously invoked to explain the clustering of distant Kuiper belt orbits in physical space, we carry out a series of numerical experiments which elucidate the physical process though which highly inclined Kuiper belt objects with semi-major axes smaller than a < 100 AU are generated. The identified dynamical pathway demonstrates that enigmatic members of the Kuiper belt such as Drac and Niku are derived from the extended scattered disk of the solar system. 1. INTRODUCTION The detection and observational characterization of the Kuiper belt have caused a qualitative shift in our un- derstanding of the solar system's post-nebular evolution. The gradual unveiling of the trans-Neptunian region's or- bital distribution has led to the replacement of a largely static solar system formation scenario (Cameron 1988; Lissauer 1993) with a dynamic picture, wherein long- range planetary migration is facilitated by the onset of a transient dynamical instability (Tsiganis et al. 2005; Morbidelli et al. 2008). This new class of instability- driven models, collectively known as the Nice model, has been remarkably successful in explaining a number of perplexing features within the solar system. The Nice model's list of accolades begins with the re- production of the solar system's planetary architecture itself (Tsiganis et al. 2005; Batygin & Brown 2010). The Nice model further explains the origins of Jovian and Neptunian co-orbital (Trojan) populations (Morbidelli et al. 2005; Nesvorn´y & Vokrouhlick´y 2009), while si- multaneously providing a natural trigger for the Lunar late heavy bombardment (Gomes et al. 2005; Levison et al. 2011). Finally, the Nice model successfully ac- counts for the dynamically excited orbital distribution of the resonant, hot classical, as well as scattered disk sub- populations of the Kuiper belt1 (Levison et al. 2008). De- spite these successes, however, the Nice model fails to ex- plain a notable, highly inclined subset of trans-Neptunian objects (TNOs), leaving the physical mechanism respon- sible for their nearly orthogonal and retrograde orbits elusive. The origin of this remarkable group of small bodies is the primary focus of this paper. Dynamical emplacement of icy debris into the Kuiper belt during the epoch of Neptune's migration can yield inclinations as large as i ∼ 40 − 60 deg (the exact value [email protected] 1 The so-called cold classical component of the Kuiper belt likely formed in-situ, and the Nice model largely preserves its primordial unperturbed state (Batygin et al. 2011; Nesvorn´y 2015a). depends on the details of Neptune's assumed evolution; Nesvorn´y 2015b). Although substantial, such inclina- tion are dwarfed by the nearly-perpendicular orbits of the TNOs Drac (2008 KV42; a = 41 AU, e = 0.5, i = 103 deg; Gladman et al. 2009) and Niku (2011 KT19; a = 36 AU, e = 0.3, i = 110 deg; Chen et al. 2016). Even more dramatically, the object 2016 NM56 (a = 74 AU, e = 0.9, i = 144 deg) occupies a retrograde orbit that is relatively close to the plane of the solar system. Simply put, there exists no physical mechanism to pro- duce such inclinations within the framework of the Nice model. How then, are such highly inclined orbits gener- ated? In a recent study (Batygin & Brown 2016), we pro- posed that the physical alignment of TNO orbits with semi-major axes greater than a (cid:38) 150 − 250 AU and perihelion distance beyond q (cid:38) 30 AU can be explained by the existence of an additional Neptune-like planet, which resides on a distant, eccentric, and moderately in- clined orbit. In this work, we refer to this object as "Planet Nine." Numerical simulations reported in Baty- gin & Brown (2016) and Brown & Batygin (2016a) sug- gest that this body has a mass of m9 ∼ 10 m⊕, perihe- lion distance of q9 ∼ 250 AU, and a semi-major axis of a9 ∼ 400 − 700 AU. Gravitational torques exerted by Planet Nine onto the small bodies it shepherds, manifest in an exten- sive web of mean-motion resonances, which maintain a rough co-linearity of distant TNO orbits over multi-Gyr timescales. Importantly, however, not all trajectories remain physically confined: Kozai-Lidov type interac- tions (Lidov 1962; Kozai 1962) driven by Planet Nine can dramatically modulate the eccentricities and incli- nations of distant KBOs, thereby reproducing a popula- tion of highly inclined (i > 40 deg), large semi-major axis (a > 100 AU) Centaurs (Gomes et al. 2015). With this notion in mind, here we investigate the possibility that Planet Nine can self-consistently explain the unusual in- clinations of Drac, Niku, 2016 NM56, as well as every other member of the trans-Neptunian population with in- 2 Batygin & Brown Fig. 1.- Orbital distribution of the trans-Neptunian region. The current observational census of all TNOs with a ∈ (30, 100) AU is shown. Regular objects with i < 60 deg are depicted with small black dots, and the anomalous objects with i > 60 deg are emphasized with blue circles. The three retrograde orbits of Drac, Niku, and 2016 NM56 are labeled explicitly. The theoretically computed orbital distribution is shown as a density histogram that underlies the observational data. Gray and green colors denote orbital paths traced by particles with dynamical lifetimes in the range of 3 − 4 Gyr and in exceess of 4 Gyr respectively. Simultaneously, transparency is used as a proxy for the amount of time spent by particles in a given region of orbital element space, with solid color corresponding to a higher probability of dynamical emplacement. Clearly, the entire collection of i > 60 deg objects, including those occupying retrograde orbits, is adequately explained by the existence of Planet Nine. clinations greater than i > 60 deg and semi-major axes below a < 100 AU i.e. outside of the range of possibili- ties of the Nice model and outside Planet Nine's region of direct gravitational influence. The paper is organized as follows. We describe the details of our numerical experiments in section 2, and present the results in section 3. We summarize and dis- cuss the implications of our findings in section 4. 2. NUMERICAL EXPERIMENTS In order to investigate the dynamical behavior of small bodies under the influence of Jupiter, Saturn, Uranus, Neptune, as well as Planet Nine, we have carried out a series of numerical N -body experiments. The per- formed simulations were evolutionary in nature: a syn- thetic solar system, with initial conditions correspond- ing to the final stages of the Nice model (Levison et al. 2008; Batygin et al. 2011; Nesvorn´y 2015b), was evolved forward in time for 4 Gyr. Following the numerical calculations reported in Batygin & Brown (2016) and Brown & Batygin (2016a), the starting configuration of the Kuiper belt was represented by a disk of 3,200 test particles, uniformly distributed across the semi-major axes range a ∈ (150, 550) AU and perihelion distance q ∈ (30, 50) AU. The initial inclinations were drawn from a half-normal distribution with a standard deviation of σi = 15 deg, while the remaining orbital angles (namely, argument of perihelion, longitude of ascending node, and mean anomaly) were assigned random values between 0 and 360 deg. In contrast with our previous models, here we did not mimic the orbit-averaged gravitational field of Jupiter, Saturn, and Uranus with an enhanced quadrupolar com- ponent of the Sun's potential. Instead, we modeled the planets in a direct N−body fashion, fully resolving their Keplerian motion. For our nominal run, the known giant planets were placed on their current orbits, while Planet Nine was chosen to have a mass m9 = 10 m⊕ and initial- ized on an orbit with a9 = 600 AU, e9 = 0.5 , i9 = 30 deg, and ω9 = 150 deg. We note that although this choice of parameters is marginally different from the a9 = 700 AU, e9 = 0.6 AU, i9 = 30 deg Planet Nine considered in Baty- gin & Brown (2010), it generates a synthetic Kuiper belt that provides an optimal match to the existing observa- tions (Brown & Batygin 2016a). A detailed analysis of the synthetic distant Kuiper belt sculpted in this sim- ulation is presented in our companion paper (Brown & Batygin 2016b). The calculations were carried out using the mercury6 gravitational dynamics software package (Chambers 1999). A hybrid Wisdom-Holman/Bulirsch-Stoer algo- rithm (Wisdom & Holman 1992; Press et al. 1992) was employed for all simulations, adopting a time-step of τ = 300 days. The presence of the terrestrial planets was ignored, and any object that attained a radial distance smaller than r < 5 AU or larger than r > 10, 000 AU was removed from the simulation. All orbital elements were measured with respect to Jupiter's orbital plane, which precesses by a few degrees over the lifetime of the solar system due to the gravitational influence of Planet Nine (Bailey et al. 2016; Gomes et al. 2016). In addition to our nominal case, we carried out a series of simulations with lower particle count, sampling the fa- vorable locus of parameter space identified in Brown & Batygin (2016a). To this end, we found that the low- inclination component of the distant Kuiper belt is far more sensitive to the specific orbit of Planet Nine than 305070901020304006018012030901500601801203090150inclination (deg)inclination (deg)semi-major axis (AU)perihelion distance (AU)JDracNikuDracNikuNM56retrograderetrogradeNM56Nice + P9P9SUNa9=600AUe9=0.5i9=30deg Make the Kuiper Belt Great Again! 3 the high-inclination component of the sculpted test par- ticle population (which is the primary focus of this pa- per). As a result, here we restrict ourselves to presenting only the results from our nominal calculation, keeping in mind that they can be deemed representative for any reasonable choice of Planet Nine's parameters. 3. RESULTS The current observational census of multi-opposition in the range a ∈ TNOs with semi-major axes (30, 100) AU is presented in Figure (1). Objects with inclinations greater than i > 60 deg are emphasized, as they represent the anomalous component of the orbital distribution. The left and right panels show inclination as a function of semi-major axes and perihelion distance respectively. Recalling that the entire small body population is ini- tialized with a > 150 AU in the simulations, we have an- alyzed the dynamical evolution of the test particles with an eye towards identifying objects that veer into the re- gion depicted in Figure (1). Owing to chaotic variations of the orbits, instances where objects enter this domain are in fact quite common. However, these excursions are often followed by ejection from the solar system. As a result, it is sensible to focus only on objects whose dy- namical lifetime is comparable to the age of the sun. The simulated orbital distribution of long-term stable bodies is shown as a density histogram that underlies the observational data in Figure (1). Green squares cor- respond to the orbital distribution traced out by objects that are stable over the full 4 Gyr integration period, while the gray squares represent bodies with dynamical lifetimes between 3 and 4 Gyr. Transparency of the color is used as a logarithmic proxy for the amount of time the particles spend in a given box, with solid colors corre- sponding to regions of higher visitation probability. Clearly, the entire observational data set of TNOs with i > 60, including the retrograde orbits of Drac, Niku, and 2016 NM56, is well explained by the simulation results. Simultaneously, it is noteworthy that the theoretical in- clination distribution is not uniform. Instead, it is com- prised of two components: one that extends from i = 0 to i ∼ 110 deg and a second, somewhat less densely pop- ulated component that is centered around i ∼ 150 deg. These two constituents also differ in their characteristic perihelion distances, with the lower inclination part ex- tending from q ∼ 5 AU to q ∼ 35 AU, and the higher inclination part characterized by substantially lower val- ues of q ∼ 10 AU. What is the physical mechanism through which the simulated particles acquire these unusual orbits? Our calculations indicate that a sequential combination of Kozai-Lidov cycles driven by Planet Nine and close en- counters with Neptune plays a dominant role. First, Kozai-Lidov interactions induce large-scale oscillations in the inclinations and eccentricities of particles with ini- tially large semi-major axes. Thus, the perihelion dis- tance of a typical low-inclination scattered disk object can grow to larger values, only to recede back down to hug the orbit of Neptune at a much higher inclination. Then, close encounters that inevitably ensue, facilitate a stochastic evolution of the semi-major axis, occasionally reducing it below a < 100 AU. To illustrate this process, Figure (2) shows the time- Fig. 2.- Time-series of a subset of simulated particles that at- tain highly inclined orbits with a < 100 AU within the span of the integration. The colors hold no physical meaning and sim- ply label different bodies. The orbital parameters of the three retrograde members of the observational census are shown with horizontal lines on each panel, and points in time when simulated objects attain orbits that are close to their observed counterparts are emphasized with black squares. The two objects with longest dynamical lifetimes are particularly notable, as they reproduce the orbits of Drac, Niku, and 2016 NM56 almost exactly. Additionally, the object depicted in blue exemplifies a reversal of the dynamical pathway for generation of Drac-type orbits from the distant Kuiper belt, as it reacquires a typical scattered disk orbit towards the end of the simulation. series of a subset of simulated objects that come to re- semble Drac and Niku, or 2016 NM56 at one point in time. The top, middle, and bottom panels depict semi- major axes, eccentricities, and inclinations as functions of time respectively, while the boxes indicate the phases of dynamical evolution where simulated particles attain orbits that are close to those of the observed retrograde objects. As can be deduced by examining the individual paths of the particles in detail, the aforementioned qual- itative picture holds, although the orbits generally ex- hibit chaotic motion. Among the exemplified evolutions 012341801206000.20.40.8301003001000time (Gyr)inclination (deg)eccentricitysemi-major axis (AU)retrogradeprogradeNice model max inclinationNeptune's orbitDracNikuNikuDracNikuDracNM56NM56NM56 4 Batygin & Brown (which represent only a small subset of the entire simu- lation suite), the two objects with the longest dynamical lifetimes (shown in black and red) are particularly no- table. Starting out on low-inclination, high-eccenticity orbits with semi-major axes beyond a > 300 AU, these bodies evolve to attain orbital states that are almost ex- act replicas of Drac, Niku, and 2016 NM56. Conventional N−body simulations of TNOs in this class (e.g. Gladman et al. 2009; Chen et al. 2016; see also Dones et al. 1996; Di Sisto & Brunini 2007) yield strictly unstable orbits with dynamical lifetimes of order ∼ 10 Myr−1 Gyr. Although future ejection of these bod- ies is indeed a distinct possiblity, our simulations sug- gest that the existence of Planet Nine can potentially prolong their lifetimes, through a reversal of their de- livery process. An example of such behavior is demon- strated by the orbit depicted in blue in Figure (2) - after reaching a semi-major axis of a ∼ 35 AU at t ∼ 2.2 Gyr, the object's perihelion distance and semi-major axes in- crease to values comparable to their starting conditions. The concurrent decrease in the inclination means that by the time the object escapes from the solar system at t ∼ 3.5 Gyr, it bears semblance to a regular member of the distant Kuiper belt. Therefore, our simulations not only reveal that Drac, Niku and other TNOs occupy- ing highly inclined orbits are sourced from the extended scattered disk, they also point to an evolutionary future where some of these bodies will once again return to more conventional, low-inclination orbits. 4. DISCUSSION In this work, we have carried out a sequence of nu- merical experiments, with an eye towards exploring the dynamical effects of Planet Nine onto the a < 100 AU portion of the Kuiper belt. Adopting the same orbital parameters for Planet Nine as those required to generate physical confinement among distant (a > 150 AU) TNO orbits (Batygin & Brown 2016; Brown & Batygin 2016a), we have shown that Planet Nine is capable of explaining the full range of inclinations observed in the Kuiper belt. The origin of these unusual objects had remained elusive until now (Gladman et al. 2009; Nesvorn´y 2015b), and the existence of Planet Nine provides a resolution for this puzzle. While the orbital domain currently occupied by the highly inclined sub-population of the Kuiper belt lies outside of Planet Nine's direct gravitational reach, our calculations suggest that these bodies had substantially larger semi-major axes in the past. Specifically, the sim- ulations reveal a dynamical pathway wherein long pe- riod TNOs undergo Kozai-Lidov oscillations facilitated by Planet Nine, and subsequently scatter inwards due to close encounters with Neptune (as well as other known gi- ant planets). This dynamical pathway is time-reversible, and the numerical experiments reveal examples of trajec- tories that originate within the extended scattered disk, proceed to become nearly orthogonal members of the classical Kuiper belt, and subsequently reacquire long orbital periods and low inclinations. This finding places large semi-major axis Centaurs (Gomes et al. 2015) and retrograde Kuiper belt objects such as Drac and Niku into the same evolutionary context. Although our model explains extant data adequately, it can be tested with further observations. To this end, we note that the simulations predict a rather specific orbital distribution in a−q−i space. The explicit non-uniformity of this theoretical expectation renders our model readily falsifiable: detection of bodies within the empty region can constitute significant evidence against the dynam- ical mechanism described herein (although we caution that the empty region diminishes somewhat for simu- lations with smaller values of a9, e.g. 500 AU). We fur- ther note that there exist high inclination Centraurs with a < 30 AU in the observational data set. Here we have chosen to ignore this component of the data, because our model does not have sufficient resolution in terms of particle count to adequately model the inter-planetary region. However, we speculate that the evolutionary his- tories of these objects may be connected to the high- inclination bodies with a > 30 AU, and modeling their generation presents an interesting avenue for future re- search. In concluding remarks, we wish to draw attention to the recent proposition of Chen et al. (2016), who noted that all Centaurs and TNOs with a < 100 AU, q > 10 AU, and i > 60 deg appear to occupy a common plane. Our simulations do not show the existence of such a plane, and predict that future observations will reveal objects that do not correspond to the pattern pointed out by Chen et al. (2016). Importantly, the latest addition to the observational census of trans-Neptunian objects, 2016 NM56, conforms to the aforementioned selection criteria, but does not lie in the same plane as the other objects in the group. This supports the notion that the apparent clustering of ascending node is not sta- tistically significant. Future observations will continue to test the theoretical expectation outlined by our model. Acknowledgments We are thankful to Ira Flatow and Chris Spalding for inspirational conversations. REFERENCES Bailey, E., Batygin, K., & Brown, M. E. 2016, arXiv:1607.03963 Batygin, K., & Brown, M. E. 2010, ApJ, 716, 1323 Batygin, K., Brown, M. E., & Fraser, W. C. 2011, ApJ, 738, 13 Batygin, K., & Brown, M. E. 2016, AJ, 151, 22 Brown, M. E., & Batygin, K. 2016a, ApJ, 824, L23 Brown, M. E., & Batygin, K. 2016b, in prep. Cameron, A. G. W. 1988, ARA&A, 26, 441 Chambers, J. E. 1999, MNRAS, 304, 793 Chen, Y.-T., Lin, H. W., Holman, M. J., et al. 2016, ApJ, 827, L24 Dones, L., Levison, H. F., & Duncan, M. 1996, Completing the Inventory of the Solar System, 107, 233 Gladman, B., Kavelaars, J., Petit, J.-M., et al. 2009, ApJ, 697, L91 Gomes, R., Levison, H. F., Tsiganis, K., & Morbidelli, A. 2005, Nature, 435, 466 Gomes, R. S., Soares, J. S., & Brasser, R. 2015, Icarus, 258, 37 Gomes, R., Deienno, R., & Morbidelli, A. 2016, arXiv:1607.05111 Kozai, Y. 1962, AJ, 67, 591 Levison, H. F., Morbidelli, A., Van Laerhoven, C., Gomes, R., & Di Sisto, R. P., & Brunini, A. 2007, Icarus, 190, 224 Tsiganis, K. 2008, Icarus, 196, 258 Make the Kuiper Belt Great Again! 5 Levison, H. F., Morbidelli, A., Tsiganis, K., Nesvorn´y, D., & Gomes, R. 2011, AJ, 142, 152 Lidov, M. L. 1962, Planet. Space Sci., 9, 719 Lissauer, J. J. 1993, ARA&A, 31, 129 Morbidelli, A., Levison, H. F., Tsiganis, K., & Gomes, R. 2005, Nature, 435, 462 Nesvorn´y, D. 2015a, AJ, 150, 68 Nesvorn´y, D. 2015b, AJ, 150, 73 Press, W. H., Teukolsky, S. A., Vetterling, W. T., & Flannery, B. P. 1992, Numerical recipes in FORTRAN. The art of scientific computing, Cambridge: University Press, 2nd ed. Tsiganis, K., Gomes, R., Morbidelli, A., & Levison, H. F. 2005, Morbidelli, A., Levison, H. F., & Gomes, R. 2008, The Solar Nature, 435, 459 System Beyond Neptune, 275 Nesvorn´y, D., & Vokrouhlick´y, D. 2009, AJ, 137, 5003 Wisdom, J., & Holman, M. 1992, AJ, 104, 2022
1811.06550
2
1811
2019-01-29T21:10:14
Predicted Yield of Transits of Known Radial Velocity Exoplanets from the TESS Primary and Extended Missions
[ "astro-ph.EP" ]
Radial velocity (RV) surveys have detected hundreds of exoplanets through their gravitational interactions with their host stars. Some will be transiting, but most lack sufficient follow-up observations to confidently detect (or rule out) transits. We use published stellar, orbital, and planetary parameters to estimate the transit probabilities for nearly all exoplanets that have been discovered via the RV method. From these probabilities, we predict that $25.5^{+0.7}_{-0.7}$ of the known RV exoplanets should transit their host stars. This prediction is more than double the amount of RV exoplanets that are currently known to transit. The Transiting Exoplanet Survey Satellite (TESS) presents a valuable opportunity to explore the transiting nature of many of the known RV exoplanet systems. Based on the anticipated pointing of TESS during its two-year primary mission, we identify the known RV exoplanets that it will observe and predict that $11.7^{+0.3}_{-0.3}$ of them will have transits detected by TESS. However, we only expect the discovery of transits for $\sim$3 of these exoplanets to be novel (i.e., not previously known). We predict that the TESS photometry will yield dispositive null results for the transits of $\sim$125 RV exoplanets. This will represent a substantial increase in the effort to refine ephemerides of known RV exoplanets. We demonstrate that these results are robust to changes in the ecliptic longitudes of future TESS observing sectors. Finally, we consider how several potential TESS extended mission scenarios affect the number of transiting RV exoplanets we expect TESS to observe.
astro-ph.EP
astro-ph
Draft version January 31, 2019 Typeset using LATEX preprint style in AASTeX62 Predicted Yield of Transits of Known Radial Velocity Exoplanets from the TESS Primary and Extended Missions Paul A. Dalba,1 Stephen R. Kane,1 Thomas Barclay,2, 3 Jacob L. Bean,4 Tiago L. Campante,5, 6 Joshua Pepper,7 Darin Ragozzine,8 and Margaret C. Turnbull9 1Department of Earth Sciences, University of California Riverside, 900 University Avenue, Riverside CA 92521, USA 2NASA Goddard Space Flight Center, 8800 Greenbelt Road, Greenbelt, MD 20771, USA 3University of Maryland, Baltimore County, 1000 Hilltop Circle, Baltimore, MD 21250, USA 4Department of Astronomy & Astrophysics, University of Chicago, 5640 S. Ellis Avenue, Chicago, IL 60637, USA 5Instituto de Astrof´ısica e Ciencias do Espa¸co, Universidade do Porto, Rua das Estrelas, PT4150-762 Porto, Portugal 6Departamento de F´ısica e Astronomia, Faculdade de Ciencias da Universidade do Porto, Rua do Campo Alegre, s/n, PT4169-007 Porto, Portugal 7Department of Physics, Lehigh University, 16 Memorial Drive East, Bethlehem, PA 18015, USA 8Brigham Young University, Department of Physics and Astronomy, N283 ESC, Provo, UT 84602, USA 9SETI Institute, Carl Sagan Center for the Study of Life in the Universe, Off-Site: 2801 Shefford Drive, Madison, WI 53719, USA (Received October 22, 2018; Revised November 13, 2018; Accepted November 14, 2018) Submitted to Publications of the Astronomical Society of the Pacific (PASP) ABSTRACT Radial velocity (RV) surveys have detected hundreds of exoplanets through their gravitational interactions with their host stars. Some will be transiting, but most lack sufficient follow-up observations to confidently detect (or rule out) transits. We use published stellar, orbital, and planetary parameters to estimate the transit probabilities for nearly all exoplanets that have been discovered via the RV method. From these probabilities, we predict that 25.5+0.7 −0.7 of the known RV exoplanets should transit their host stars. This prediction is more than double the amount of RV exoplanets that are currently known to transit. The Transiting Exoplanet Survey Satellite (TESS) presents a valuable opportunity to explore the transiting nature of many of the known RV exoplanet systems. Based on the anticipated pointing of TESS during its two-year primary mission, we identify the known RV exoplanets that it will observe and predict that 11.7+0.3 −0.3 of them will have transits detected by TESS. However, we only expect the discovery of transits for ∼3 of these exoplanets to be novel (i.e., not previously known). We predict that the TESS photometry will yield dispositive null results for the transits of ∼125 RV exoplanets. This will represent a substantial increase in the effort to refine ephemerides of known RV exoplanets. We demonstrate that these results are robust to changes in the ecliptic longitudes of future TESS observing sectors. Finally, we consider Corresponding author: P. A. Dalba [email protected] 2 Dalba et al. how several potential TESS extended mission scenarios affect the number of transiting RV exoplanets we expect TESS to observe. Keywords: planets and satellites: detection -- surveys -- methods: statistical 1. INTRODUCTION The vast majority of presently known exoplanets were discovered by dedicated transit surveys using space-based (e.g., Kepler, Borucki et al. 2010; Twicken et al. 2016) and ground-based observatories (e.g., KELT, HATNet, and WASP, Pepper et al. 2003; Bakos et al. 2004; Pollacco et al. 2006). The sheer number of stars that can be photometrically monitored in such surveys clearly overwhelms the probability of having the proper viewing geometry to witness a transit. However, the inverse proportionality between transit probability and orbital semi-major axis (e.g., Beatty & Gaudi 2008) has largely limited the exoplanet characteristics inferred from transit surveys to the inner regions of planetary systems, within a few tenths of an AU from the host star (e.g., Howard et al. 2012). The detection of exoplanets in radial velocity (RV) measurements of their host stars (e.g., Pepe et al. 2004; Howard et al. 2010) is less biased toward short-period exoplanets than the transit method. As a result, RV surveys have revealed exoplanet characteristics at moderate orbital distances (tenths of an AU to several AU from the host star), which has greatly benefited theoretical investigations of planet formation (e.g., Dawson & Murray-Clay 2013; Santerne et al. 2016). At the intersection of the sensitivities of the transit and RV methods exists an opportunity to im- prove the accuracy of exoplanet demographics and properties through the validation and synthesis of independent observations. The combination of transit and RV occurrence rates better constrains the true underlying exoplanet population and identifies the biases of each method (e.g., Howard et al. 2012; Wright et al. 2012; Dawson & Murray-Clay 2013; Guo et al. 2017). Furthermore, the combi- nation of planet mass and radius -- the basic properties inferred from RV and transit observations, respectively -- advances the status of an exoplanet from merely a detected world, to one that can be characterized in depth. An exoplanet's bulk density (derived from planet mass and radius) provides a window to its internal composition (e.g., Weiss & Marcy 2014). Furthermore, mass and radius measurements are imperative for atmospheric characterization, which may place even more precise constraints on planetary interior (e.g., Miller & Fortney 2011; Thorngren et al. 2016) and formation processes (e.g., Garaud & Lin 2007; Oberg et al. 2011). One of the following two approaches is typically employed to obtain measurements of an exoplanet's mass and radius. Either the exoplanet is detected in a transit survey and follow-up RV observations covering the orbital phase of the planet are acquired, or the exoplanet is detected in an RV survey and follow-up photometry within the predicted transit window is acquired. Here, we consider the latter, which is potentially the riskier of the two approaches. RV exoplanet detections are biased toward edge-on orbital inclinations, but substantial photometric follow-up campaigns may nonetheless fail to measure the planet radius in the event that the exoplanet is non-transiting. In many cases, this risk factor has precluded sufficient photometric monitoring in search of transits for known RV exoplanets. This is especially true for exoplanets with relatively long orbital periods. Substantial efforts have previously been made to improve orbital ephemerides known exoplanets. For exoplanets discovered in transit surveys, observations of subsequent transits are typically needed to constrain transit timing variations (e.g., Dalba & Muirhead 2016; Wang et al. 2018). For exoplan- TESS Yield of Transits from Known Radial Velocity Exoplanets 3 ets discovered in RV surveys, multiple types of follow-up observations are usually required. Efforts to follow-up known RV exoplanets such as the Transit Ephemeris Refinement and Monitoring Survey (TERMS; Kane et al. 2009) acquire spectroscopic and photometric observations to search for tran- sits and characterize host stars. TERMS and other similar efforts have successfully ruled out (e.g., Kane et al. 2011a) and discovered (e.g., Moutou et al. 2009) transits of a handful of RV exoplanets. However, there are simply not enough telescope resources to target all known RV exoplanets indi- vidually. Instead, this luxury can potentially be afforded by dedicated large-scale transit-hunting missions. With the launch of the Transiting Exoplanet Survey Satellite (TESS, Ricker et al. 2015) came the new opportunity to acquire light curves of known hosts of RV exoplanets. The potential yield of exoplanet discoveries from TESS has been thoroughly explored (Sullivan et al. 2015; Barclay et al. 2018; Ballard 2018; Huang et al. 2018), yet little effort has been focused on the yield of transits from known RV exoplanets. Yi et al. (2018) evaluated the detectability of RV exoplanets from the anticipated Characterizing Exoplanets Satellite (CHEOPS ; Broeg, C. et al. 2013). However, their investigation focused on making accurate predictions of RV exoplanet radii and subsequently transit signal-to-noise ratio (SNR) to maximize the efficiency of the CHEOPS observing strategy. Yi et al. (2018) calculated single point-estimates of transit probability and provided a qualitative interpreta- tion, but their predicted yield of CHEOPS transits was set by SNR arguments (Yi et al. 2018). An extension of the transit probability calculation to the expected yield of transits for all known RV exoplanets, including those to be observed by TESS, remains to be done. In this paper, we take a probabilistic approach to predicting the number of RV exoplanets that transit their host stars and the fraction of those that TESS may detect. In Section 2, we explain the selection criteria used to generate our sample of known RV exoplanets. In Section 3, we describe the determination of all necessary exoplanet parameters and the Monte Carlo simulation of each exoplanet's transit probability1. For each exoplanet in our sample, we report parameters describing the probability density function for its transit probability. In Section 4, we predict the total number of RV exoplanets that transit their host stars and offer comparison to the currently known sample of transiting RV exoplanets. In Section 5, we refine the predicted number of transiting RV exoplanets to only those that may be observed by TESS during its primary mission. We then extend this consideration of TESS to several potential extended mission scenarios. In Section 6, we discuss the influence of several simplifying assumptions that we employ in our analysis on our results. Finally, in Section 7, we summarize our findings. 2. SELECTION OF RV EXOPLANETS All exoplanet data used here were acquired from the NASA Exoplanet Archive2 (NEA). Of the total 3778 confirmed exoplanets, we select only the 677 exoplanets that had "Radial Velocity" as the discovery method. A small fraction of this subset of exoplanets has since been found to transit their host stars. We do not exclude these exoplanets from our analysis so that we may compare the number of RV exoplanets that are currently known to transit with the number that we predict to transit based on our statistical arguments. 1 The code developed here to predict the transit probabilities of RV exoplanets will be made publicly available at https://github.com/pdalba/transit prob. 2 Accessed 2018 August 28. 4 Dalba et al. 3. TRANSIT PROBABILITY For an exoplanet on an eccentric orbit about its host star, the probability that a distant observer witnesses a transit (ptransit) can be expressed as ptransit = (Rp + R⋆)[1 + e cos (π/2 − ω)] a(1 − e2) (1) where Rp is the planetary radius, R⋆ is the stellar radius, a is the orbital semi-major axis, e is the orbital eccentricity, and ω is the longitude of periastron (e.g., Kane 2007; Winn 2010). This relation assumes a uniform distribution for the cosine of the orbital inclination and ignores the prior probability distribution for the exoplanet mass. Therefore, when applied to RV exoplanets, Equation (1) is the prior transit probability, not the posterior transit probability (Ho & Turner 2011; Stevens & Gaudi 2013). For Jupiter-like exoplanets, the prior and posterior transit probabilities are nearly equal (Stevens & Gaudi 2013). Given that the sample of known RV exoplanets is comprised of mostly Jupiter-like exoplanets, we continue our analysis using the relation for prior transit probability (Equation 1). The implications of this choice are considered in Section 6. The NEA data have been federated from multiple sources, so some exoplanets lack one or more of the parameters needed to evaluate Equation (1). Furthermore, the vast majority of exoplanets in our sample do not have reported Rp values, as they either do not transit or have not been found to transit their host stars. Before evaluating transit probabilities, we give special consideration to missing parameter values and the estimation of planet radii. 3.1. Missing Parameter Values In some cases, the NEA lacks values of the physical parameters necessary to evaluate Equation (1). Here, we explain our decision process regarding missing data. If the NEA does not contain a value of the longitude of periastron, we assume ω = π/2. If the NEA does not contain a value of orbital eccentricity, we assume e = 0 (i.e., a circular orbit). The assumption of e = 0 is somewhat of a simplification. The long-period nature of many of the known RV exoplanets suggests that they likely have nonzero orbital eccentricities (Kipping 2013). The assumption of e = 0 therefore yields a lower limit for the transit probability. The resulting effect on the full set of transit probability calculations is likely minimal, but we encourage any future extensions of this work to consider physically-motivated eccentricities, such as Beta (Kipping 2013) or Rayleigh (e.g., Fabrycky et al. 2014) distributions. If the NEA does not contain a value of stellar radius, we estimate its value through the following methods in the following order. First, if values of stellar surface gravity (log g) and stellar mass (M⋆) are available, we estimate R⋆ from the relation (Smalley 2005) log g − log g⊙ = log(cid:18) M⋆ M⊙(cid:19) − 2 log(cid:18) R⋆ R⊙(cid:19) (2) where the solar surface gravity is log g⊙ = 4.4374, the solar mass is M⊙ = 1.989× 1030 kg, and the solar radius is R⊙ = 6.96× 108 m. Second, if values of stellar B-band and V -band magnitudes are available, we estimate R⋆ with a B − V color look-up table under the assumption that the exoplanet host is a dwarf star (Cox 2000, page 388). Third, if B- or V -band magnitude is unavailable or if the B − V color is outside of the range of the look-up table, we exclude this source from further analysis (four exclusions in total: HD 37124 b, HD 37124 c, HD 37124 d, and HD 41004 B b). TESS Yield of Transits from Known Radial Velocity Exoplanets 5 If the NEA does not contain a value of orbital semi-major axis, we use the stellar mass, orbital period (P ), and planet mass (Mp) to estimate a from Kepler's Third Law under the assumption that the orbit has an edge-on inclination. This assumption slightly biases the value of Mp used in Kepler's Third Law. However, since Mp ≪ M⋆ in all cases, the effect of this bias on the sum of stellar and planetary mass is negligible. Only 27 exoplanets in our sample lack an a value in the NEA, and each of those contain values for P , M⋆, and Mp. Hence, none are excluded from further analysis at this step. After all exclusions, our final sample contains 673 exoplanets (of a total 677 exoplanets). 3.2. Estimation of Exoplanet Radii Although a few of the exoplanets in our sample have measured radii, we ignore this information to maintain consistency with the vast majority of RV exoplanets. We estimate Rp for all of the exoplanets in our sample using the forecaster tool (Chen & Kipping 2017). forecaster applies probabilistic mass-radius relations to an input exoplanet mass and returns an exoplanet radius along with lower and upper uncertainties based on a posterior probability distribution (Chen & Kipping 2017). 3.3. Monte Carlo Simulations of Transit Probability Using measurements or estimates of each of the necessary stellar, planetary, and orbital parameters, we conduct a Monte Carlo simulation of the transit probabilities from Equation (1). We begin by approximating a probability density function (PDF) for each physical parameter from which parameter values can be drawn. For parameters with symmetric errors (i.e., equal upper and lower errors), we use a normal PDF peaked at the parameter value and having a standard deviation equal to the parameter uncertainty. For parameters with asymmetric errors, we employ a skew-normal PDF (Azzalini 1985). A skew- normal PDF is defined by three functional parameters, which control location, shape, and skewness. We obtain these functional parameters through the non-linear least-squares regression technique de- scribed by Espinoza & Jord´an (2015, Appendix C). Briefly, the physical parameter value and its upper and lower uncertainties are assumed to have been derived from the 16th, 50th, and 84th per- centiles of their PDF. The residuals between these percentiles and those of a skew-normal distribution defined by the location, shape, and skewness parameters are then minimized. For parameters with no uncertainties provided by the NEA, we assume symmetric errors of 10% the parameter value. This decision primarily influences stellar radii determined through Equation (2). For our sample of exoplanets, many of those lacking values of R⋆ also lack uncertainties on M⋆, despite having a listed value for M⋆. Propagating arbitrarily large errors on M⋆ through Equation (2) leads to even larger errors on R⋆, which artificially inflate the transit probability. We find that a value of 10% error does not inflate ptransit. We evaluate the transit probability (Equation 1) 1× 106 times, by drawing parameter values from the aforementioned PDFs. Only probabilities in the range 0 ≤ ptransit ≤ 1 are allowed. The result is a PDF for ptransit, an example of which is shown in Figure 1. The PDFs for ptransit are not necessarily normal, so we describe them using percentiles instead of the mean and standard deviation. In Table 1, we describe the location and scale of the ptransit PDFs for all exoplanets in our sample using the 16th, 50th, and 84th percentiles of the distributions. 4. PREDICTED NUMBER OF RV EXOPLANETS THAT TRANSIT THEIR HOST STARS 6 Dalba et al. 1600 1400 1200 1000 800 600 400 200 y t i s n e D y t i l i b a b o r P 0 0.0010 0.0015 0.0020 0.0025 0.0030 0.0035 0.0040 Transit Probability Figure 1. Normalized histogram of the transit probability (ptransit), for the RV exoplanet HD 29012 b as an example, found by a Monte Carlo simulation of Equation (1). The 50th percentile is marked by a dashed black line, and the 16th and 84th percentiles are marked by solid black lines. The distribution of ptransit values is slightly skewed toward higher values of ptransit. We use the transit probability PDFs for each exoplanet in our sample to predict the total number (N) of RV exoplanets that have the proper orientation to transit their host stars. First, we let ni be a random variable representing the number of transits the ith exoplanet contributes to the total. The value of ni can either be zero or unity. The expectation of ni is simply E[ni] = ptransit,i ni = ptransit,i. It follows that the total number of transiting planets in the entire sample of 673 RV exoplanets is described by N = 673 Xi=1 ptransit,i (3) For each of the 1× 106 evaluations of the transit probability, Equation (3) is also evaluated, resulting in a PDF for N (Figure 2). The distribution of N has a mean of 25.5 exoplanets and a standard deviation of 0.66 exoplanet. The 16th, 50th and 84th percentiles of the distribution of N values are 24.8, 25.5, and 26.2 exoplanets, respectively, which (to significant figures) yields a final estimate of N = 25.5+0.7 −0.7 exoplanets. How does the predicted number of RV exoplanets that transit compare to the number of RV exoplanets that have actually been found to do so? Of our sample of 673 RV exoplanets, only 12 have been found to transit their host stars3. Hence, the current number of transiting exoplanets from our sample is more than 19 standard deviations below the expectation. This suggests that additional photometric monitoring of RV exoplanets -- and especially those with high transit probabilities (see Table 1) -- is likely to reveal transits. 3 The RV exoplanets already known to transit are those in the NEA with values of "1" in the "pl tranflag" column including HD 189733 b, HD 209458 b, HD 17156 b, 55 Cnc e, GJ 436 b, HD 97658 b, HD 149026 b, HD 219134 b, HD 219134 c, 30 Ari B b, HD 80606 b, and GJ 3470 b. TESS Yield of Transits from Known Radial Velocity Exoplanets 7 Simulation Normal 0.6 0.5 0.4 0.3 0.2 0.1 y t i s n e D y t i l i b a b o r P 0.0 22 23 24 25 26 27 28 29 30 Predicted Number of RV Exoplanets that Transit Figure 2. Normalized histogram of N , the predicted number of RV exoplanets in our sample that transit their host stars. The black curve is a normal distribution with the same mean and standard deviation of the distribution of N values. Based on our Monte Carlo simulation of Equation (1), we predict that 25.5+0.7 −0.7 RV exoplanets have the proper orientation to transit their host stars. 12 exoplanets within our sample are already known to transit, suggesting that there ∼13 more transiting RV exoplanets that have yet to be identified. We consider whether the 12 known transiting RV exoplanets have relatively high transit probabil- ities compared to the rest of the sample. The 12 RV exoplanets currently known to transit have a mean transit probability of 11.6%. To compare this value to the rest of the sample, we randomly draw 1× 104 sets of 12 RV exoplanets (not including those known to transit) and record their mean transit probability. The final distribution of mean transit probabilities has a mean of 3.6% and a standard deviation (σ) of 1.5%. The 11.6% mean transit probability of the known transiting RV exoplanets is more than 5σ higher than the mean of the distribution. This suggests that the sample of RV exoplanets currently known to transit has significantly higher transit probabilities than the rest of the RV exoplanet sample. This result is not surprising as telescope resources that provide follow-up for RV exoplanets are limited and the exoplanets that are most likely to transit have been prioritized. Efforts to detect or rule out transits of RV exoplanets include the Transit Ephemeris Refinement and Monitoring Survey (TERMS; Kane et al. 2009), which has been refining transit ephemerides and conducting photometric transit searches for RV exoplanets for nearly a decade. TERMS, and other similar efforts, have thoroughly ruled out transits for the exoplanet hosts GJ 581 (e.g., L´opez-Morales et al. 2006; Dragomir et al. 2012a), GJ 876 (e.g., Shankland et al. 2006), HD 114762 (Kane et al. 2011a), HD 63454 (Kane et al. 2011b), HD 192263 (Dragomir et al. 2012b), HD 38529 (b planet, Henry et al. 2013), HD 130322 (b planet, Hinkel et al. 2015a), 70 Vir (Kane et al. 2015), HD 6434 (Hinkel et al. 2015b), and HD 20782 (Kane et al. 2016). Less confident null detections of transits have also been reported for HD 168443 b (Pilyavsky et al. 2011), HD 37605 b (Wang et al. 2012), HD 156846 b (Kane et al. 2011c), and Proxima Cen b (e.g., Kipping et al. 2017; Blank et al. 8 Dalba et al. 2018). However, resources for individual efforts such as those listed here are limited. Depending on observing strategy, large-scale photometric monitoring programs may be especially helpful for detect- ing or ruling out transits from known RV exoplanets. The ground-based transit surveys listed above (KELT, SuperWASP, and HATNet/HATSouth) have years of photometric observations of much of the sky. However, the typical magnitudes of known RV planet hosts tend to be brighter than even the bright limit of these surveys. Space-based transit surveys, as discussed in the following section, may then provide the best opportunity to follow-up on known RV exoplanets. 5. PREDICTED NUMBER OF TRANSITING RV EXOPLANETS TO BE OBSERVED BY TESS TESS was launched in April 2018, and is actively searching bright stars for transiting exoplanets. Over the course of its two-year primary mission, TESS will search for transits across most of the sky, including in many of the currently known RV exoplanet systems. TESS has four cameras positioned in a 1x4 array that provide a combined ∼2300 deg2 field of view. TESS scans approximately half of the sky in 13 sectors, each of which receives ∼27 days of continuous observation (at either 2- or 30-minute cadence). Depending on the ecliptic longitude and latitude of a known RV exoplanet host, it may be observed in as many as 13 consecutive sectors. We determine the sectors and cameras in which all 673 exoplanets in our sample will be (or were) observed. We uploaded the NEA right ascension and declination for each exoplanet in our sample to the Web TESS Viewing Tool4 (WTV; Mukai & Barclay 2017), which estimates the pointing of TESS's cameras using predicted spacecraft ephemerides. At the time of writing, the WTV tool was only implemented for Cycle 1 of the TESS primary mission. We estimate the TESS pointings for Cycle 2 by assuming that the ecliptic longitudes of the sector boundaries are the same as Cycle 1, and mirroring the ecliptic latitudes from the southern to the northern ecliptic hemispheres. The output sector(s) and camera(s) (if any) for each RV exoplanet in our sample are provided in Table 1. We explore how this assumption -- and the potential alteration of TESS's predicted pointing -- affects our results by shifting the ecliptic longitudes of the Cycle 2 sector boundaries in increments of +5◦, up to a maximum of +30◦, and repeating our analysis. This amount of shifting is sufficient to give TESS its original total field of view. We find that specific longitude shifts cause specific stars to receive different amounts of observation or to fall outside of TESS's field of view (e.g., in between adjacent sectors at low ecliptic latitudes). However, variation of the ecliptic longitudes of the Cycle 2 sector boundaries do not alter the results for all RV exoplanets discussed below in a statistically significant manner. This outcome is expected because the distribution of RV exoplanets and their transit probabilities on the sky is essentially random (Figure 3). Nonetheless, this outcome is significant because it suggests that the search for transits of known RV exoplanets is resilient to changes in the ecliptic longitudes of TESS's sector boundaries. Stars at different ecliptic latitudes receive different amounts of TESS observation. If the ith exo- planet host is observed in a total of s TESS sectors, then the total duration (τi) of TESS observation for that exoplanet host can be approximated by τi = 27s days. If this exoplanet's orbital period (Pi) is greater than τi, then its transit probability as observed by TESS is reduced by a factor of (τi/Pi). If instead Pi < τi, then the limited TESS baseline does not affect the probability of detecting a transit. This method of estimating transiting probabilities neglects the orbital ephemerides -- except 4 https://heasarc.gsfc.nasa.gov/cgi-bin/tess/webtess/wtv.py TESS Yield of Transits from Known Radial Velocity Exoplanets 9 75◦ 60◦ 45◦ 30◦ 30◦ 80◦ 130◦ 180◦ 230◦ 280◦ 330◦ 15◦ 0◦ -15◦ -30◦ -45◦ -60◦ -75◦ 1.0 0.1 0.01 0.001 y t i l i b a b o r P t i s n a r T S S E T Figure 3. All known RV exoplanets in our sample displayed in ecliptic coordinates and colored by the probability that TESS observes a transit (ptransit,TESS). Exoplanets symbolized by black "x" markers are not expected to be observed by TESS. For stars hosting multiple exoplanets, the color represents the highest ptransit,TESS value. for orbital period -- of our known RV exoplanets. A justification and discussion of this choice is given in Section 6. We repeat the Monte Carlo simulation of Section 3.3, but we multiply each value of ptransit,i by (τi/Pi) to account for TESS's limited observational baseline. The probability (ptransit,TESS,i) that TESS observes a transit of the ith exoplanet in our sample becomes Pi(cid:19) ptransit,TESS,i = ptransit,i(cid:18) τi (4) where (τi/Pi) is forced to unity for any exoplanet with an orbital period shorter than the duration of TESS observation. Values of ptransit,TESS for each exoplanet in our sample are provided in Table 1 and are displayed in Figure 3. Following Section 3.3, we combined the individual ptransit,TESS distributions to predict the total number (NTESS) of known RV exoplanets that will be observed to transit their host stars by TESS (Figure 4). We find that NTESS = 11.7+0.3 −0.3 exoplanets, meaning that we predict that TESS will observe transits of 11 or 12 known RV exoplanets in our sample during its primary mission. We note that our full sample of 673 RV exoplanets includes 12 that have previously been found to transit their host stars. Of these 12 exoplanets, 8 (HD 189733 b, HD 209458 b, HD 17156 b, 55 Cnc e, GJ 436 b, HD 149026 b, HD 219134 b, and HD 219134 c) are expected to have TESS observational baselines that are equal to or longer than their orbital periods. Hence, these 8 exoplanets are essentially guaranteed to be observed in transit by TESS. Two of the remaining four RV exoplanets already known to transit (GJ 3470 b and HD 97658 b) are not expected to be observed by TESS. GJ 3470 b has an ecliptic latitude of ∼ −5◦, placing it outside of TESS's anticipated field of view. HD 97658 b has an ecliptic latitude of ∼ +19◦, but it falls in a field of view gap between adjacent 10 Dalba et al. 1.4 1.2 1.0 0.8 0.6 0.4 0.2 y t i s n e D y t i l i b a b o r P 0.0 10 Simulation Normal 11 12 13 14 Number of Transiting RV Exoplanets to be Observed by TESS Figure 4. Normalized histogram of NTESS, the predicted number of known RV exoplanets that will observed to transit their host stars by TESS. The solid line is a normal distribution with the same mean and standard deviation as the distribution of NTESS values. Based on our Monte Carlo simulation of Equation (4), we expect that TESS will observe transits from 11.7+0.3 −0.3 known RV exoplanets during its primary mission. Accounting for the RV exoplanets that TESS will observe that are already known to transit reduces NTESS to ∼3 exoplanets (see the text). sectors5. The other two RV exoplanets already known to transit (30 Ari B b and HD 80606 b) have (τ /P ) ratios of 0.08 and 0.24, respectively, so their respective TESS transit probabilities ptransit,TESS also equal 0.08 and 0.24. The contribution of RV exoplanets that are already known to transit can be removed from the predicted value of NTESS by subtracting 8 to account for the "guaranteed" transits and by subtracting an additional 0.32 to account for the ptransit,TESS values of 30 Ari B b and HD 80606 b. In summary, during its primary mission, we predict that TESS will discover transits of ∼ 3 known RV exoplanets that were not previously known to transit. The low value of NTESS can be attributed to the long-period nature of the sample of known RV exoplanets, especially when contrasted with the relatively short observational baseline that most of the TESS fields will receive. The median and mean orbital periods of our exoplanet sample are 381 days and 867 days, respectively, demonstrating the distribution is significantly skewed toward long periods. Of the 673 known RV exoplanets in our sample, 125 have (τ /P ) ≥ 1, meaning that the TESS baseline does not reduce the probability of observing a transit. For these exoplanets, TESS will be able to confidently refute or confirm the transiting nature of the system. This will represent a substantial contribution to ephemeris refinement efforts such as TERMS (Kane et al. 2009), and one that would otherwise necessitate a tremendous amount of ground-based resources. TESS's contribution to ephemeris refinement will be especially valuable for exoplanets with long 5 If the actual Cycle 2 sector boundaries are defined differently than we have assumed here (in Section 5) -- such that HD 97658 is in the field of view of TESS -- the transit of HD 97658 b (with P ≈ 9.5 days) will be observed. TESS Yield of Transits from Known Radial Velocity Exoplanets 11 120 100 80 60 40 20 r e b m u N 0 100 101 102 103 104 105 106 Predicted Transit Depth [ppm] Figure 5. Histogram of the predicted transit depth as approximated by (Rp/R⋆)2 for the known RV exoplanets in our sample. Rp values are estimated using measured exoplanet masses and the mass-radius relations of Chen & Kipping (2017). R⋆ values are either measured or estimated as described in Section 3.1. The majority of this sample have predicted transit depths of ∼1%, suggesting that the sample is composed mainly of giant exoplanets. periods. The long-period end of 125 exoplanets with (τ /P ) ≥ 1 includes HD 40307 g (P ≈ 198 days, ptransit,TESS ≈ 0.95%), HD 65216 c (P ≈ 153 days, ptransit,TESS ≈ 0.83%), and HD 156279 b (P ≈ 131 days, ptransit,TESS ≈ 3.5%) among others (see Table 1). 5.1. Signal-to-noise Considerations for TESS Our prediction for the number of known RV exoplanets for which TESS will discover transits has so far lacked any consideration of the depths of the potential transits, the brightness of the host stars, and the anticipated precision of the TESS photometry. Here, we briefly consider these factors in relation to detecting transits of RV exoplanets in TESS observations. In Figure 5, we show the predicted distribution of transit depths for all RV exoplanets in our sample. We approximate transit depth as (Rp/R⋆)2, where Rp and R⋆ are determined as described in Section 3. The distribution of the predicted transit depths spans ∼2 -- 2× 105 parts per million (ppm) and peaks near 1× 104 ppm, which is the equivalent of Jupiter transiting the Sun. This is expected as our sample of RV exoplanets is known to contain mostly giant planets. It is expected that 1% precision should be readily achievable for most TESS targets. However, we investigate the TESS detection efficiency of the known RV exoplanet sample more carefully by estimating the signal-to-noise ratio (SNRTESS) of their potential transit signals. First, we assume circular, edge-on transits so that the transit duration (tdur) is given by (Winn 2010) tdur = P π arcsin(cid:20) R⋆ a (cid:18)1 + Rp R⋆(cid:19)(cid:21) . (5) 12 Dalba et al. The transit duration is evaluated with the same parameters used previously to estimate transit prob- ability. Then, for a transit signal containing N transits, we estimate SNRTESS following Barclay et al. (2018): SNRTESS = (Rp/R⋆)2 σTESS (6) √N tdur (1 + d) where d is the dilution of the transit signal due to nearby sources and σTESS is the photometric noise model for TESS from Figure 14 of Sullivan et al. (2015). Predicted values of d are acquired from the TESS Candidate Target List (Stassun et al. 2018) in the "contratio" column. In Equation (6), tdur has units of hr and σTESS has units of hr1/2. We estimate the photometric noise for our sources using their TESS-band magnitudes as a proxy for the standard Johnson -- Cousins IC band. In calculating SNRTESS, we consider the total amount of observing time that a star is expected to receive by defining N as the integer value of τ /P (rounding down). For 0 < (τ /P ) < 1, we ascribe N = 1, as if a single transit is detected. The purpose of this calculation is not to consider the probability of such a transit detection, but instead to compare its signal to the noise assuming a transit was to occur. Figure 6 shows the predicted distribution of SNRTESS for the all of the known RV exoplanets that are expected to be observed by TESS. We compare this distribution to the nominal detection threshold of 7.3, which was derived by Sullivan et al. (2015) for TESS following a similar calculation for the original Kepler mission (Jenkins et al. 2010). Of the known RV exoplanets expected to be observed by TESS, 93% are predicted to have SNRTESS ≥ 7.3. Alternatively, eight exoplanets are expected to have SNRTESS < 1. These include the five known exoplanets orbiting GJ 667 C (which suffers from extreme flux dilution) and three others orbiting host stars with anomalously large radii values listed in the NEA (i.e., HD 208527, BD+20 2457, and HD 96127). Since the vast majority of RV exoplanets would likely produce transit signals that are significantly larger than TESS's detection threshold, it is unlikely that the yield of transits from the RV exoplanets would be significantly reduced due to insufficient photometric precision. 5.2. Potential TESS Extended Missions The long-term stability of TESS's lunar resonance orbit (Gangestad et al. 2013) coupled with the anticipated use of spacecraft fuel makes an extended mission a realistic possibility. Here, we consider how the predicted number of transits of known RV exoplanets increases for three potential extended mission scenarios. The scenarios discussed here are by no means an exhaustive list of all possibilities. In this discussion, we continue to only use orbital periods and ignore the other orbital ephemerides when estimating TESS transit probabilities (see Section 6). The first strategy for an extended mission involves obtaining a long observational baseline over a rel- ative small portion of the sky. Scenarios such as "pole" (Bouma et al. 2017) or "C3PO" (Huang et al. 2018) achieve this by pointing TESS toward either of the ecliptic poles for timescales of a year or longer. As displayed in Figure 3, the known RV exoplanets are distributed approximately evenly across the sky. Therefore, focusing observations on a small area would likely minimize the number of transits detected from known RV exoplanets. This strategy perhaps allows for the recovery of transits from known RV exoplanets with long orbital periods, up to the observational baseline (see Section 7). However, the transit probability for long-period exoplanets is inherently lower, meaning that a transit is less likely to occur regardless of how long the host star is observed. TESS Yield of Transits from Known Radial Velocity Exoplanets 13 SNR = 7.3 102 r e b m u N 101 100 100 101 10-2 104 10-3 Predicted SNR of Transit Signal in TESS Observations 10-1 102 103 Figure 6. Histogram of the predicted SNR of the potential transit signals of the known RV exoplanets. Only exoplanet hosts that are expected to be observed by TESS (i.e., τ > 0 days) are included in this distribution. The vertical dashed line is plotted at SNR=7.3, the detection threshold adopted by Sullivan et al. (2015). 93% of the plotted distribution has SNR values above 7.3, suggesting that photometric precision is unlikely to reduce TESS's ability to recover transits of known RV exoplanets. On the other hand, the second strategy for an extended mission involves repeating the nominal primary mission: dividing the ecliptic hemisphere into 13 sectors, observing each sector for ∼27 days, and repeating for the opposite ecliptic hemisphere (e.g., "hemi," Bouma et al. 2017; Huang et al. 2018). For the ith known RV exoplanet, this strategy yields an extended observational baseline τext,i that satisfies τext,i = (1 + fext,i)τi (7) where fext,i is the number of times during the extended mission that TESS surveys the ecliptic hemisphere containing the ith exoplanet. By substituting τext,i into Equation (4), we find that ptransit,TESS increases by as much as a factor of (1 + fext,i), so long as (τext,i/Pi) ≤ 1. If this extended mission strategy is employed for only one year in the southern ecliptic hemisphere, then the total value of NTESS (including both the primary and extended missions) would increase to 12.2+0.3 −0.3 exoplanets. The same calculation for a one-year extended mission facing the northern ecliptic hemisphere yields NTESS = 12.6+0.3 −0.3 exoplanets. Following this strategy for a six-year extended mission (three years facing south, three years facing north) would yield NTESS = 15.0+0.3 −0.3. We find that, on average, the "hemi" extended mission scenario leads to the discovery of transits for roughly one RV exoplanet per year of the extended mission. The third strategy for an extended mission involves observing both ecliptic hemispheres over the course of one year using the same pointing as the nominal primary mission, but reducing the sector baseline by 50%, to ∼14 days each (e.g.,"allsky," Bouma et al. 2017; Huang et al. 2018). For the ith known RV exoplanet, this strategy yields an extended observational baseline τext,i that satisfies τext,i = (1 + 0.5fext,i)τi , (8) 14 Dalba et al. Table 1. Transit probabilities and predicted TESS observation details for the known RV exoplanets in our sample b ptransit 0.5+0.5 −0.5 0.4+0.4 −0.4 Exoplanet Name TIC IDa NGC 2682 Sand 364 b 294684871 99734092 BD+20 2457 b 332064670 0.291+0.005 55 Cnc e −0.004 323919676 HD 240237 b 428673146 HD 102956 b HD 86081 b 27073220 441712711 24 Boo b 268634769 HD 96127 b 54237857 HD 88133 b HD 24064 b 286923464 120960812 HD 118203 b 20096620 GJ 674 b HD 1690 b 37763102 68418284 WASP-94 B b 161314759 HD 212301 b HD 13189 b 63564248 148563075 BD+20 2457 c 0.29+0.1 −0.08 0.28+0.03 −0.03 0.27+0.03 −0.03 0.26+0.03 −0.02 0.2+0.1 −0.1 0.23+0.03 −0.03 0.22+0.04 −0.03 0.21+0.02 −0.02 0.2+0.2 −0.2 0.21+0.07 −0.06 0.21+0.02 −0.02 0.2+0.02 −0.02 0.2+0.03 −0.03 0.2+0.2 −0.2 TESS Cycle 1 TESS Cycle 2c ptransit,TESS b Sector(s) Camera(s) Sector(s) Camera(s) ··· 0.03+0.03 −0.03 0.291+0.005 −0.004 0.021+0.007 −0.006 ··· 0.27+0.03 −0.03 0.26+0.03 −0.02 0.01+0.005 −0.005 0.23+0.03 −0.03 0.022+0.004 −0.003 0.21+0.02 −0.02 0.2+0.2 −0.2 0.01+0.003 −0.003 0.21+0.02 −0.02 0.2+0.02 −0.02 0.008+0.009 −0.008 ··· ... ··· ··· ··· ··· ··· 8 ··· ··· ··· ··· ··· 12 3 1 1,13 ··· ··· ... ··· ··· ··· ··· ··· 1 ··· ··· ··· ··· ··· 1 1 1 3,3 ··· ··· ... ··· 8 7 3,4 ··· ··· 9,10 8 8 5,6 9 ··· ··· ··· ··· ··· 8 ... ··· 1 1 3,3 ··· ··· 3,3 2 1 2,2 3 ··· ··· ··· ··· ··· 1 ... ... ... ... Note -- Table 1 will be published in its entirety in the machine-readable format as Supplementary Data. aUnique identifier from the TESS Input Catalog (TIC, Stassun et al. 2018). bTransit probabilities and uncertainties are approximated using the 16th, 50th, and 84th percentiles of their probability distributions. cCycle 2 sectors and cameras are determined using the ecliptic longitudes and latitudes of Cycle 1 mirrored to the northern ecliptic hemisphere (see the text). which only differs from the "hemi" strategy if the extended mission lasts for an odd number of years. In the event of a single-year extended mission that follows the "allsky" strategy, we estimate that NTESS would increase to 12.5+0.3 −0.3. In the interest of detecting transits from known RV exoplanets that are currently not known to transit, TESS extended mission scenarios that provide broad sky coverage, as opposed to long obser- vational baselines, are preferred. For a single or multi-year extended mission, repeating the nominal primary mission pointing with the same sector durations or reduced sector durations produce similar results. Although the predicted number of transiting RV exoplanets rises slowly as a function of extended mission duration, the number of RV exoplanets for which we can rule out transits will increase more rapidly. These dispositive null results will comprise a valuable portion of the science return for known RV exoplanets from the TESS mission. TESS Yield of Transits from Known Radial Velocity Exoplanets 15 6. DISCUSSION In the following sections, we describe several of the simplifying assumptions we employed in our prediction of the yield of transiting RV exoplanets. We also evaluate the significance of these as- sumptions and any effects they may have had on the results. 6.1. Coplanarity of Multi-planet Systems Throughout this work, the transit probability of each exoplanet is estimated independently. That is, transit probabilities of exoplanets in multi-planet systems do not take into account the coplanarity of the system or the exoplanets' mutual inclinations. This simplifying assumption merits mention as our sample consists of 100 multi-planet systems comprising 254 total exoplanets. A more accurate approach would be to convolve the transit probabilities within a system of exoplanets with the likelihood of coplanarity (Brakensiek & Ragozzine 2016). Depending on a particular system's properties, considering coplanarity may either increase or de- crease the transit probabilities of the members of that system. When considered for the full sample, we do not expect coplanarity to significantly alter the predicted number (N) of RV exoplanets that transit or the predicted number (NTESS) that TESS is expected to observe. However, our choice to ignore coplanarity should be considered when interpreting the transit probabilities of exoplanets belonging to multi-planets systems as listed in Table 1. 6.2. Orbital Inclination Bias of the RV Method As is true for the transit method, the detection efficiency of the RV method of exoplanet discovery increases for more edge-on orbital inclinations. This means that the sample of RV exoplanets dis- cussed in this work is biased toward edge-on orbits. The calculation of transit probabilities according to Equations (1) and (4) ignore this bias in favor of the simplicity offered by the assumption of random inclinations. Therefore, the predictions for the number (N) of RV exoplanets that transit and the number (NTESS) of RV exoplanets that TESS should see in transit are lower limits. There is an additional subtlety to the inclination bias of the RV method. For a given RV precision, a more massive exoplanet is less affected by the inclination bias than a less massive exoplanet because the RV detection efficiency also scales with planet mass. Therefore, less massive (e.g., several M⊕) RV exoplanets are more likely to have edge-on inclinations than more massive (e.g., several hundred M⊕) exoplanets. This subtle point further demonstrates that our predictions for N and NTESS are lower limits. However, low mass exoplanets comprise only a minor portion of our sample of RV exoplanets, so this bias has a similarly minor effect. 6.3. Prior versus Posterior Transit Probability for RV Exoplanets The transit probability of an exoplanet detected via the RV method depends on the prior probability density of exoplanet mass as well as prior probability density of orbital inclination (e.g., Ho & Turner 2011). Because Equation (1) does not consider the a priori mass distribution of exoplanet, it does not represent the a posteriori transit probability of an RV exoplanet. Stevens & Gaudi (2013) provided a thorough investigation of posterior transit probabilities that included the derivation of multiplicative factors useful for determining posterior transit probability from prior transit probability. The factors, however, required the assumption of upper limits for exoplanet masses, which we do not have for our sample of known RV exoplanets. Instead of introducing uncertainty through the assumption of 16 Dalba et al. mass upper limits, we chose to accept the approximation of Equation (1) as the RV exoplanet transit probability. The impact of this choice can be evaluated qualitatively using Table 2 of Stevens & Gaudi (2013). Dividing our sample of RV exoplanets into mass regimes based on their measured minimum masses, we find that 12% are "Earth/Super-Earths" (0.1 M⊕ -- 10 M⊕), 13% are "Neptunes" (10 M⊕ -- 100 M⊕), 48% are "Jupiters" (100 M⊕ -- 103 M⊕), 23% are "Super-Jupiters" (103 M⊕ -- 13 MJupiter), and 4% are "Brown Dwarfs" (13 MJupiter -- 0.07 M⊙). The posterior transit probabilities for the Earth/Super- Earth, Neptune, and Super-Jupiter regimes (i.e., 48% of our total sample) are larger than the prior transit probabilities by 12 -- 19% on average. The posterior transit probabilities for the Jupiter regime (i.e., 48.0% of our total sample) are only 1% smaller than the prior transit probabilities on average. The posterior transit probabilities for the Brown Dwarf regime (i.e., 4% of our total sample) are 14% smaller than the prior transit probabilities on average. Had we included the a priori distribution of exoplanet masses in the calculation of transit probability, the vast majority of transit probabilities would have either stayed nearly the same or increased. This caveat again demonstrates that our predictions for N and NTESS are lower limits. 6.4. Ignoring RV Exoplanet Ephemerides In determining the fraction of an exoplanet's orbital phase sampled by TESS observation, we neglect all ephemeris information except for orbital period. This choice is motivated by the lack of precision in the current ephemerides of most RV exoplanets. Figure 2 of Kane et al. (2009) demonstrated that, for orbital periods longer than ∼100 days, the size of the exoplanet's transit window was comparable to its orbital period. Hence, the positions of those exoplanets in their orbits were essentially unknown. In the intervening 10 years since that study, most of the RV exoplanets have not received follow-up observation, so the transit windows have become even longer. Only a small handful of RV exoplanets (e.g., the TERMS targets mentioned in Section 4) have been observed sufficiently frequently and recently to have precise ephemerides that enable a meaningful comparison between the transit windows and the dates of the TESS sectors. Therefore, for consistency, we have assumed that the portions of all exoplanet orbits sampled by TESS observations are random. 7. SUMMARY AND CONCLUSIONS We use published stellar, orbital, and planetary parameters to estimate the transit probabilities for nearly all exoplanets that have been discovered via the RV method. Monte Carlo simulations of these transit probabilities enable the prediction that 25.5+0.7 −0.7 of the RV exoplanets in our sample have the proper orbital inclination to transit their host stars (Figure 2). At the time of writing, the number of exoplanets in our sample that are known to transit is only 12, which is more than 19 standard deviations below the expectation. Hence, we predict that follow-up photometric observations of known RV exoplanets are likely to yield transit discoveries. The ability of the transit-hunting TESS mission to observe transits of known RV exoplanets is also considered. The vast majority (93%) of RV exoplanets have predicted transit signals with SNR greater than the nominal detection threshold of 7.3 (Sullivan et al. 2015), suggesting that TESS is a suitable observatory for follow-up (Figure 6). We identify the known RV exoplanets that TESS is likely to observe (Figure 3) and recalculate their transit probabilities to account for TESS's finite observational baseline. Monte Carlo simulations of these updated transit probabilities enable the prediction that TESS will observe transits for 11.7+0.3 −0.3 of the known RV exoplanets in our sample TESS Yield of Transits from Known Radial Velocity Exoplanets 17 (Figure 4). Of these, only ∼3 exoplanets will likely not have been known to transit previously. We find that our prediction is robust to changes in the ecliptic longitudes of TESS's sector boundaries. We also expect TESS to be able to confidently rule out transits for ∼125 known RV exoplanets that have orbital periods that are less than or equal to the anticipated baselines of TESS observations. Transit probabilities and predicted TESS observation information for each exoplanet in our sample are provided in Table 1. We consider several possible scenarios for a TESS extended mission in regard to following up known RV exoplanets. The distribution of known RV exoplanets on the sky is essentially even, so extended missions that emphasize all-sky coverage will maximize the number of orbital ephemerides constraints for known RV exoplanets. We find that a multi-year extended mission with an all-sky observing strategy can increase the total number of RV exoplanets that are found to transit, but only by roughly one exoplanet per year. This increase is not strongly dependent on the specific all-sky observing strategy (i.e., repeating the strategy of the primary mission or observing the full sky each year). Alternatively, an extended mission scenario that focuses on a smaller portion of sky (e.g., an ecliptic pole) in order to acquire a longer baseline would also complement efforts to follow-up known RV exoplanets. Indeed, 23 of the known RV exoplanets in our sample have ecliptic latitudes within 13◦ of a pole, and 13 of these have orbital periods longer than 100 days. Detecting a transit of a known long-period RV exoplanet would be an exceptionally valuable, albeit rare, discovery. Having a measured mass and radius, this exoplanet could extend relations of planet density versus stellar insolation and potentially serve as a solar system analog. During a subsequent transit, its transmission spectrum could also be observed, an endeavor that would likely prove fruitful for terrestrial (e.g., Robinson & Reinhard 2018) or giant (e.g., Dalba et al. 2015) long-period exoplanets alike. The observational strategy employed by the TESS primary mission is largely geared toward short-period exoplanets. This is evident by the relatively small number of known RV exoplanets that we predict TESS will observe in transit. An extended mission that focuses on longer period exoplanets, both new and previously known, would likely yield a valuable and complementary science return. We wish to thank the anonymous referee for thoughtful suggestions that improved this work. This research made use of the NASA Exoplanet Archive, which is operated by the California Institute of Technology, under contract with the National Aeronautics and Space Administration under the Exoplanet Exploration Program. Some of the data presented in this paper were obtained from the Mikulski Archive for Space Telescopes (MAST). STScI is operated by the Association of Universities for Research in Astronomy, Inc., under NASA contract NAS5-26555. T. L. Campante acknowledges support from the European Union's Horizon 2020 research and innovation program under the Marie Sk lodowska-Curie grant agreement No. 792848. Funding for the TESS mission is provided by NASA's Science Mission directorate. P. A. Dalba and S. R. Kane acknowledge financial support from the NASA Exoplanet Research Program through grant 17-XRP17 2-0148. 18 Dalba et al. Software: forecaster(Chen & Kipping2017),tvguide(Mukai & Barclay2017),astropy(Astropy Collaboration et al. 2013) REFERENCES Astropy Collaboration, Robitaille, T. P., Tollerud, Fabrycky, D. C., Lissauer, J. J., Ragozzine, D., E. J., et al. 2013, A&A, 558, A33, doi: 10.1051/0004-6361/201322068 Azzalini, A. 1985, Scandinavian Journal of Statistics, 12, 171, doi: 10.2307/4615982 Bakos, G., Noyes, R. W., Kov´acs, G., et al. 2004, PASP, 116, 266, doi: 10.1086/382735 Ballard, S. 2018, ArXiv e-prints. https://arxiv.org/abs/1801.04949 Barclay, T., Pepper, J., & Quintana, E. V. 2018, ApJS, 239, 2, doi: 10.3847/1538-4365/aae3e9 Beatty, T. G., & Gaudi, B. S. 2008, ApJ, 686, 1302, doi: 10.1086/591441 et al. 2014, ApJ, 790, 146, doi: 10.1088/0004-637X/790/2/146 Gangestad, J. W., Henning, G. A., Persinger, R. R., & Ricker, G. R. 2013, ArXiv e-prints. https://arxiv.org/abs/1306.5333 Garaud, P., & Lin, D. N. C. 2007, ApJ, 654, 606, doi: 10.1086/509041 Guo, X., Johnson, J. A., Mann, A. W., et al. 2017, ApJ, 838, 25, doi: 10.3847/1538-4357/aa6004 Henry, G. W., Kane, S. R., Wang, S. X., et al. 2013, ApJ, 768, 155, doi: 10.1088/0004-637X/768/2/155 Blank, D. L., Feliz, D., Collins, K. A., et al. 2018, Hinkel, N. R., Kane, S. R., Henry, G. W., et al. AJ, 155, 228, doi: 10.3847/1538-3881/aabded Borucki, W. J., Koch, D., Basri, G., et al. 2010, Science, 327, 977, doi: 10.1126/science.1185402 Bouma, L. G., Winn, J. N., Kosiarek, J., & McCullough, P. R. 2017, ArXiv e-prints. https://arxiv.org/abs/1705.08891 2015a, ApJ, 803, 8, doi: 10.1088/0004-637X/803/1/8 Hinkel, N. R., Kane, S. R., Pilyavsky, G., et al. 2015b, AJ, 150, 169, doi: 10.1088/0004-6256/150/6/169 Ho, S., & Turner, E. L. 2011, ApJ, 739, 26, Brakensiek, J., & Ragozzine, D. 2016, ApJ, 821, doi: 10.1088/0004-637X/739/1/26 47, doi: 10.3847/0004-637X/821/1/47 Howard, A. W., Johnson, J. A., Marcy, G. W., Broeg, C., Fortier, A., Ehrenreich, D., et al. 2013, EPJ Web of Conferences, 47, 03005, doi: 10.1051/epjconf/20134703005 Chen, J., & Kipping, D. 2017, ApJ, 834, 17, doi: 10.3847/1538-4357/834/1/17 Cox, A. N. 2000, Allen's Astrophysical Quantities (New York: Springer Science + Business Media) Dalba, P. A., & Muirhead, P. S. 2016, ApJL, 826, et al. 2010, ApJ, 721, 1467, doi: 10.1088/0004-637X/721/2/1467 Howard, A. W., Marcy, G. W., Bryson, S. T., et al. 2012, ApJS, 201, 15, doi: 10.1088/0067-0049/201/2/15 Huang, C. X., Shporer, A., Dragomir, D., et al. 2018, ArXiv e-prints. https://arxiv.org/abs/1807.11129 L7, doi: 10.3847/2041-8205/826/1/L7 Jenkins, J. M., Caldwell, D. A., Chandrasekaran, Dalba, P. A., Muirhead, P. S., Fortney, J. J., et al. 2015, ApJ, 814, 154, doi: 10.1088/0004-637X/814/2/154 H., et al. 2010, ApJL, 713, L87, doi: 10.1088/2041-8205/713/2/L87 Kane, S. R. 2007, MNRAS, 380, 1488, Dawson, R. I., & Murray-Clay, R. A. 2013, ApJL, doi: 10.1111/j.1365-2966.2007.12144.x 767, L24, doi: 10.1088/2041-8205/767/2/L24 Dragomir, D., Matthews, J. M., Kuschnig, R., et al. 2012a, ApJ, 759, 2, doi: 10.1088/0004-637X/759/1/2 Dragomir, D., Kane, S. R., Henry, G. W., et al. 2012b, ApJ, 754, 37, doi: 10.1088/0004-637X/754/1/37 Kane, S. R., Henry, G. W., Dragomir, D., et al. 2011a, ApJL, 735, L41, doi: 10.1088/2041-8205/735/2/L41 Kane, S. R., Mahadevan, S., von Braun, K., Laughlin, G., & Ciardi, D. R. 2009, PASP, 121, 1386, doi: 10.1086/648564 Kane, S. R., Dragomir, D., Ciardi, D. R., et al. Espinoza, N., & Jord´an, A. 2015, MNRAS, 450, 1879, doi: 10.1093/mnras/stv744 2011b, ApJ, 737, 58, doi: 10.1088/0004-637X/737/2/58 TESS Yield of Transits from Known Radial Velocity Exoplanets 19 Kane, S. R., Howard, A. W., Pilyavsky, G., et al. Ricker, G. R., Winn, J. N., Vanderspek, R., et al. 2011c, ApJ, 733, 28, doi: 10.1088/0004-637X/733/1/28 Kane, S. R., Boyajian, T. S., Henry, G. W., et al. 2015, ApJ, 806, 60, doi: 10.1088/0004-637X/806/1/60 Kane, S. R., Wittenmyer, R. A., Hinkel, N. R., et al. 2016, ApJ, 821, 65, doi: 10.3847/0004-637X/821/1/65 Kipping, D. M. 2013, MNRAS, 434, L51, doi: 10.1093/mnrasl/slt075 2015, JATIS, 1, 014003, doi: 10.1117/1.JATIS.1.1.014003 Robinson, T. D., & Reinhard, C. T. 2018, ArXiv e-prints. https://arxiv.org/abs/1804.04138 Santerne, A., Moutou, C., Tsantaki, M., et al. 2016, A&A, 587, A64, doi: 10.1051/0004-6361/201527329 Shankland, P. D., Rivera, E. J., Laughlin, G., et al. 2006, ApJ, 653, 700, doi: 10.1086/508562 Smalley, B. 2005, MSAIS, 8, 130 Kipping, D. M., Cameron, C., Hartman, J. D., Stassun, K. G., Oelkers, R. J., Pepper, J., et al. et al. 2017, AJ, 153, 93, doi: 10.3847/1538-3881/153/3/93 2018, AJ, 156, 102, doi: 10.3847/1538-3881/aad050 L´opez-Morales, M., Morrell, N. I., Butler, R. P., & Stevens, D. J., & Gaudi, B. S. 2013, PASP, 125, Seager, S. 2006, PASP, 118, 1506, doi: 10.1086/508904 Miller, N., & Fortney, J. J. 2011, ApJL, 736, L29, doi: 10.1088/2041-8205/736/2/L29 Moutou, C., H´ebrard, G., Bouchy, F., et al. 2009, A&A, 498, L5, doi: 10.1051/0004-6361/200911954 Mukai, K., & Barclay, T. 2017, tvguide: A tool for determining whether stars and galaxies are observable by TESS, 1.0, https://github.com/tessgi/tvguide Oberg, K. I., Murray-Clay, R., & Bergin, E. A. 2011, ApJL, 743, L16, doi: 10.1088/2041-8205/743/1/L16 Pepe, F., Mayor, M., Queloz, D., et al. 2004, A&A, 423, 385, doi: 10.1051/0004-6361:20040389 Pepper, J., Gould, A., & Depoy, D. L. 2003, AcA, 53, 213 Pilyavsky, G., Mahadevan, S., Kane, S. R., et al. 2011, ApJ, 743, 162, doi: 10.1088/0004-637X/743/2/162 Pollacco, D. L., Skillen, I., Collier Cameron, A., et al. 2006, PASP, 118, 1407, doi: 10.1086/508556 933, doi: 10.1086/672572 Sullivan, P. W., Winn, J. N., Berta-Thompson, Z. K., et al. 2015, ApJ, 809, 77, doi: 10.1088/0004-637X/809/1/77 Thorngren, D. P., Fortney, J. J., Murray-Clay, R. A., & Lopez, E. D. 2016, ApJ, 831, 64, doi: 10.3847/0004-637X/831/1/64 Twicken, J. D., Jenkins, J. M., Seader, S. E., et al. 2016, AJ, 152, 158, doi: 10.3847/0004-6256/152/6/158 Wang, S., Wu, D.-H., Addison, B. C., et al. 2018, AJ, 155, 73, doi: 10.3847/1538-3881/aaa253 Wang, Sharon, X., Wright, J. T., Cochran, W., et al. 2012, ApJ, 761, 46, doi: 10.1088/0004-637X/761/1/46 Weiss, L. M., & Marcy, G. W. 2014, ApJL, 783, L6, doi: 10.1088/2041-8205/783/1/L6 Winn, J. N. 2010, in Exoplanets, ed. S. Seager (Tucson: Univ. of Arizona Press), 55 -- 77 Wright, J. T., Marcy, G. W., Howard, A. W., et al. 2012, ApJ, 753, 160, doi: 10.1088/0004-637X/753/2/160 Yi, J. S., Chen, J., & Kipping, D. 2018, MNRAS, 475, 3090, doi: 10.1093/mnras/sty102
1910.07146
1
1910
2019-10-16T03:24:32
Asteroid Discovery and Light Curve Extraction Using the Hough Transform -- A Rotation Period Study for Sub-Kilometer Main-Belt Asteroids
[ "astro-ph.EP" ]
The intra-night trajectories of asteroids can be approximated by straight lines, and so are their intra-night detections. Therefore, the Hough transform, a line detecting algorithm, can be used to connect the line-up detections to find asteroids. We applied this algorithm to a high-cadence Pan-STARRS 1 (PS1) observation, which was originally designed to collect asteroid light curves for rotation period measurements (Chang et al., 2019). The algorithm recovered most of the known asteroids in the observing fields and, moreover, discovered 3574 new asteroids with magnitude mainly of 21.5 < w_p1 < 22.5 mag. This magnitude range is equivalent to sub-kilometer main-belt asteroids (MBAs), which usually lack of rotation period measurements due to their faintness. Using the light curves of the 3574 new asteroids, we obtained 122 reliable rotation periods, of which 13 are super-fast rotators (SRFs; i.e., rotation period of < 2 hr). The required cohesion to survive these SFRs range from tens to thousands of Pa, a value consistent with the known SFRs and the regolith on the Moon and Mars. The higher chance of discovering SFRs here suggests that sub-kilometer MBAs probably harbor more SFRs.
astro-ph.EP
astro-ph
Asteroid Discovery and Light Curve Extraction Using the Hough Transform -- A Rotation Period Study for Sub-Kilometer Main-Belt Asteroids Kai-Jie Lo1, Chan-Kao Chang2, Hsing-Wen Lin3, Meng-Feng Tsai1, Wing-Huen Ip2,4, Wen-Ping Chen2, Ting-Shuo Yeh2, K. C. Chambers4, E. A. Magnier4, M. E. Huber4, R. J. Wainscoat4 [email protected] Received ; accepted 1Department of Computer Science and Information Engineering, National Central Uni- versity, Jhongli, Taiwan 2Institute of Astronomy, National Central University, Jhongli, Taiwan 3Department of Physics, University of Michigan, Ann Arbor, Michigan 48109, USA 4Space Science Institute, Macau University of Science and Technology, Macau 5Institute for Astronomy, University of Hawaii, Honolulu, Hawaii 96822, USA -- 2 -- ABSTRACT The intra-night trajectories of asteroids can be approximated by straight lines, and so are their intra-night detections. Therefore, the Hough transform, a line de- tecting algorithm, can be used to connect the line-up detections to find asteroids. We applied this algorithm to a high-cadence Pan-STARRS 1 (PS1) observation, which was originally designed to collect asteroid light curves for rotation period measurements (Chang et al. 2019). The algorithm recovered most of the known asteroids in the observing fields and, moreover, discovered 3574 new asteroids with magnitude mainly of 21.5 < wp1 < 22.5 mag. This magnitude range is equivalent to sub-kilometer main-belt asteroids (MBAs), which usually lack of rotation period measurements due to their faintness. Using the light curves of the 3574 new asteroids, we obtained 122 reliable rotation periods, of which 13 are super-fast rotators (SRFs; i.e., rotation period of < 2 hr). The required cohesion to survive these SFRs range from tens to thousands of P a, a value consistent with the known SFRs and the regolith on the Moon and Mars. The higher chance of discovering SFRs here suggests that sub-kilometer MBAs probably harbor more SFRs. Subject headings: surveys - minor planets, asteroids: general -- 3 -- 1. Introduction Finding asteroids is the fundamental step for asteroid research. With the advances in wide-field cameras, robotic observations, and computing resources, several telescopes have been dedicated to asteroid discovery, such as the Catalina Sky Survey (CSS)1, the Pan-STARRS 1 (Chambers et al. 2016), the Asteroid Terrestrial-impact Last Alert System (ATLAS; Tonry et al. 2018), and the Chinese Near-Earth-Object Survey Telescope (CNEOST)2. To optimize the discovery rate of asteroid, each project customizes its survey rates (i.e., scanning sky area over a period of observation time) to a certain cadence that meets the basic requirement of discovering an asteroid, such as three detections within a night or several detections over a few nights for a single asteroid discovery. Therefore, each project develops its own asteroid-detecting pipeline and optimizes it to the scientific interests and survey strategy. In some cases, the pipeline does not utilize all the detections for some reasons. The most common case is a filter to exclude the low signal-to-noise ratio detections, which usually contain many noises. Nevertheless, if a survey was conducted using a relatively short cadence, asteroids could be much easier to be distinguished from the background noises and field stars. A good example of high-cadence observation is the rotation period survey for MBAs carried out in October, 2016 using the PS1 (Chang et al. 2019), in which eight PS1 fields were continuously scanned using ∼ 10 min cadence over six straight nights. From the survey, the intra-night detections of asteroids appeared as straight lines due to their approximately linear motion during the short period of time, which obviously stand out from the stationary sources and noises. Therefore, we can make use of the Hough Transform to connect these intra-night line-up detections to find asteroids. The Hough Transform was originally proposed by Hough (1959) and then further utilized by 1http://www.lpl.arizona.edu/css/ 2http://www.cneost.org/ -- 4 -- Duda & Hart (1972, i.e., generalized Hough transform). This method has been widely used to extract the linear features on two-dimension images and recently started to be applied to astronomical images to detect cosmic rays (Keys, & Pevtsov 2010), streaks produced from fast-moving small solar system bodies (SSSBs) (Virtanen et al. 2014; Kim 2016; Bektesevi´c, & Vinkovi´c 2017; Nir et al. 2018), and long tracks of space debris (Hickson 2018). Instead of working on images, we applied the Hough Transform to the source catalogs obtained from the aforementioned intra-night observations. The implementation of the Hough Transform is relatively simple and, moreover, it does not have to exclude low signal-to-noise detections in advance. Therefore, we were able to utilize the data set much closer to its actual limiting magnitude (i.e., ∼ 22.5 mag) and found more than three thousand asteroids mainly of 21.5 < wp1 < 22.5 mag. With this magnitude, a study of rotation period for sub-kilometer MBAs was carried out for the first time3. In this work, the PS1 observation and the data set are briefly described in Section 2, the Hough Transform is illustrated in Section 3, the result of asteroid recovery, discovery, and rotation period analysis are presented in Section 4, and a short summary is given in the end. 2. The PS1 Observation and the Data Set The PS1 was designed to discover small solar system bodies and, especially, those potentially hazardous objects. It is a 1.8 m Ritchey-Chretien reflector located on Haleakala, Maui equipped with the Gigapixel Camera having a field of view of 7 deg2. Six filters are used, including gP 1 (∼ 400 − 550 nm), rP 1 (∼ 550 − 700 nm), iP 1 (∼ 690820 nm), zP 1 3Most sub-kilometer sized or smaller objects with available rotation periods are all near- earth asteroids. -- 5 -- (∼ 820 − 920 nm), yP 1 (> 920 nm), and wP 1 (i.e., combination of gP 1, rP 1, and iP 1), in which the wP 1 filter is especially designed for moving object discovery (Kaiser et al. 2010; Tonry et al. 2012; Chambers et al. 2016). The images obtained by the PS1 are processed by the Image Processing Pipeline (IPP; Chambers et al. 2016; Magnier et al. 2016a,b,c; Waters et al. 2016) to provide source catalogs as the final product. During October 26-31, 2016, a high-cadence observation was carried out using the PS1 to collect a large sample of asteroid light curves for rotation period measurement (Chang et al. 2019). The observation was conducted using a cadence of ∼ 10 minutes to repeatedly scan eight consecutive PS1 fields in wP 1 band. Therefore, each field had ∼ 30 exposures each night, except for the last two nights which only had few exposures were taken due to unstable weather. Moreover, several exposures in each gP 1, rP 1, iP 1, and zP 1 band were also taken in the first night between the exposures of wP 1 band to obtain asteroid colors. Table 1 shows the summary of the observation. Although ∼ 3500 objects, mostly < 21.5 mag, have been discovered and reported to the Minor Planet Center4 using this observation, a lot of unknown objects close to the actual limiting magnitude of ∼ 22.5 have not been found yet. This high-cadence observation is suitable for asteroid discovery using the Hough Transform. Therefore, the source catalogs obtained from the observation were used as our running cases. 3. The Algorithm 3.1. The Hough Transform The intra-night trajectory of an asteroid can be approximated by a straight line, and so do its detections obtained from intra-night observations. Therefore, we can find asteroids 4https://www.minorplanetcenter.net/ -- 6 -- through these line-up intra-night detections. The situation is illustrated in Figure 1 and 2, in which all the detections, obtained from all the source catalogs of the intra-night observations for one particular field, are stacked into a single frame (hereafter, the master frame). On the master frame, we see that the intra-night detections of asteroids on that field appear as straight lines with correct time sequence. Using the Hough Transform, a line on a two dimension image can be expressed in the Hesse normal form as r = x cos θ + y sin θ, (1) where (x, y) is the coordinate of the pixel on the image, r is the distance to the closet point of the line to the reference point, and θ is the angle between this line and the x axis (see Figure 3a). In our case, the (x, y) is the sky coordinate, RA and Dec. When a group of detections belong to the same straight line, they would share common/similar (θ, r) (see Figure 3b). Therefore, we can explore the parameter space of (θ, r) for all the intra-night detections on the master frame (see Figure 3c) and search for those sharing common (θ, r) to locate MBAs (see Figure 3d. 3.2. The Implementation 3.2.1. Detection clean-up and master frame generation Before stacking the intra-night detections of a field to create the mater frame, any detection was first removed from a source catalog if it is on the noisy area of the PS1 detector chip (Waters et al. 2016, see Figure 4 therein) or around any stationary source within 2(cid:48)(cid:48). Then, the remaining intra-night detections were stacked into a master frame. According to the bulk motion of MBAs within 10 minutes, one asteroid detection was expected to have at least another companion on the master frame. Therefore, any detection -- 7 -- on the master frame was further removed if it does not have any other neighbor within 1.5 to 30(cid:48)(cid:48). Typically, each field each night has ∼ 260000 detections left on the master frame. 3.2.2. Hough Transform calculation and segment identification on the master frame After the clean-up step, the Hough Transform, Equation (3.1), was applied to each detection on the master frame, in which r was calculated by exploring θ in a range of 1 to 179 degree with a step of 0.1 degree. Next, we searched for the detections sharing common (θ, r) to identify the line-up detections (hereafter, segment). Here, a common (θ, r) means the same value of θ and the same value of r till the fourth digit after the decimal point (i.e., within 0.0001 degree difference)5. Only the segments with five or more detections were passed to the following processes. Typically, it takes ∼ 25 min to complete this step for a master frame on a computer cluster of 5 nodes6. 3.2.3. Segment combination on the master frame In some cases, the neighboring segments on the master frame share one or more identical detections. When the difference in θ between these segments is less than 0.1 degree, they are combined into a new segment, otherwise only the longest segment was remained. Moreover, two segments on the master frame are combined if they have the 5The 0.0001 degree or 0.36(cid:48)(cid:48) difference in r is considered as the uncertainty of a detection coordinates, which is much smaller than the average seeing, ∼ 1.2(cid:48)(cid:48), during the observations. 6The computer cluster contains 5 identical nodes, in which one is for job distribution and the other four are for computation. Each node has a 4-core Intel [email protected] GHz cup and 16 GB memory. -- 8 -- following situation: (a) their mean locations can be mutually predicted within 7(cid:48)(cid:48), (b) their difference in θ is less than 0.1 degree, and (c) their time spans have no overlap. A detection is further excluded from a segment when it has incorrect epoch in the time sequence (i.e., simple increase/decrease in observation epochs) or has a magnitude 3σ away from the mean magnitude. 3.2.4. Segment Linkage across the master frames To connect segments across different nights and fields, a linkage was performed to all the master frames. If two segments can mutually predict each other's mean locations within 60(cid:48)(cid:48) and have a difference in θ less than 0.1 degree, they are identified as the same asteroid. In the end, the light curves of the remaining segments were extracted for further analysis. The flow chart of the Hoguh Transform implementations is given in Figure 4 3.3. Orbital Determination and Diameter Estimation To determine the preliminary orbits and absolute magnitudes in wP 1 band of the newly discovered asteroids, the Find Orb was adopted7. Figure 5 shows the orbital distributions of the new asteroids along with the known ones in the observing fields. We see that the distributions are similar in the semi-major axis and inclination and, however, the new asteroids has a mean eccentricity slightly smaller than the known. This is because we only have four-straight-night arcs at the opposition, which provides relatively better constraints 7https://www.projectpluto.com/fo.htm. Note that the absolute magnitude was calcu- lated simply assuming a slope of 0.15 in the H − G system (Bowell et al. 1989). -- 9 -- on the semi-major axis and inclination8. To estimate the diameters of the new asteroids, we used D = 10−Hw/5, 1329√ pv (2) where Hw is absolute magnitude in wp1 band, D is diameter in kilometer, pv is geometric albedo in V band, and 1329 is the conversion constant. We assumed pv = 0.20, 0.08, and 0.04 for asteroids in the inner (2.1 < a < 2.5AU ), mid (2.5 < a < 2.8 AU) and outer (a > 2.8 AU) main belts, respectively (Tedesco, Cellino & Zappal´a 2005)9. 4. Asteroid Recovery and Discovery To check the recovery rate of the Hough Transform, we used the known asteroids as of January 6, 2018 to compare with our result. The ephemerides of the known asteroids in the 8The medium uncertainties in our semi-major axis, inclination, and eccentricity are ∼ 7%, ∼ 45%, and ∼ 84%, respectively. 9The conversion from Hw to Hv requires color information, which is not available for our samples. Therefore, Hw was used in Equation (2) instead of Hv. Typically, Hv is ∼ 0.1 to 0.3 mag fainter than Hw for MBAs when applying the transformation listed in Tonry et al. (2012) and assuming a g − r color varying between 0.3 - 0.8 (Ivezi´c et al. 2001). This magnitude difference makes ∼ 10% difference in diameter estimation. In addition, the median uncertainty of our semi-major axes is about ∼ 7%. This makes a difference in Hw ∼ 0.4 mag, corresponding to an additional uncertainty in diameter ∼ 20%. Considering this two factors, a typically uncertainty in our diameter estimation could be ∼ 30 to 40%. However, this uncertainty does not take the assuming albedos into account, which could make several times difference in diameter estimation. -- 10 -- observing fields were obtained from the JPL/HORIZONS system, and then a cross match was performed against the PS1 source catalogs to extract their detections using a radius of 2(cid:48)(cid:48)(hereafter, ephemeris-matching method). If an asteroid have five or more detections, we defined it as a successful recovery. In our observing fields, there were 3870 known asteroids recovered by the ephemeris-matching method, out of which 3819 were also recovered by the Hough Transform. For the asteroids lost by the Hough Transform, we found that the detections of these objects remaind on the master frames are relatively small and, therefore, they have a lower chance to be recovered by the Hough Transform. Two reasons account for this: (a) the detection number is intrinsically small (i.e., ∼ 80% of them have < 20 detections using the ephemeris-matching method), and (b) the majority of the detections have been removed from the master frames in the clean-up step (i.e., ∼ 20% of them have tens of detections matched using the ephemeris-matching method). We then compare the detection number for the known asteroids recovered by the both methods. Figure 6 shows the distribution of the ratio of the detection number recovered by the Hough Transform to that extracted by the ephemeris-matching method, where we see that the Hough Transform has smaller detection number in average. This is because (a) a relatively strict criterion has been applied in the definition of a segment in the Hough Transfrom (i.e., 0.0001 degree difference in r for a segment which is much smaller than the 2(cid:48)(cid:48) radius used in ephemeris-matching method; see Section 3.2.2), and (b) some detections have been removed from the master frame in the clean-up step (see Section 3.2.1). However, a small portion of the known asteroids have much smaller ratio of detection number (i.e., < 50%). We found that these low-ratio asteroids have multiple segments in the Hough Transform, which were not probably connected in the linking steps to become long segments (see the orange color in Figure 6). -- 11 -- 4.1. Asteroid Discovery In total, we found 3574 new asteroids which have five or more detections. Figure 7 shows the distribution of the detection number for these new asteroids, in which the 5 to 10 bin has much more objects. We believe that the high raise in that bin is mainly because some of these short segments were not probably connected in the linkage steps to be identified as the same asteroids, a situation similar the multiple-segment known asteroids mentioned in Section 4. Despite these short-segment asteroids, the high detection number of the new asteroids suggests that they are very unlikely to be false discoveries. Although it is probably difficult to trace them back at this moment due to the orbital uncertainty, we still submitted their detections to the Minor Planet Center in any case of accidental recovery in the near future. The summary table and the entire data set of light curve of these new asteroids can be found in the Table 5 and 6. Figure 8 shows the magnitude distribution of the new asteroids along with the known ones in our observing fields, where the new asteroids are mainly of 21.5 ≤ wp1 ≤ 22.5 mag, ∼ 1 mag fainter than the samples of Chang et al. (2019, ; i.e., roughly 30% smaller in diameter). This suggests that the Hough Transform has pushed the asteroid discovery to the actual limiting magnitude (∼ 22.5 mag) of the observation. With this magnitude range, we are able to reach hundred-meter MBAs (see Figure 9) and have a unique chance to conduct a rotation-period study for them. In general, the Hough Transform is useful to discover asteroids when applied to the high-cadence observation. Moreover, its implementation is relatively easy and it works effectively on a data set containing massive noises. -- 12 -- 5. Rotation Period Analysis for New PS1 Asteroids Out of the 3574 new asteroids, 2853 objects, which have 10 or more detections in wP 1 band in their light curves, were performed a rotation period analysis using a 2nd-order Fourier series fitting (Harris et al. 1989): 2(cid:88) k=1 (cid:20)2πk P (cid:21) (cid:20)2πk P (cid:21) Mi,j = Bk sin (tj − t0) + Ck cos (tj − t0) + Zi, (3) where Mi,j are the reduced magnitudes in wP 1 band measured at the epoch, tj; Bk and Ck are the coefficients in the Fourier series; P is the rotation period; and t0 is an arbitrary epoch. We also introduced a constant value, Zi, to correct the possible offsets in magnitude between the measurements obtained from different nights. In the end, we obtained 122 reliable rotation periods. Their information are summarized in the Table 2 and their folded light curves can be found in Figure 10 and 11. Among the 122 objects, 22 have rotation periods of < 2 hr, of which 13 objects were assigned as SFRs due to their highly convincing folded light curves (see Figure 12) and the other 9 objects were temporarily seen as SFR candidates (see Figure 13). The information of the SFRs and the SFR candidates are listed in Table 3 and 4, respectively. Figure 14 shows plot of the diameters vs. rotation periods for these 122 objects on top of the samples obtained from Chang et al. (2019) and the U ≥ 2 objects in the light-curve database (LCDB; Warner, Harris & Pravec 2009). These SFRs are unlikely to be explained by the rubble-pile structure unless they have unusual high bulk density. Compared to the similar studies, e.g., Chang et al. (2015, 2016, 2019), the chance of finding SRFs is higher here. As shown in the simulation carried out by Chang et al. (2019, ; see Figure 12 therein), the recovery rates are similar for the spin rates of ≥ 3 rev/day within each magnitude interval10. This suggests that more SFRs found here is not because we are in 10The simulation of rotation period recovery in Chang et al. (2019) is used to explain our -- 13 -- favor of detecting short period. Therefore, we believe that small MBAs probably harbor more SFRs. One possible explanation is that these SFRs are monoliths and, moreover, small asteroids are more possible to be monoliths. Consequently, more SFRs could be expected in small MBAs. However, this is difficult to explain why the relatively large SFRs (i.e., diameter of ∼ 1 km) could avoid numerous collisions after their formations and remain monolithic. Another possible explanation is the size-dependent cohesion model in which the cohesion play a significant role for small asteroids, allowing them to rotate faster without breaking apart (Holsapple 2007). This is another way to expect more SFRs in small asteroids. We adopted the equations shown in Holsapple (2007); Rozitis et al. (2014); Polishook et al. (2016); Chang et al. (2017) and assumed a bulk density of ρ = 2 g cm−3 to estimate the cohesion required to survive these SFRs under their super-fast spinning. The calculated cohesion of each SFR are given in Table 3 and 4, in which most cases need a cohesion of tens to thousands of P a, a value consistent with the previously reported SFRs (see Table 4 in Chang et al. 2019) and the regolith on the Moon and Mars (such as the Moon and Mars; Mitchell 1974; Sullivan et al. 2011). However, two SFRs, 510525 and 710033, need a cohesion up to thousands of P a, which is several times larger than the average value. This kind of SFRs, larger than kilometer and rotation period of < 1 hr, can put a critical constraint on the cohesion model. Therefore, it is encouraged to conduct a comprehensive rotation period survey on the asteroids of few kilometer in size. 6. Summary The intra-night detections of asteroids show up as straight lines and, therefore, the Hough Transform can be used to locate these line-up detections to discover new result here because the data sets used in both works are the same. -- 14 -- asteroids. We applied this algorithm to the PS1 high cadence observation, a survey was originally conducted to collect a large sample of asteroid light curves for rotation period measurement during October 26-31, 2016 (Chang et al. 2019). Most of the known asteroids in the observing fields were recovered and, moreover, 3574 new asteroids, mainly of 21.5 < m < 22.5 mag, were found. Using the light curves of the new asteroids, we obtained 122 reliable rotation periods, of which 13 are SFRs. Compared to the previous surveys, the chance of discovering SFR is much higher here (i.e., ∼ 11%). Since the data set used here is the same as Chang et al. (2019), except our samples are mainly of 21.5 ≤ wp1 ≤ 22.5 mag (i.e.,one magnitude fainter than theirs; ∼ 30% smaller in size), our result suggests that sub-kilometer MBAs possibly harbor more SFRs. This work is equally contributed by C.-K. Chang and K.-J. Lo. This work is supported by the National Science Council of Taiwan under the grants MOST 107-2112-M-008-009- MY2. We thank the anonymous referee for his comments and suggestions, which makes a great improvement on this work. -- 15 -- REFERENCES Bottke, W. F., Nolan, M. C., Greenberg, R., & Kolvoord, R. A. 1994, Icarus, 107, 255 Bowell, E., Hapke, B., Domingue, D., et al. 1989, Asteroids II, 524 Chambers, K. C., Magnier, E. A., Metcalfe, N., et al. 2016, arXiv:1612.05560 Chang, C.-K., Ip, W.-H., Lin, H.-W., et al. 2014a, ApJ, 788, 17 Chang, C.-K., Waszczak, A., Lin, H.-W., et al. 2014b, ApJ, 791, LL35 Chang, C.-K., Ip, W.-H., Lin, H.-W., et al. 2015, ApJS, 219, 27 Chang, C.-K., Lin, H.-W., & Ip, W.-H. 2016, ApJ, 816, 71 Chang, C.-K., Lin, H.-W.,
1806.01154
2
1806
2018-06-15T13:17:08
The methane distribution and polar brightening on Uranus based on HST/STIS, Keck/NIRC2, and IRTF/SpeX observations through 2015
[ "astro-ph.EP" ]
HST/STIS observations of Uranus in 2015 show that the depletion of upper tropospheric methane has been relatively stable and that the polar region has been brightening over time as a result of increased aerosol scattering. This interpretation is confirmed by near-IR imaging from HST and from the Keck telescope using NIRC2 adaptive optics imaging. Our analysis of the 2015 spectra, as well as prior spectra from 2012, shows that there is a factor of three decrease in the effective upper tropospheric methane mixing ratio between 30\deg N and 70\deg N. The absolute value of the deep methane mixing ratio, which probably does not vary with latitude, is lower than our previous estimate, and depends significantly on the style of aerosol model that we assume, ranging from a high of 3.5$\pm$0.5% for conservative non-spherical particles with a simple Henyey-Greenstein phase function to a low of 2.7%$\pm$0.3% for conservative spherical particles. Our previous higher estimate of 4$\pm$0.5% was a result of a forced consistency with occultation results of Lindal et al. (1987, JGR 92, 14987-15001). That requirement was abandoned in our new analysis because new work by Orton et al. (2014, Icarus 243, 494-513) and by Lellouch et al. (2015, Astron. & AstroPhys. 579, A121) called into question the occultation results. For the main cloud layer in our models we found that both large and small particle solutions are possible for spherical particle models. At low latitudes the small-particle solution has a mean particle radius near 0.3 $\mu$m, a real refractive index near 1.65, and a total column mass of 0.03 mg/cm$^2$, while the large-particle solution has a particle radius near 1.5 $\mu$m, a real index near 1.24, and a total column mass 30 times larger. The pressure boundaries of the main cloud layer are between about 1.1 and 3 bars, within which H$_2$S is the most plausible condensable.
astro-ph.EP
astro-ph
Journal reference: Icarus (2018) Under review. Preprint typeset using LATEX style emulateapj v. 12/16/11 8 1 0 2 n u J 5 1 . ] P E h p - o r t s a [ 2 v 4 5 1 1 0 . 6 0 8 1 : v i X r a THE METHANE DISTRIBUTION AND POLAR BRIGHTENING ON URANUS BASED ON HST/STIS†, KECK/NIRC2, AND IRTF/SPEX OBSERVATIONS THROUGH 2015 L. A. Sromovsky1, E. Karkoschka2, P. M. Fry1, I. de Pater3, H. B. Hammel4 5 Journal reference: Icarus (2018) Under review. ABSTRACT Space Telescope Imaging Spectrograph (STIS) observations of Uranus in 2015 show that the de- pletion of upper tropospheric methane has been relatively stable and that the polar region has been brightening over time as a result of increased aerosol scattering. This interpretation is confirmed by near-IR imaging from HST and from the Keck telescope using NIRC2 adaptive optics imaging. Our analysis of the 2015 spectra, as well as prior spectra from 2012, shows that there is a factor of three decrease in the effective upper tropospheric methane mixing ratio between 30◦ N and 70◦ N. The absolute value of the deep methane mixing ratio, which probably does not vary with latitude, is lower than our previous estimate, and depends significantly on the style of aerosol model that we assume, ranging from a high of 3.5±0.5% for conservative non-spherical particles with a simple Henyey-Greenstein phase function to a low of 2.7%±0.3% for conservative spherical particles. Our previous higher estimate of 4±0.5% was a result of a forced consistency with occultation results of Lindal et al. (1987, JGR 92, 14987-15001). That requirement was abandoned in our new analysis because new work by Orton et al. (2014, Icarus 243, 494-513) and by Lellouch et al. (2015, Astron. & AstroPhys. 579, A121), called into question the occultation results. For the main cloud layer in our models we found that both large and small particle solutions are possible for spherical particle models. At low latitudes the small-particle solution has a mean particle radius near 0.3 µm, a real refractive index near 1.65, and a total column mass of 0.03 mg/cm2, while the large-particle solution has a particle radius near 1.5 µm, a real index near 1.24, and a total column mass 30 times larger. The pressure boundaries of the main cloud layer are between about 1.1 and 3 bars, within which H2S is the most plausible condensable. Subject headings: Uranus, Uranus Atmosphere; Atmospheres, composition, Atmospheres, structure 1. INTRODUCTION The visible and infrared spectra of Uranus are both dominated by the absorption features of methane, its third most abundant gas. In some spectral regions, how- ever, the effects of collision-induced absorption (CIA) by hydrogen can be seen competing with methane. Those wavelengths provide constraints on the number den- sity of methane with respect to hydrogen, and thus on the volume mixing ratio of methane. From analysis of 1 University of Wisconsin - Madison, Madison WI 53706, USA 2 University of Arizona, Tucson AZ 85721, USA 3 University of California, Berkeley, CA 94720, USA 4 AURA, 1212 New York Ave. NW, Suite 450, Washington, DC 20005, USA 5 Space Science Institute, Boulder, CO 80303, USA † Based in part on observations with the NASA/ESA Hub- ble Space Telescope obtained at the Space Telescope Science Institute, which is operated by the Association of Universities for Research in Astronomy, Incorporated under NASA Contract NAS5-26555. HST/STIS spatially resolved spectra of Uranus obtained in 2002, 5 years before equinox and limited to latitudes south of 30◦ N, Karkoschka and Tomasko (2009), hence- forth referred to as KT2009, discovered that good fits to the latitudinal variation of these spectra required a latitudinal variation in the effective volume mixing ratio of methane. They inferred that the southern polar region was depleted in methane with respect to low latitudes by about a factor of two. This suggested a possible merid- ional circulation in which upwelling methane-rich gas at low latitudes was dried out by condensation then moved to high latitudes, where descending motions brought the methane-depleted gas downward, with a return flow at deeper levels. Because post equinox groundbased observations re- vealed numerous small "convective" features in the north polar region that had not been seen in the south polar region just prior to equinox, Sromovsky et al. (2012b) suggested that the downwelling movement of methane- 2 depleted gas would suppress convection in the south po- lar region, providing a plausible explanation for the lack of discrete cloud features there, and further suggested that the presence of discrete cloud features at high north- ern latitudes might mean that methane is not depleted there. However, using 2012 HST/STIS observations de- signed to test that hypothesis, Sromovsky et al. (2014) showed that the depletion was indeed symmetric, with both polar regions depleted by similar amounts, and from imaging observations taken near equinox using discrete narrow band filters that sampled methane-dominated and hydrogen-dominated wavelengths, they showed that the symmetry was also present at equinox and thus prob- ably a stable feature of the Uranian atmosphere. In spite of the apparent general stability of the lat- itudinal distribution of methane, there were significant post-equinox increases in the brightness of the north po- lar region, as well as some evidence for brightening at low latitudes. New HST/STIS observations were obtained in 2015 to further investigate possible changes in the methane distribution and the nature of the polar bright- ening that was taking place. Other relevant developments occurred since our last analysis of the HST/STIS spectra of Uranus. The first is the inference of new mean temperature and methane profiles for Uranus by Orton et al. (2014a), which are in disagreement with the occultation-based profiles of Lindal et al. (1987) and also with those using a reduced He/H2 volume mixing ratio derived by Sromovsky et al. (2011). The second development is an independent deter- mination of the methane volume mixing ratio in the lower stratosphere and upper troposphere by Lellouch et al. (2015) using Herschel far infrared and sub-mm obser- vations. Since both of these results question the valid- ity of Uranus occultation results in general, and more specifically the validity of using them at all times and all latitudes, we decided to modify our analysis so that models would be constrained by STIS spectral observa- tions alone, and abandoned the requirement that our low- latitude methane profiles also be consistent with occul- tation constraints. We also took a fresh look at how to best model the aerosol structure, and found that a much simpler 2-3 layer structure could produce fits as accurate as the more complex five-layer structure we had used in our previous analysis of the 2012 STIS observations. In the following we first describe the approach to con- straining the methane mixing ratio on Uranus, then dis- cuss the new thermal profile for Uranus and its im- plications. We follow that with a discussion of our STIS, WFC3, and Keck supporting observations, the calibration of the STIS spectra, direct comparisons of STIS spectra, comparisons of imaging observations at key wavelengths, description of our approach to radia- tive transfer modeling, results of modeling cloud struc- ture and the distribution of methane, interpretation of those results, how models can be extended to longer wavelengths to match spectra obtained at NASA's In- frared Telescope Facility (IRTF), comparisons with other models, and a final summary and conclusions. 2. HOW OBSERVATIONS CONSTRAIN THE METHANE MIXING RATIO ON URANUS Constraining the mixing ratio of CH4 on Uranus is based on differences in the spectral absorption of CH4 and H2, illustrated by the penetration depth plot of Fig. 1. There methane absorption can be seen to dominate at most wavelengths, while hydrogen's Collision Induced Absorption (CIA) is relatively more important in nar- row spectral ranges near 825 nm, which is covered by our STIS observations, and also near 1080 nm, which we were able to sample with imaging observations using the NIC- MOS F108N filter and Keck He1 A filter. Model calcula- tions that don't have the correct ratio of methane to hy- drogen lead to a relative reflectivity mismatch near these wavelengths. Karkoschka and Tomasko (2009) used the 825-nm spectral constraint to infer a methane mixing ra- tio of 3.2% at low latitudes, but dropping to 1.4% at high southern latitudes. Sromovsky et al. (2011) analyzed the same data set, but used only temperature and mixing ratio profiles that were consistent with the Lindal et al. (1987) refractivity profiles. They confirmed the depletion but inferred a somewhat higher mixing ratio of 4% at low latitudes and found that better fits were obtained if the high latitude depletion was restricted to the upper tropo- sphere (down to ∼2-4 bars). Subsequently, 2009 ground- based spectral observations at the NASA Infrared Tele- scope Facility (IRTF) using the SpeX instrument, which provided coverage of the key 825-nm spectral region, were used by Tice et al. (2013) to infer that both northern and southern mid latitudes were weakly depleted in methane, relative to the near equatorial region, which was enriched by at least 9%. Their I-band analysis yielded a broad peak centered at 6◦S, which was 32±24% above the min- imum found at 44◦N. These low IRTF-based values for the latitudinal variation might be partly a result of lower spatial resolution combined with worse view angles into higher latitudes than obtained by HST observations. 3. HST/STIS 2015 OBSERVATIONS 3 Level at which 2-way Optical Depth=1.00 (all gas absorption + Rayleigh) CH4 only CIA only N 5 9 0 F A 1 E H N 8 0 1 F A T E B A P T N O C H 10-1 100 101 ) s r a b ( e r u s s e r P 0.4 0.6 0.8 Jan17-112804-2018 1.0 Wavelength (µm) 1.2 1.4 1.6 Fig. 1.- Penetration depth vs. wavelength as limited by different opacity sources assuming the Orton et al. (2014a) temperature profile and a 3.5% deep methane mixing ratio. Where absorption dominates, penetration is about one optical depth, but when Rayleigh scattering dominates, light can penetrate many optical depths. Transmission profiles of key HST/NICMOS (dot-dash) and Keck NIRC2 (solid) narrow-band filters are also plotted. Our 2015 spectral observations of Uranus (Cycle 23 HST program 14113, L. Sromovsky P.I.) used four HST orbits, three of them devoted to STIS spatial mosaics and one devoted to Wide Field Camera 3 (WFC3) sup- port imaging. The STIS observations were taken on 10 October 2015 and the WFC3 observations on 11 October 2015. Observing conditions and exposures are summa- rized in Table 1. 3.1. STIS spatial mosaics. Our STIS observations used the G430L and G750L gratings and the CCD detector, which has ∼0.05 arc- second square pixels covering a nominal 52′′×52′′ square field of view (FOV) and a spectral range from ∼200 to 1030 nm (Hernandez et al. 2012). Using the 52′′×0.1′′ slit, the resolving power varies from 500 to 1000 over each wavelength range due to fixed wavelength dispersion of the gratings. Observations had to be carried out within a few days of Uranus opposition (12 October 2015) when the telescope roll angle could be set to orient the STIS slit parallel to the spin axis of Uranus. One STIS orbit produced a mosaic of half of Uranus using the CCD detector, the G430L grating, and the 52′′×0.1′′ slit. The G430L grating covers 290 to 570 nm with a 0.273 nm/pixel dispersion. The slit was aligned with Uranus' rotational axis, and stepped from the evening limb to the central meridian in 0.152 arcsec- ond increments (because the planet has no high spatial resolution center-to-limb features at these wavelengths we used interpolation to fill in missing columns of the mosaic). Two additional STIS orbits were used to mo- saic the planet with the G750L grating. We intended to use the 52′′×0.1′′ slit (524-1027 nm coverage with 0.492 nm/pixel dispersion) for both orbits, but an error in the program resulted in half of the half-disk covered with the nominal 0.05′′ slit. This produced a higher spectral resolution at the cost of a significant reduction in signal to noise ratio. The limb to central meridian stepping was at 0.0562 arcsecond intervals for the G750L grating. Aside from the slit width error, this was the same proce- dure that was used successfully for HST program 9035 in 2002 (E. Karkoschka, P.I.) and for HST program 12894 in 2012 (L. Sromovsky, P.I.). As Uranus' equatorial radius was 1.85 arcseconds when observations were performed, stepping from one step off the limb to the central merid- ian required 13 positions for the G430L grating and 36 for the G750L grating. Two orbits were needed to com- plete the G750L grating observations, spanning a total time of 2 h 17 m, during which Uranus rotated 47◦. This rotation was not a problem because of the high degree of zonal symmetry of Uranus and because our analysis rejected any small scale deviations from it. Exposure times were similar to those used in the 2002 and 2012 programs, with 70-second exposures for G430L and 84-second exposures for G750L gratings, using the 1 electron/DN gain setting. These exposures yielded single-pixel signal-to-noise ratios of around 10:1 at 300 nm, > 40:1 from around 400 to 700 nm, and decreasing to around 20:1 (methane windows) to < 10:1 (methane absorption bands) at 1000 nm. For the G750L grating, the part of the planet covered by the narrow slit setting (from 0.9′′ from the planet center to the central meridian) the signal to noise at continuum wavelengths was reduced by a factor of ∼1.7 at short wavelengths to ∼2.1 at the longest wavelengths. In methane bands, where readout noise dominates, the degradation factor was ∼4.5. 3.2. Supporting WFC3 imaging. 4 Science exposures from 2015 HST program 14113. TABLE 1 Instrument Filter or Exposure No. of Relative Start Start Orbit 1 1 2 3 19 19 19 19 19 19 19 19 19 19 19 Date (UT) Time (UT) 2015-10-10 2015-10-10 2015-10-10 2015-10-10 2015-10-11 2015-10-11 2015-10-11 2015-10-11 2015-10-11 2015-10-11 2015-10-11 2015-10-11 2015-10-11 2015-10-11 2015-10-11 13:54:06 14:09:48 15:28:02 17:03:25 18:30:59 18:32:44 18:34:24 18:35:48 18:38:17 18:40:24 18:42:11 18:44:05 18:50:43 19:02:25 19:09:25 STIS STIS STIS STIS WFC3 WFC3 WFC3 WFC3 WFC3 WFC3 WFC3 WFC3 WFC3 WFC3 WFC3 Grating MIRVIS G430L G750L G750L F336W F467M F547M F631N F665N F763M F845M F953N FQ889N FQ937N FQ727N (sec) 5.0 70.0 84.0 84.0 30.0 16.0 6.0 65.0 52.0 26.0 35.0 250.0 450.0 150.0 210.0 Phase Exp. Angle (◦) 0.09 0.09 0.09 0.09 0.04 0.04 0.04 0.04 0.04 0.04 0.04 0.04 0.04 0.04 0.04 1 13 19 19 1 1 1 1 1 1 1 1 1 1 1 On October 10 the sub-observer planetographic latitude was 31.7◦ S, the observer range was 18.984 AU (2.8400×109 km), and the equatorial angular diameter of Uranus was 3.7126 arcseconds. The first two STIS orbits used the 52′′×0.1′′ slit and the third inadvertently used the 52′′×0.05′′ slit The complex radiometric calibration of the STIS spec- tra relies on calibrated WFC3 images to provide the final wavelength dependent correction functions. To ensure that this function was determined as well as possible for the Cycle 23 observations in 2015, and to cross check the extensive spatial and spectral corrections that are required for STIS observations, we used one additional orbit of WFC3 imaging at a pixel scale of 0.04 arcseconds with eleven different filters spread over the 300-1000 nm range of the STIS spectra. These WFC3 images are dis- played in Fig. 2, along with synthetic images with the same spectral weighting constructed from STIS spatially resolved spectra, as described in the following section. The filters and exposures are provided in Table 1. 3.3. Supporting near-IR imaging HST/NICMOS and groundbased Keck and Gemini imaging at near-IR wavelengths help to extend and fill gaps in the temporal record of changes occurring in the atmosphere of Uranus. Fig. 3 shows that the differ- ence between polar and low latitude cloud structures has evolved over time. The relatively rapid decline of the bright "polar cap" in the south and its reforma- tion in the north is faster than seems consistent with the long radiative time constants of the Uranian atmo- sphere (Conrath et al. 1990). In following sections we will show that the polar brightness in 2015 (and presum- ably also in 1997) is not due very much to latitudinal variations in aerosol scattering, but is mainly due to a much lower degree of methane absorption at high lati- tudes. This latitudinal variation of methane absorption appears to be stable over time according to infrared ob- servations (Sromovsky et al. 2014). Thus, at times when the polar region was as dark as low latitudes (compared at the same view angles), it appears not that methane ab- sorption was greater then, but instead that aerosol scat- tering was reduced, a causal relationship we will here further confirm regarding polar brightness increases be- tween 2012 and 2015. The aforementioned interpretation of the bright polar region on Uranus can be partly inferred from the charac- teristics of near-IR images at key wavelengths that have different sensitivities to methane and hydrogen absorp- tion, as illustrated in Fig. 4. Images at hydrogen dom- inated wavelengths (panels A and E) reveal relatively bright low latitudes, and high latitudes that were either darker, as at equinox (A), or comparably bright, as in 2015 (E). At methane dominated wavelengths, low lati- tudes are relatively darker, especially in 2015, where the excess methane absorption at low latitudes is obvious from comparing panels E and F. 4. STIS DATA REDUCTION AND CALIBRATION. The STIS pipeline processing used at STScI is just the first step of a rather complex calibration procedure, which is described by KT2009 for the 2002 observations, and by Sromovsky et al. (2014) for the 2012 observa- tions. Essentially the same procedure was followed for the 2015 observations. Additionally, 2002 and 2012 STIS cubes were recalibrated using WFPC2 and WFC3 im- 5 Near-IR observations from HST, Keck, and Gemini observatories. TABLE 2 PID 7429 11118 11118 11118 11190 2010B- Q-110 Telescope /Instrument HST/NICMOS Keck II/NIRC2 Keck II/NIRC2 HST/NICMOS HST/NICMOS HST/NICMOS Keck II/NIRC2 Keck II/NIRC2 HST/NICMOS Gemini-N/NIRI Keck II/NIRC2 Keck II/NIRC2 Keck II/NIRC2 Keck II/NIRC2 Obs. Time Obs. Date 09:50:24 1997-07-28 07:14:51 2003-10-06 2004-07-11 11:30:32 2007-07-28 04:39:xx 2007-07-28 04:22:30 04:39:13 2007-07-28 14:39:28 2007-07-31 14:32:33 2007-07-31 2007-08-16 07:32:32 Phase S.O. Filter Angle CLat PI, Notes F165M H H F095N F095N F108N PaBeta Hcont F165M -40.3 Tomasko, 1 -18.1 Hammel, 2 -11.1 de Pater, 2 0.61 Sromosvky, 3 0.58 Sromovsky, 3 0.58 Sromovsky, 3 0.51 Sromovsky, 2 0.49 Sromovsky, 2 -0.0 Trafton, 3 2.0 1.87 2010-11-02 07:08:57 H G0203 9.3 Sromovsky, 4 2014-08-06 2015-08-29 2015-08-29 2015-08-30 13:42:06 12:09:05 12:04:41 10:56:55 H He1A PaBeta H 28.4 de Pater, 2 31.96 de Pater, 2 31.96 de Pater, 2 31.9 de Pater, 2 NOTES: 1: pscale = 0.0431 as/pixel; 2: pscale = 0.009942 as/pixel ; 3: pscale = 0.0432 as/pixel ; 4: pscale= 0.02138 as/pixel ages newly reduced using the best available detector re- sponsivity functions and filter throughput functions. All three calibrated STIS cubes and related information can be found online in the HST MAST archive as described in the supplemental material section at the end of the paper. In the discussion that follows, we first describe the processing of supporting WFC3 imaging. In the case of 2012 recalibration, WFC3 imaging was also utilized, but for the 2002 recalibration, WFPC2 images were used. We then describe the creation of our calibration correc- tion function, describe our spectral cube construction, and finally our comparison of STIS synthetic images with bandpass filter images. Each WFC3 image was deconvolved with an appro- priate Point-Spread Function (PSF) obtained from the tiny tim code of Krist (1995), optimized to result in data values close to zero in the space view just off the limb of Uranus. To match the spatial resolution of the STIS images, the WFC3 images were then reconvolved with an approximation of the PSF given in the analysis supplemental file of Sromovsky et al. (2014). The im- ages were then converted to I/F using the best available header PHOTFLAM values [given in WFC3 ISR 2016- 001] and the Colina et al. (1996) solar flux spectrum, av- eraged over the WFC3 filter band passes. (PHOTFLAM is a multiplier used to convert instrument counts of elec- trons/second to flux units of ergs/s/cm2/A.) To obtain a disk-averaged I/F, the planet's light was integrated out to 1.15 times the equatorial radius, then averaged over the planet's cross section in pixels, which was computed using NAIF ephemerides (Acton 1996) and SPICELIB limb ellipse model (SPICELIB is NAIF toolkit software used in generating navigation and ancillary instrument information files.) The disk-averaged I/F (using the ini- tial calibration) was also computed for each of the STIS monochromatic images, and the filter- and solar flux- weighted I/F was computed for each of the WFC3 filter pass bands that we used. By comparing the synthetic disk-averaged STIS I/F, using the initial calibration, to the corresponding WFC3 values, we constructed a correction function to improve the radiometric calibration of the STIS cube. Figure 5C shows the ratios of STIS to WFC3 disk-integrated bright- ness, and the quadratic function that we fit to these ratios as a function of wavelength, for the 2015 data set and re- calibrations of the previous two data sets. We heavily weighted the broadband filters, and computed an effec- tive wavelength weighted according to the product of the solar spectrum and the I/F spectrum of Uranus. The RMS deviation of individual filters from the calibration curve given in Fig. 5 is about 1% RMS for 2012 and 2015 correction curves, but about 2 % RMS for the 2002 cal- ibration curve (fit points not shown). For narrow filters such as FQ937N, typical deviations are some three times larger. The difference between the 2002 and the later calibration curves is mostly due to use of different slit lo- cations, which result in different light paths through the monochrometer. The 2012 and 2015 curves use the same 6 A G M R WFC3 15 STIS 15 STIS/WFC3 WFC3 15 STIS 15 STIS/WFC3 B C D E F F336W F467M F547M F631N F665N FQ727N H F336W F467M I F547M J K L F631N F665N FQ727N F336W F467M F547M F631N F665N FQ727N N O P Q F763M F845M FQ889N FQ937N F953N S T U V F763M F845M FQ889N FQ937N F953N F763M F845M FQ889N FQ937N F953N Fig. 2.- WFC3 images of Uranus taken on 11 October 2015 (A-F and M-Q) compared to synthetic band-pass filter images (G-L and R-V) created from weighted averages of STIS spectral data cubes using WFC3 throughput and solar spectral weighting. The north pole is at at the right. Portions of the synthetic images east of the central meridian are obtained by reflection of the images west of the central meridian. The ratio images are stretched to make 0.8 black and 1.2 white. A B C D E F NICMOS, F165M 1997-07-28, 09:50 Keck II, H 2004-07-11, 11:30 NICMOS, F165M 2007-08-16, 07:32 Gemini-North, H 2010-11-02, 07:08 Keck II, H 2014-08-06, 13:42 Keck II, H 2015-08-30, 10:56 Fig. 3.- H-band (1.6-µm) images of Uranus from 1997 through 2015, from observatories/instruments given in the legends. The bright south polar region seen in 1997 (A), 10 years before equinox, is similar to the bright north polar region seen in 2015 (F), eight years after equinox. Images taken during the 2007 equinox year (C) found that neither polar region was bright. The longitude and planetographic latitude grid lines are at 30◦ and 15◦ intervals respectively. A B C D E F 7 NICMOS, F108N 2007-07-28, 04:39 Keck II, PaBeta (1.29) 2007-07-31, 14:39 Keck II, Hcont (1.58) 2007-07-31, 14:32 NICMOS, F095N 2007-07-28, 04:22 Keck II, He1_A (1.09) 2015-08-29, 12:09 Keck II, PaBeta (1.29) 2015-08-29, 12:04 Fig. 4.- Near-IR narrow-band images of Uranus obtained near the 2007 equinox (A-D) and in August 2015 (E, F) from observato- ries/instruments given in the legends. The NICMOS F108N image (A) and the Keck II He1 A image (E) sample wavelengths at which hydrogen absorption dominates. The remaining images sample wavelengths of comparable absorption but due entirely to methane. Note that low latitudes are relatively darker than high latitudes at methane dominated wavelengths (e.g. in B and F), which is not the case for wavelengths dominated by hydrogen absorption (as in A and E). Grid lines are the same as in Fig. 3. slit location and thus should be the same, and indeed they are consistent to about 1%. If one assumes that the spectrum of Uranus varies only slowly with time, one can add many other filters to plots of Fig. 5 where images are available somewhat close to the time of STIS spectroscopy. The medium and wide filters plot quite consistently near the same curve while many narrow filters show significant offsets, suggesting that an improved calibration weights the narrow filters much less in the fitted curve. This consideration changes the calibration by about 1% and thus does not make a big difference with respect to our previous adopted calibra- tion, but our new calibration is more reliable because it is less dependent on unreliable data from narrow filters. The final calibrated cubes contain 150 pixels parallel to the spin axis of Uranus and 75 pixels perpendicular to its spin axis, with a spatial sampling interval of of 0.015×RU km/pixel (384 km/pixel), which is equivalent to 0.028 arc seconds per pixel for 2015 observations. (Here RU is the equatorial radius of Uranus). The center of Uranus is located at coordinates (74, 74), where (0, 0) is the lower left corner pixel. The spatial resolution of the final cube is defined by a point-spread function with a FWHM of 3 pixels. The cube contains an image for each of 1800 wavelengths sampled at a spacing of 0.4 nm, with a uni- form spectral resolution of 1 nm. Navigation backplanes are provided, in which the center of each pixel is given a planetographic latitude and longitude, solar and observer zenith angle cosines, and an azimuth angle. As a sanity check on the STIS processing we compared WFC3 images to synthetic WFC3 images created from our calibrated STIS data cubes, as shown in Fig. 2. Ra- tio plots of STIS/WFC3 show the desired flat behavior, except very close to the limbs, where STIS I/F values ex- ceed WFC3 values. The most significant discrepancy is in the overall I/F level computed for the FQ937N filter (note the dark ratio plot in the bottom row), a conse- quence of our calibration curve being 10% high for that filter. 5. CENTER-TO-LIMB FITTING The low frequency of prominent discrete cloud fea- tures on Uranus and its zonal uniformity make it possi- ble to characterize the smooth center-to-limb profiles of the background cloud structure without much concern about longitudinal variability, even though we observed only half the disk of Uranus. These profiles provide im- portant constraints on the vertical distribution of cloud particles and the vertical variation of methane compared to hydrogen. Because our observations were taken very close to zero phase, these profiles are a function of just one angular parameter, which we take to be µ, the co- sine of the zenith angle (the observer and solar zenith angles are essentially equal). They also have a relatively simple structure that we characterized using the same 3-parameter function KT2009 used to analyze the 2002 STIS observations, and which we also used to fit the 2012 observations. For each 1◦ of latitude from 30◦ S to 87◦ N, all image samples within 1◦ of the selected latitude and with µ > 0.175 are collected and fit to the empirical function I(µ) = a + bµ + c/µ, (1) assuming all samples were collected at the desired lati- tude and using the µ value for the center of each pixel of the image samples. Fitting this function to center- to-limb (CTL) variations at high latitudinal resolution makes it possible to separate latitudinal variations from those associated with view angle variations. Before fitting the CTL profile for each wavelength, the spectral data were smoothed to a resolution of 2.88-nm to improve signal to noise ratios without significantly blur- ring key spectral features. (Our prior analysis was con- ducted in the wavenumber domain and used smoothing to a resolution of 36 cm−1.) Sample fits are provided in Fig. 6. Most of the scatter about the fitted profiles is due to noise, which is often amplified by the deconvo- lution process. Because the range of observed µ values 8 l o d e b A c i r t e m o e G l o d e b A 5 1 0 2 / o d e b A l ) 3 C F W 2 C P F W ( / I S T S . m i l e r P ) 3 C F W 2 C P F W ( / I S T S . l a C A B C D 0.6 0.4 0.2 0.0 1.2 1.1 1.0 0.9 0.8 0.7 0.9 0.8 0.7 0.6 0.5 1.05 1.00 0.95 0.90 2002 2012 2015 2002 New 2012 New 2015 N 3 5 9 F N 7 3 9 Q F N 9 8 8 Q F STIS Photometric Correction Functions W 6 3 3 F M 7 6 4 F M 7 4 5 F N 5 6 6 F N 1 3 6 F M 3 6 7 F N 7 2 7 Q F M 5 4 8 F 400 600 Wavelength (nm) 800 1000 Fig. 5.- A: Radiometrically calibrated disk-averaged I/F spectra for three year of STIS observations. STIS observations from 2002 and 2012 have been recalibrated from previous incarnations (Karkoschka and Tomasko 2009; Sromovsky et al. 2014). B: Ratio of each year's disk-averaged I/F to 2015. C: STIS photometric calibration functions (raw stis albedo divided by WFC3 albedo). The functions are a fit to ratios constructed using synthetic band-pass filter disk-integrated I/F values (preliminary-calibration) divided by corresponding I/F values obtained from WFC3 measurements (circles and horizontal bars indicate filter effective wavelength and full-width half maximum transmission for 2012 and 2015). D: Ratio of final calibrated STIS synthetic disk-averaged I/F values to WFC3 reflectivities, showing scatter of ratios relative to the fits, for 2012 and 2015. decreases away from the equator at high southern and northern latitudes, we chose a moderate value of µ = 0.7 as the maximum view-angle cosine to provide a rea- sonably large unextrapolated range of 16◦ S to 77◦ N. Ranges for other years and for a µ range of 0.3 to 0.6 are given in Table 3. Unless otherwise noted all our results are derived without extrapolation. The CTL fits can also be used to create zonally smoothed images by replacing the observed I/F for each pixel by the fitted value. Results of that procedure are displayed in a later section. Latitude ranges for two different view-angle cosine TABLE 3 ranges. Year 0.3 ≤ µ ≤ 0.6 0.3 ≤ µ ≤ 0.7 2002 74◦S – 33◦N 67◦S – 26◦N 2012 35◦S – 72◦N 28◦S – 65◦N 2015 23◦S – 77◦N 16◦S – 77◦N 6. DIRECT COMPARISONS OF STIS SPECTRA CTL curves at glat 0.00o 552.0 nm, X 1.0 620.0 nm, X 1.0 826.8 nm, X 2.0 890.0 nm, X 4.0 935.2 nm, X 2.0 wlen res = 2.88 nm limbcorr 0.8 0.6 F / I 0.4 0.2 0.0 limbcorr 0.8 0.6 F / I 0.4 0.2 0.0 CTL curves at glat 60.00o 552.0 nm, X 1.0 620.0 nm, X 1.0 826.8 nm, X 2.0 890.0 nm, X 4.0 935.2 nm, X 2.0 wlen res = 2.88 nm 0.2 PS file: minn_ctl_diag.ps 0.4 0.6 cos(Emission Angle) 0.8 1.0 0.2 Aug11-100905-2016 PS file: minn_ctl_diag.ps 0.4 0.6 cos(Emission Angle) 9 0.8 1.0 Aug11-101023-2016 Fig. 6.- Sample center-to-limb fits for 0◦ N (left) and 60◦N (right), as described in the main text. STIS I/F samples and fit lines with uncertainty bands are shown for five different wavelengths indicated in the legends. The latitude bands sampled for these fits are darkened in the inset images of the half-disk of Uranus. A rough assessment of the changes between 2012 and 2015 and the differences between high and low latitudes in these two years can be made with the help of direct comparisons of STIS spectra, as in Fig. 7. Note that at 10◦ N there is almost no difference between 2012 and 2015 spectra (panels A and B). This is also the case for µ values of 0.3 and 0.5, which are not shown in the figure. For 2015, (see panel E) the lack of any I/F difference be- tween 10◦ N and 60◦ N at 0.83 µm, which is a wavelength at which hydrogen absorption dominates, suggests that there is not much difference in aerosol scattering between these two latitudes. A similar lack of difference at 0.93 µm, a wavelength of weak (but dominant) methane ab- sorption, suggests that at very deep levels, there may not be much of a latitudinal difference in methane mixing ra- tios, or that there is an aerosol layer blocking visibility down to levels that might sense such a difference. Yet the fact that wavelengths of intermediate methane absorp- tion do show a significant increase in I/F with latitude suggests that at upper tropospheric levels the methane mixing ratio does decline with latitude, which is a known result from previous work, and is refined by the analysis presented in following sections. Somewhat different re- sults are seen, for 2012 (in panel G). The 10◦ N and 60◦ N I/F values at 0.83 µm and 0.93 µm do differ (panels G and H), which we will later show is a result of differ- ences in aerosol scattering. The small size of continuum differences between 2012 and 2015 (panels D and H) is partly a result of the relatively smaller impact of par- ticulates at short wavelengths where Rayleigh scattering is more significant. At absorbing wavelengths for which gas absorption is important, the optical depth and verti- cal distribution of particulates have a greater fractional effect on I/F and thus small secular changes in these pa- rameters can be more easily noticed. 7. DIRECT COMPARISON OF METHANE AND HYDROGEN ABSORPTIONS VS. LATITUDE. If methane and hydrogen absorptions had the same de- pendence on pressure, then it would be simple to estimate the latitudinal variation in their relative abundances by looking at the relative variation in I/F values with lati- tude for wavelengths that produce similar absorption at some reference latitude. Although this idea is compro- mised by different vertical variations in absorption, which means that latitudinal variation in the vertical distribu- tion of aerosols can also play a role, it is nevertheless useful in a semi-quantitative sense. Thus we explore sev- eral cases below. 7.1. Image comparisons at key near-IR wavelengths in 2007 and 2015 Our first example compares an HST/NICMOS image made with an F108N filter (centered at 1080 nm), which is dominated by H2 CIA, to a KeckII/NIRC2 image made with a PaBeta filter (centered at 1290 nm), which is dom- inated by methane absorption. The images are shown in panels A and B in Fig. 4 and latitude scans at fixed view angles are shown in Fig. 8. The NICMOS obser- vation was taken on 28 July 2007 at 4:39 UT and the Keck observation on 31 July 2007 at 14:39 UT (see Ta- ble 2 for more information). That these two observations sense roughly the same level in the atmosphere is demon- strated by the penetration depth plot in Fig. 1, which also displays the filter transmission functions. The ab- solute (unscaled) I/F profiles for these two images near the 2007 Uranus equinox are displayed for µ = 0.6 and µ = 0.8 by thinner lines in Fig. 8. At high latitudes in both hemispheres, profiles at the two wavelengths agree closely, and both increase towards the equator. But as low latitudes are approached the two profiles diverge dramatically, with the I/F for the hydrogen-dominated 10 A 0.6 2015 I/F at 10oN, µ = 0.7 2012 I/F at 10oN, µ = 0.7 E 0.6 2015 I/F at 10oN, µ = 0.7 2015 I/F at 60oN, µ = 0.7 0.4 F / I 0.2 0.0 0.06 0.04 0.02 0.00 -0.02 -0.04 -0.06 0.6 e c n e r e f f i D 0.4 F / I 0.2 0.0 0.06 0.04 0.02 0.00 -0.02 -0.04 -0.06 e c n e r e f f i D B C 0.4 0.6 0.8 1.0 2015 I/F at 60oN, µ = 0.7 2012 I/F at 60oN, µ = 0.7 D 0.4 0.6 Wavelength (µm) 0.8 1.0 0.4 F / I 0.2 0.0 0.06 0.04 0.02 0.00 -0.02 -0.04 -0.06 0.6 e c n e r e f f i D 0.4 F / I 0.2 0.0 0.06 0.04 0.02 0.00 -0.02 -0.04 -0.06 e c n e r e f f i D F G 0.4 0.6 0.8 1.0 2012 I/F at 10oN, µ = 0.7 2012 I/F at 60oN, µ = 0.7 H 0.4 0.6 Wavelength (µm) 0.8 1.0 Fig. 7.- Comparison of 2012 and 2015 STIS spectra at 10◦ N (A) and 60◦ N (C), and comparison of STIS 10◦ N and 60◦ N spectra from 2015 (E) and 2012 (G), with difference plots shown in panels B, D, F, and H respectively. The dotted curve in panel H is a copy of the 2015 difference curve from panel F. Latitudes are planetographic. Note the nearly exact equality (in A) of 10◦ N spectra from 2012 and 2015. wavelength ending up 50% greater than for the methane- dominated wavelength, indicating much greater methane absorption at low latitudes than at high latitudes. As noted by Sromovsky et al. (2014) this suggests that up- per tropospheric methane depletion (relative to low lat- itudes) was present at both northern and southern high latitudes in 2007, at least roughly similar to the pattern that was inferred by Tice et al. (2013) from analysis of 2009 IRTF SpeX observations. Latitudinal variations in aerosol scattering could distort these results somewhat, but because they affect both wavelengths to similar de- grees, most of the effect is likely due to methane varia- tions. He1 A filter). In 2015 the I/F in the He1 A filter is relatively independent of latitude as shown by the im- age in panel E of Fig. 4, indicating that aerosol scatter- ing must have a relatively weak latitudinal dependence. Note that latitude scans at fixed view angles for these fil- ters (shown in Fig. 8) exhibit a low-latitude divergence of the hydrogen-dominated and methane-dominated wave- lengths which has about the same magnitude in 2015 as seen for the 2007 observations, indicating a similar in- crease of methane absorption at low latitudes. A second example is shown by the thicker lines in Fig. 8, which display latitudinal scans of 2015 images shown in panels E and F of Fig. 4. These were made by the KeckII/NIRC2 camera with He 1A and PaBeta filters on 29 August 2015 (see Table 2). The He 1A filter is similar to the NICMOS F108N filter, as shown in Fig. 1. The 2015 observations present a picture that is somewhat different from the 2007 observations, with high north- ern latitudes much brighter (at the same view angles) than in 2007. This change appears to be entirely due to increased aerosol scattering. This conclusion is sup- ported by the characteristics of images obtained with H2-dominated filters (NICMOS F108N filter and Keck 7.2. Direct comparison of key STIS wavelength scans A comparison of the STIS latitude scans at methane dominated wavelengths with scans at H2 CIA domi- nated wavelengths is also informative. By selecting wave- lengths that at one latitude provide similar I/F values but very different contributions by H2 CIA and methane, one can then make comparisons at other latitudes to see how I/F values at the two wavelengths vary with lati- tude. If aerosols did not vary at all with latitude, then any observed I/F variation would be a clear indicator of variation in the ratio of CH4 to H2. Fig. 9 displays a detailed view of I/F in the spectral region where hy- drogen CIA exceeds methane absorption (see Fig. 1 for penetration depths). Near 827 nm (A) and 930 nm (C) the I/F values are similar but the former is dominated e c n a t c e l f e R e v i t a e R l 0.12 0.10 0.08 0.06 0.04 0.02 0.00 A µ = 0.60 2015 2007 Keck/NIRC2 PaBeta (1290 nm, all CH4) Keck/NIRC2 He1_A (1080 nm, mostly H2) Keck/NIRC2 PaBeta (1290 nm, all CH4) NIC F108N (1080 nm, mostly H2) 0.12 0.10 0.08 0.06 0.04 0.02 0.00 2015 2007 60 80 Solid lines: mostly H2 absorption Dashed lines: all CH4 absorption 11 B µ = 0.80 -60 -40 -20 Planetographic Latitude (o) 20 40 0 60 80 -60 -40 -20 Planetographic Latitude (o) 20 40 0 Fig. 8.- Latitudinal profiles at fixed zenith angle cosines of 0.6 (A) and 0.8 (B) for F108N (HST/NICMOS) and PaBeta (Keck/NIRC2) filters (light solid and dashed lines respectively) taken near the Uranus equinox in 2007, and Keck/NIRC2 filter He1 A and PaBeta filters (thick solid and dashed lines respectively) in 2015. In 2007 the southern hemisphere was still generally brighter than the northern hemisphere and the 38◦ S - 58◦ S southern bright band was still better defined and considerably brighter than the corresponding northern bright band. The relatively low equatorial I/F values for the methane-dominated PaBeta filter (1290 nm) implies greater CH4/H2 absorption at low latitudes. We scaled the Keck 2015 observations to approximately match the 2007 observations at 10◦N, where we found almost no change between 2012 and 2015 at CCD wavelengths (Fig. 7A). Note that between 2007 and 2015, the north polar region has brightened by comparable amounts at both hydrogen-dominated and methane-dominated wavelengths, indicating that it the brightening is due to increased aerosol scattering, not a temporal change in the methane mixing ratio at high latitudes. More information about the observations is given in Table 2. by hydrogen absorption (dot-dash curve) and the latter by methane absorption (dashed curve). Near 835 nm (B) there is a relative minimum in hydrogen absorption, while methane absorption is still strong. For the lati- tude and view angle of this figure (50◦ N and µ = 0.6), I/F values are nearly the same at all three wavelengths, suggesting that they all produce roughly the same atten- uation of the vertically distributed aerosol scattering. Figure 10 displays the latitudinal scans for the three wavelengths highlighted in Fig. 9 for the STIS obser- vations in 2002 (shown by thin lines), 2012 (thick gray lines), and 2015 (thick black lines). This is for a view angle cosine of µ = 0.6, chosen as a compromise be- tween amplitude of variation and coverage in latitude. The 2012 I/F for the hydrogen-dominated wavelength increases towards low latitudes, while the I/F for the methane-dominated wavelength decreases substantially, indicating an increase in the amount of methane relative to hydrogen at low latitudes. Similar effects are seen in 2002 (providing the best view of southern latitudes) and 2015 (providing the best view of the northern latitudes). The hydrogen-dominated wavelengths have relatively flat latitudinal profiles of I/F in the southern hemisphere in 2002 and in the northern hemisphere in 2015, while the methane-dominated wavelengths show strong decreases towards the equator, beginning at about 45◦ S and 50◦ N. For µ=0.8 (not shown), which probes more deeply, the latitudinal variation for the methane dominated wave- lengths is somewhat greater (a 30% decrease in I/F at low latitudes vs. a 20% decrease for µ = 0.6). The spectral comparisons in Fig. 10 also reveal sub- stantial secular changes between 2002 and 2012 and be- tween 2012 and 2015. At wavelengths for which methane and/or hydrogen absorption are important, the north- ern low-latitudes have brightened substantially, while the southern low latitudes have darkened. The bright band between 38◦ and 58◦ N continued to brighten. Its bright- ening and the darkening of the corresponding southern band was already apparent from a comparison of 2004 and 2007 imaging (Sromovsky et al. 2009). The most dramatic change between 2012 and 2015 is the increased brightness of the polar region. The nearly identical brightening at all wavelengths, shown by the ratio plot in panel B of Fig. 10, argues that the brightening is due to aerosol scattering rather than a decrease in the amount of methane. We will confirm this with radiation transfer modeling in Section 9. A color composite of the highlighted wavelengths (us- ing R = 930 nm, G = 834.6 nm, and B = 826.8 nm) is shown in Fig. 11, where the three components are bal- 12 0.20 0.15 F / I 0.10 0.05 0.00 800 Latitude= 50o, µ = 0.60 solid = I/F spectrum dashed = CH4 abs. coeff., (km-am)-1 dot-dash = scaled H2 CIA, (km-am2)-1 820 A B 840 860 880 Wavelength (nm) 900 920 C 940 Fig. 9.- I/F and absorption spectra comparing the equilibrium H2 CIA coefficient spectrum (divided by 1.2×10−7, shown as dot-dash curve) and methane absorption coefficient spectrum (dashed). Note that the I/F spectrum has nearly equal I/F values at 826.8 nm (A), 834.6 nm (B), and 930 nm (C), but H2 absorption is much greater at A than at B, while the opposite is true of methane absorption, and at C only methane absorption is present. In a reflecting layer model, changes in cloud reflectivity should affect wavelengths A-C by the same factor, but changes in methane mixing ratio would affect C most and A least. From Sromovsky et al. (2014). A 0.20 826.8 nm (mostly H2, solid) 834.6 nm (mostly CH4, dot-dash) 930.0 nm (all CH4, dotted) 0.15 F / I 2002 0.10 µ = 0.60 2015 2015 2012 2012 2 1 0 2 o t o i t a r F / I B 2002 2012 2015 1.6 1.4 1.2 1.0 0.8 0.6 0.4 2015 2002 -80 -70 -60 -50 -40 -30 -20 -10 0 10 20 30 40 50 60 70 80 -80 -70 -60 -50 -40 -30 -20 -10 0 10 20 30 40 50 60 70 80 Planetographic Latitude (o) Planetographic Latitude (o) Fig. 10.- I/F vs. latitude at µ = 0.6 (A) and ratio of I/F scans to the 2012 scans (B) for three wavelengths with different amounts of methane and hydrogen absorption. Thin curves are from 2002, thick gray curves from 2012, and thick black curves from 2015. These are plots of center-to-limb fitted values instead of raw image data. In all cases the methane-dominated wavelengths have much reduced I/F at low latitudes, compared to hydrogen-dominated wavelengths, indicating higher CH4 /H2 ratios at low latitudes. Also note the increased north polar brightness at all displayed wavelengths between 2012 and 2015, indicating that the temporal change is due to increased aerosol scattering. anced to produce comparable dynamic ranges for each wavelength. This results in nearly blue low latitudes where absorption at the two methane dominated wave- lengths is relatively high and green/orange polar regions as a result of the decreased absorption by methane there. The spatial structure in high-resolution near-IR H-band Keck II images is also shown in each panel, revealing that small discrete cloud features remain visible in the north polar regions even in the 2015 images, taken after the polar region brightened significantly between 2012 and 2015. Also noteworthy, is the lack of such features in the south polar region (panel A). This asymmetry is some- what surprising. As noted by Sromovsky et al. (2014), because both polar regions are depleted in methane, the suggested downwelling motions that could produce such depletion would be expected to inhibit convection in both polar regions. 8. RADIATIVE TRANSFER MODELING OF METHANE AND AEROSOL DISTRIBUTIONS 8.1. Radiation transfer calculations In contrast to our prior analysis (Sromovsky et al. 2014), which was carried out in the wavenumber do- main to accommodate our Raman scattering code, here we worked in the wavelength domain, which is better suited to the uniform wavelength resolution of our cali- brated STIS data cubes. We also used an approximation for the effects of Raman scattering rather than carrying out the full Raman scattering calculations. We again used the accurate polarization correction described by Sromovsky (2005b) instead of carrying out the time con- suming rigorous polarization calculations. To assess the adequacy of our approximations, we did sample calcu- lations that included Raman scattering and polarization effects on outgoing intensity using the radiation transfer code described by Sromovsky (2005a). Examples in Fig. 12 show that at most wavelengths the errors from these approximations do not exceed a few percent. We improved our characterization of methane ab- sorption at CCD wavelengths by using correlated-k model fits by Irwin et al. (2010), which are available at http://users.ox.ac.uk/∼atmp0035/ktables/ in files ch4 karkoschka IR.par.gz and ch4 karkoschka vis.par.gz. These fits are based on band model results of Karkoschka and Tomasko (2010). To model collision- induced opacity of H2-H2 and He-H2 interactions, we in- terpolated tables of absorption coefficients as a function of pressure and temperature that were computed with a program provided by Alexandra Borysow (Borysow et al. 13 2000), and available at the Atmospheres Node of NASA'S Planetary Data System. We assumed equilibrium hydro- gen, following KT2009 and Sromovsky et al. (2011). After trial calculations to determine the effect of differ- ent quadrature schemes on the computed spectra, we se- lected 12 zenith angle quadrature points per hemisphere and 12 azimuth angles. Test calculations with 10 and 14 quadrature points in each variable changed fit param- eters by only about 1%, which is much less than their uncertainties. 8.2. Thermal profiles for Uranus Assuming the helium volume mixing ratio (VMR) of 0.152 inferred by Conrath et al. (1987), Lindal et al. (1987) used radio occultation measurements of refrac- tivity versus altitude to infer a family of thermal and methane profiles, with each profile distinguished by the assumed methane relative humidity above the cloud level, and the resultant deep volume mixing ratio (VMR) of methane below the cloud level. The cloud was posi- tioned at the point where the refractivity profile had a sharp change in slope. None of these profiles achieved methane saturation inside the cloud layer, even the pro- file with the highest physically realistic humidity level (limited by the requirement that lapse rates could not be superadiabatic). This high-humidity profile also had the highest deep temperatures and the largest deep methane VMR of 4%. By allowing the He/H2 ratio to take on values near the low end of the uncertainty range given by Conrath et al. (1987), Sromovsky et al. (2011) were able to find solutions that achieved methane saturation inside the cloud layer, as well as deep methane mixing ratios somewhat greater than 4%. The above results are based on the Voyager ingress profile, which sampled latitudes from 2◦S to 6◦S. As to whether this local sample can be taken to roughly rep- resent a global mean profile, some guidance is provided by the results that Hanel et al. (1986) derived from the Voyager 2 Infrared Interferometer Spectrometer (IRIS) observations. Inversion of spectral samples near both poles and near the equator yielded temperature profiles that differed by less than 1 K from about 150 mbar to 600 mbar, and the equator and south pole profiles re- mained within 2 K from 60 mbar to 150 mbar, with the north polar profile deviating up to 4 K above the tropopause. More significant variations can be seen at middle latitudes, however, especially in the 60 – 200 mbar range where average temperatures are 3.5 K higher than the latitudinal average near the equator and 4.5 K 14 A B C 2003 / 2002 2012 2015 Fig. 11.- Color composites of fitted center-to-limb smoothed images for 2002 (A, right), 2012 (B, right), and 2015 (C, right), using color assignments R=930 nm (all methane) G=834.6 nm (methane and hydrogen), and B=826.8 nm (mostly hydrogen). The blue tint at low latitudes in all years is due to locally increased methane absorption. We also show near-IR NIRC2 H-band images for 2003 (A, left), 2012 (B, left), and 2015 (C, left). The NIRC2 images are rotation removed averages following Fry et al. (2012) and processed to enhance the contrast of small spatial scales, by adding k times the difference between the original image and a 0.13-arcsecond smoothed version, where k was taken to be 30 for 2003 and 2012 images, but only 22.5 for the 2015 image because of better seeing conditions during its acquisition. 1.00 F / I 0.10 A µ= 0.8 µ= 0.6 µ= 0.4 1.00 F / I 0.10 C µ= 0.8 µ= 0.6 µ= 0.4 line = Raman scattering and Polarization ignored + = true Raman + true Polarization line = modified Wallace + Sromovsky pol. correction + = true Raman + true Polarization l n a m a r + o p e u r t / m a r , l o p e r o n g i 0.5 0.6 0.7 0.8 0.9 1.0 0.01 0.4 1.20 1.15 1.10 1.05 1.00 0.95 0.90 0.85 B 0.4 0.5 0.6 0.7 Wavelength (µm) 0.8 0.9 1.0 l n a m a r + o p e u r t / n a m a r + o p x o r p p a l 0.5 0.6 0.7 0.8 0.9 1.0 0.01 0.4 1.20 1.15 1.10 1.05 1.00 0.95 0.90 0.85 D 0.4 0.5 0.6 0.7 Wavelength (µm) 0.8 0.9 1.0 Fig. 12.- Trial calculations showing errors produced by ignoring Raman scattering and polarization (A and B) and greatly reduced errors achieved by employing the modified Wallace approximation of Raman scattering (Sromovsky 2005a) and the approximation of polarization effects following Sromovsky (2005b). lower near 30◦S (Conrath et al. 1991). In the 200 – 1000 mbar range latitudinal excursions are within 1–1.5 K. Thus it appears that in the most important region of the atmosphere for our applications, the thermal structure was not strongly variable with latitude, at least in 1986. Models of seasonal temperature variations on Uranus by Friedson and Ingersoll (1987) suggest that the effective temperature variation at low latitudes will be extremely small, only 0.2 K peak-to-peak at the equator, increasing to a still relatively small 2.5 K at the poles. Thus, it is plausible to analyze observations during the 2012 – 2015 period with thermal profiles obtained as far back as 1986, even though they are local, but probably more appropri- ate to use thermal structures derived from observations in 2007, averaging over a wide range of latitudes, such as those inferred by Orton et al. (2014a) from nearly disk- integrated spectral observations. Sample thermal and methane profiles are displayed in Fig. 13. The profile of Orton et al. (2014a), hereafter re- ferred to as O14, is based on nearly disk-integrated spec- tral observations obtained with the Spitzer Space Tele- scope near the Uranus equinox in 2007. Among the oc- cultation profile sets, it is only those with high methane VMR values that provide decent agreement with the O14 deep temperature structure, but none of the occultation profiles are compatible with the O14 profile in the 0.30 – 1.0 bar range. One might argue that if radio occultation results agree with O14 at 100 mb and at pressures beyond 1 bar, then the disagreement in temperatures at interme- diate pressure levels is more likely due to an error in the Orton et al. profile because that profile is inferred from different spectrometers in different spectral regions that sample different altitudes, which might suffer from dif- ferences in calibrations, while the radio occultation uses the same measurement (the frequency of a radio signal) throughout the pressure range. It seems more likely that the errors in the radio profile would be in the altitude scale or in offsets due to uncertain He/H2 ratios, rather than varying in the way the differences between the ra- dio and Orton et al. profile do. A similar argument might be made in favor of the Conrath et al. (1987) pro- file over the O14 profile because the former is based on interferometric measurements using the same instrument over the entire spectral range. And the former profile is in good agreement with the occultation-based profiles in the 300–600 mbar range, where the latter is not. On the other hand, the Orton et al. profile allows higher CH4 mixing ratios without saturation in the 0.3 – 1 bar re- gion (Fig. 13) and are thus more compatible with the 15 recent Lellouch et al. (2015) CH4 VMR profile derived from Herschel far-IR and sub mm observations. −0.1) × 10−5 according to Fig. The methane VMR in the stratosphere was estimated to be no greater than 10−5 by Orton et al. (1987). A best fit estimate for the tropopause value of the methane VMR, based on more recent Spitzer observations, is (1.6+0.2 4 of Orton et al. (2014b), which is the value we assumed here in deriv- ing the new F0 profile. However, the even more recent Lellouch et al. (2015) result is three times larger. In terms of relative humidity (ratio of vapor pressure to sat- uration vapor pressure) these stratospheric mixing ratios correspond to humidities of 25% and 75% at the Orton et al. tropopause temperature of 52.4 K. The F profile of Lindal et al. (1987) was derived assuming a constant methane relative humidity of 53% above the cloud tops and a constant stratospheric mixing ratio equal to the tropopause value. In deriving the F0 profile we used linear-in-altitude interpolation of the methane humidity values between the cloud top and tropopause. The F1 profile of Sromovsky et al. (2011) followed the same pro- cedure except that the value of the tropopause mixing ratio was taken to be the earlier upper bound of 10−5 and the He VMR was taken to be 0.1155 instead of 0.152. The lower He VMR was chosen to produce a saturated methane mixing ratio inside the cloud layer. Our analysis for this paper is primarily based on the O14 thermal profile, although we did consider the effects of using these alternative profiles. From trial retrievals we found no significant difference in the absolute mixing ratios inferred for different thermal profiles. The main differences occurred when these mixing ratios were con- verted to relative humidities. We often found supersatu- ration above the condensation level for the cooler occul- tation profiles, whereas the same mixing ratios did not lead to supersaturation for the warmer O14 profile, or for the F0 profile. The F0 profile would also have been a decent baseline choice, as long as we did not also use the F0 methane profile, and instead let the STIS spectra constrain the methane mixing ratios without regard to occultation consistency. 8.3. Vertical Profiles of Methane In our prior analysis the vertical profile of methane was generally coupled to the vertical temperature pro- file so that the vertical variation of atmospheric re- fractivity was consistent with occultation measurement of refractivity. In our current analysis we uncoupled temperature and methane profiles because of questions 16 0.01 A 0.10 ) r a b ( P 1.00 Conrath et al. 1987 Orton et al. 2014 D profile F0 profile F1 profile B C 0.1 1.0 dotted = saturation VMR Lellouch et al. 2015 60 80 T (K) 100 -10 -5 0 5 T - TO (K) 10 10-5 10-4 10-3 CH4 VMR 10-2 10-1 Fig. 13.- A: Alternate Uranus T(P) profiles. F0 and D profiles were derived from radio occultation measurements (Lindal et al. 1987) assuming a helium VMR of 0.152. The F1 profile was also derived from radio occultation measurements, but using a lower helium VMR 0.1155, following Sromovsky et al. (2011). The Voyager IRIS profile of Conrath et al. (1987) (thick gray curve) is in best agreement with the F0, D, and F1 profiles. The Orton et al. (2014a) profile (solid black curve), is based on Spitzer Space Telescope spectral observations. B: Temperatures relative to the Orton et al. (2014a) profile, which strongly disagrees with the radio occultation profiles in 500 mbar – 1 bar region, where it is 3 K warmer. C: Methane VMR profiles corresponding to temperature profiles shown in A, using the same line styles, with an additional estimated profile by Lellouch et al. (2015), based on Herschel far-IR and sub-mm observations. raised about the reliability of the occultation results, especially by the new temperature structure results of Orton et al. (2014a), and by the new methane measure- ments of Lellouch et al. (2015), which imply supersatu- ration for the cooler temperature profiles obtained from occultation analysis. Another difference in our current analysis is that we included the parameters describing the methane vertical distribution as adjustable param- eters in the fitting process. We first carry out fits of spectra at different latitudes assuming a vertically in- variant (but adjustable) methane VMR (α0) below the condensation level. Slightly above the condensation level we fit a relative humidity rhc, and assume a minimum relative humidity of rhm at the tropopause between 20% and 60%, which yields mixing ratios within a factor of two of Orton et al. (2014a)). The high end of this range is in better agreement with Lellouch et al. (2015). The STIS spectra themselves are not very sensitive to the exact value at the tropopause, as evident from Fig. 1. Between the tropopause and the condensation level we interpolate relative humidity between rhc and rhm using the function rh(P ) = rhm + (rhc − rhm) × (cid:2)1 − log(Pc/P )/ log(Pc/Pm)(cid:3), (2) where Pc is the pressure at which CH4 condensation would occur for the given thermal profile and a given uniform deep methane VMR, and Pm is the pressure at which the relative humidity attains its minimum value near the tropopause. Given a deep methane VMR (α0) and a temperature profile from which a condensation pressure can be defined, Eq. 3 then defines a methane VMR as a function of pressure for P < Pc, denoted by α(P ). That profile is generated prior to application of the Sromovsky et al. (2011) "descended profile" function in which the initial mixing ratio profile α(P ) is dropped down to increased pressure levels P ′(α) using the equa- tion P ′ = P × [1 + (α(P )/α0)vx(Pd/Pc − 1)] for Ptr < P < Pd, (3) where Pd is the pressure depth at which the revised mix- ing ratio α′(P ) = α(P ′) equals the uniform deep mixing ratio α0, Pc is the methane condensation pressure be- fore methane depletion, Ptr is the tropopause pressure (100 mb), and the exponent vx controls the shape of the profile between 100 mb and Pd. Sample plots of de- scended profiles are displayed in Fig. 14. The profiles with vx = 1 are similar in form to those adopted by Karkoschka and Tomasko (2011). Our prior analysis ob- tained the best fits with vx=3, while our current analysis obtains a latitude dependent value ranging from ≥ 9 at low latitudes to 2.4±0.7 at 70◦ N. Fig. 14B displays an alternative step-function deple- tion model in which the methane mixing ratio decreases from the deep value to a lower vertically uniform value beginning at pressure Pd and continuing upward until the condensation level is reached for that mixing ratio. This is parameterized by four variables: the deep mixing ratio α0, the pressure break-point Pd , the upper mixing ratio α1, and the relative humidity immediately above the condensation level rhc. The parameters of all three of these vertical profile models are summarized in Table 4. 8.4. Cloud models 8.4.1. Prior cloud models Our prior analysis used an overly complex five-layer model that was based on the KT2009 four-layer model, with the main difference being replacement of their main Henyey-Greenstein (HG) layer with two layers, the higher of which was a Mie-scattering layer that was a putative methane condensation cloud, as illustrated in Fig. 15A. In this model the scattering properties of the three remaining Henyey-Greenstein layers were taken from KT2009, with no adjustment to improve fit qual- ity. This model was partly based on parameters tuned to fit the 2002 STIS observations, taken 13 years before our most recent ones, and thus it was appropriate to reconsider the aerosol structure. In addition, the five- layer model actually has too many parameters to mean- ingfully constrain independently with STIS observations. Our starting point consisted of three Mie-scattering sheet clouds, as illustrated in Fig. 15B. But we obtained fits of comparable quality using for the tropospheric aerosols a simpler single cloud of uniform scattering properties and uniformly mixed with the gas between top and bottom boundaries, as in in Fig. 15C. As a result, that simpler model became our baseline model. Tice et al. (2013), Irwin et al. (2015), and de Kleer et al. (2015) were suc- 17 cessful in using a similar model structure to fit near-IR spectra. 8.4.2. Simplified Mie-scattering aerosol models We have two options for our 2-cloud baseline model displayed in Fig. 15C. Both options use a sheet cloud of spherical Mie-scattering particles to approximate the stratospheric haze contribution. The parameters defining a sheet cloud of spherical particles are the size distribu- tion of particles, their refractive index, effective pressure, and optical depth. We chose the Hansen (1971) gamma distribution, characterized by an effective radius and ef- fective dimensionless variance. As spectra are not very sensitive to the variance, we chose an arbitrary value of 0.1. Based on preliminary fits we chose a particle size of 0.06 µm. Other researchers have selected a slightly larger size of 0.1 µm. We also found generally low sen- sitivity to the effective pressure as long as it was suffi- ciently low. We thus chose a somewhat arbitrary value of 50 mbar, putting the haze above the tropopause. We made an arbitrary choice of 1.4 for the layer's refractive index. At wavelengths shorter than our lower limit, the haze undoubtedly provides some absorption, as noted by KT2009, but we did not need to include stratospheric haze absorption to model its effects in our spectral range. Usually the only adjustable parameter for this layer we took to be the optical depth. Test calculations showed that an extended haze spanning pressures from 1 mbar to 200 mbar worked almost as well as our sheet cloud model. It did produce a slightly larger χ2 but using a diffuse stratospheric haze model had little effect on de- rived parameter values. The optical depth of the haze only increased by 2%, and the fractional changes in all the other fitted parameters were less than 0.4%, putting these changes well below their estimated uncertainties. In any case, our aim with the haze model was to ac- count for its spectral effects, not to accurately describe the physical characteristics of the haze itself. For a tropospheric sheet cloud of conservative particles the fitted parameters would be particle size, real refrac- tive index, effective pressure, and optical depth (4 pa- rameters). For a pair of tropospheric sheet clouds, as in Fig. 15B, there would be 8 parameters to constrain. As- suming both layers had the same scattering properties, that would drop the number of fitted parameters to 6. Replacing the pair of sheet clouds with a single diffuse layer with uniform scattering properties, as in Fig. 15C, reduces the number of optical depths to one, but keeps the number of pressure parameters to two, this time used 18 ) ) r r a a b b ( ( e e r r u u s s s s e e r r P P c c i i r r e e h h p p s s o o m m A A t t 10-1 10-1 CH4RHM = 0.35 CH4RHM = 0.22 10-1 10-1 100 100 Pcond= 1.21 bars vx = 0.5 1 3 CH4RHC = 0.6 Pd = 5 bars ) ) r r a a b b ( ( e e r r u u s s s s e e r r P P c c i i r r e e h h p p s s o o m m A A t t C H 4 Saturation Profile 100 100 Pcond= 0.93 bars CH4RHC = 0.35 Pd = 2.72 bars 101 101 A 10-5 10-5 10-4 10-4 10-3 10-3 10-2 10-2 Volume Mixing Ratio Volume Mixing Ratio CH4 He 101 101 B 10-1 10-1 10-5 10-5 10-4 10-4 CH4 He 10-1 10-1 10-3 10-3 10-2 10-2 Volume Mixing Ratio Volume Mixing Ratio Fig. 14.- A: Sample "descended gas" methane profiles with pd = 5 bars and vx = 0.5 (dashed), 1 (dot-dash), and 3 (solid). The starting profile before descent is shown in solid gray and is based on the F1 T(P) profile with methane constrained by its deep mixing ratio and the humidities above the condensation level (CH4RHC) and at the tropopause (CH4RHM), with linear in log P interpolation between these levels. B: Sample step-function vertical methane profile using the T(P) profile of Orton et al. (2014a) to define the saturation vapor pressure profile (dotted curve). This particular example fits the 2015 spectra at 40◦N. See text for further explanation. Methane vertical profile model parameters. TABLE 4 Model Type uniform deep 2-step uniform descended Parameter (description) α0 (deep mixing ratio) Pc (condensation pressure) Pt (tropopause pressure) rhc (relative humidity at 0.95 × Pc) rhm (relative humidity at Pt) α0 (mixing ratio for P > Pd) α1 (mixing ratio for Pc < P < Pd) Pd (transition pressure) Pc (condensation pressure) rhc (relative humidity at 0.95 × Pc) rhm (relative humidity at Pt) α0 (mixing ratio for P > Pd) Pd (transition pressure) vx (exponent of shape function) α′(P ) (descended VMR profile) rhc (relative humidity at 0.95 × Pc) rhm (relative humidity at Pt) Value adjustable derived from α0, P (T ) profile derived from P (T ) profile adjustable adjustable, or from Orton et al. (2014a) adjustable adjustable adjustable derived from α1, P (T ) profile adjustable fixed at various values adjustable adjustable adjustable derived by inverting Eq. 4 adjustable fixed at various values NOTE: we assumed the same mixing ratio for P < Pt as for P = Pt. For the 1 and 2-step uniform models rh(P ) for Pt < P < 0.95 × Pc is obtained from Eq. 3. for top and bottom boundaries, yielding a new total of 5 parameters for the tropospheric aerosols. Instead of fitting top and bottom pressures to control the vertical distribution, Tice et al. (2013) chose to fit the base pres- sure and the particle to gas scale height ratio. Which approach is more realistic remains to be determined. At this point we have a nominal total of 6 adjustable pa- rameters to describe our aerosol particles, one for the stratospheric sheet, and five for the vertically extended tropospheric layer. These are named m1 r, m2 r, m2 pb, m2 pt, m2 od, and m2 nr, where the characters preced- ing the number indicate the type of particle (m denotes Mie scattering spherical particle), the number is the layer number, and the type of parameter is indicated after the underscore (r for radius, pb for bottom pressure, pt for top pressure, od for optical depth, and nr for real refrac- 19 0.01 0.1 A m1 hg1 ) r a b ( e r u s s e r P 0.01 B 0.01 0.01 C m1_r, m1_p, m1_od m1_r, m1_p, m1_od 0.1 0.1 0.1 HG- or Mie-Scattering Sheets Diffuse DHG, HG or Mie Layer 1 m2, m2_p, m2_r CH4 cloud 1 m2_r, m2_n, m2_p, m2_od 1 1 hg2, hg2_p H2S? m3_r, m3_n, m3_p, m3_od Ptop (hg2_pt or m2_pt) HG: hg2_ssa, _g, _od DHG: hg2_ssa, _f1, _g1, _g2, _od Mie: m2_r, _nr, _ni, _od Pbot (hg2_pb or m2_pb) hg3 NH4SH? 10 0.001 0.01 0.1 1 10 Optical depth per bar 10 100 0.001 0.01 Optical depth 0.1 10 10 1 10 0.001 0.001 0.01 0.01 Optical depth 0.1 0.1 1 1 10 10 Fig. 15.- A: Comparison of the KT2009 model (dotted) with the similar but more complex 5-layer model used by Sromovsky et al. (2014), which replaced two diffuse layers with 3 compact layers. B: A preliminary simplified model with three compact layers, mostly defined by the two lower layers. This model was constructed with the possibility in mind that m2 might be formed from methane and m3 from H2S. C: Our baseline simplest model in which the tropospheric cloud is uniformly mixed between top and bottom pressures and has the same particle properties throughout layer 2. tive index). For these spherical (Mie-scattering) particles, wave- length dependent properties are controlled by particle size and refractive index. Even if both of these are wavelength-independent, scattering cross section (or op- tical depth) and phase function do have a wavelength dependence because of the physical interaction of light with spherical particles. Where our chosen parameters fail to provide sufficient wavelength dependence, we will also add another parameter, namely the imaginary re- fractive index m2 ni, which will in general be wavelength dependent, and have its main influence over the single- scattering albedo . We also have between two and four parameters chosen to constrain the vertical methane pro- file, yielding generally between eight and ten total pa- rameters to constrain by the non-linear regression rou- tine. 8.4.3. Non-spherical aerosol models. Because the particles in the atmosphere of Uranus are thought to be mostly solid particles, they are unlikely to be perfect spheres, and thus we also considered a more generalized description of their scattering properties. To investigate non-spherical scattering, we employed the commonly used double Henyey-Greenstein phase func- tion, in which three generally wavelength-dependent pa- rameters need to be defined: the scattering asymme- try parameter (g1 > 0) of a mainly forward scatter- ing term, the asymmetry parameter (g2 < 0) of the mainly backscattering term, and their respective frac- tional weights (f1 and 1 − f1 respectively). An addi- tional fourth parameter is the single-scattering albedo (), which might also be wavelength dependent. The double Henyey-Greenstein (DHG) phase function is given by P (θ) = f1 ×(1 − g2 +(1 − f1) × (1 − g2 1)/(1 + g2 1 − 2g1 cos(θ))3/2 (4) 2)/(1 + g2 2 − 2g2 cos(θ))3/2, where θ is the scattering angle. KT2009 modeled their results assuming g1 = 0.7 and g2 = −0.3 and used a wavelength-dependent f1 to adjust the phase function of their tropospheric cloud layers so that they would ap- pear relatively bright enough at short wavelengths. For haze layers composed of fractal aggregate particles, as inferred to exist in Titan's atmosphere, one would ex- pect both phase function and optical depth to be wave- length dependent, and modeling the fractal aggregate phase function variation with double Henyey-Greenstein functions would require wavelength dependence in g1 and g2 as well as f1, judging from the aggregate models of 20 Rannou et al. (1999). An alternate approach to match- ing observed spectra with spherical particles is to make the particles absorbing at longer wavelengths and con- servative at shorter wavelengths. The simplest DHG particle is just an HG particle char- acterized by an asymmetry parameter g, and a single scattering albedo , and for a limited spectral range, a wavelength dependence parameter, which can be taken as a linear slope in optical depth, which amounts to three parameters (g, , dτ /dλ). This is the same number needed to characterize scattering by a Mie particle (r, nr, ni). However, if we use a full DHG formulation, then there are five particle parameters to constrain (f1, g1, g2, , and dτ /dλ). An alternative way to produce the wavelength depen- dence of a spherical particle without its potentially com- plex phase function, containing features like glories and rainbows, which would not be seen in randomly oriented solid particles, is to follow the procedure of Irwin et al. (2015). They computed scattering properties of spherical particles to determine the wavelength dependence of the scattering cross section, but fit the phase function to a double HG function to smooth out the spherical particle features. The refractive index they assumed was the typ- ical value of 1.40 at short wavelengths, but was modified by the Kramers-Kronig relation to be consistent with the fitted variation of the imaginary index. Whether there are any cases of randomly oriented solid particles actually displaying these modified Mie scattering characteristics remains to be determined. 8.4.4. Fractal aggregate particles For those layers that are produced by photochemistry, it is also plausible that the hazes might consist of fractal aggregates, which have phase functions that are strongly peaked in the forward direction, but are shaped at other angles by the scattering properties of the monomers from which the aggregates are assembled. It is a con- venience to assume identical monomers, and to parame- terize the aggregate scattering in terms of the number of monomers, the fractal dimension of the aggregate, and the potentially wavelength dependent real and imaginary refractive index of the monomers (Rannou et al. 1999). If the refractive index were wavelength independent, this would require fitting of potentially five parameters (rm, Nm, dim, nr, ni), the same number as for the most gen- eral DHG particle. Assuming ni = 0, rm = fixed size, this would require fitting just three parameters (Nm, dim, nr), a tractable task, but one which we have not so far implemented in our fitting code. To better understand the wavelength dependent prop- erties of aggregates we made some sample calculations. We first considered an aggregate of 100 monomers 0.05 µm in radius with a real refractive index of 1.4, and a fractal dimension of 2.01. These particles have the mass of a particle of 0.23 µm in radius. This provides a physical connection between monomer parameters and the wavelength dependent aggregate phase function and scattering and absorption cross sections. We found that it is possible to at least roughly characterize the fractal aggregate phase functions with double Henyey- Greenstein functions, although this provides no physical connection to a wavelength dependent cross-section and single-scattering albedo unless DHG fits to the fractal ag- gregates are done for each wavelength. We found for this example that the backscatter phase function amplitude declines as wavelength decreases, opposite to the model of KT2009, while the scattering efficiency (and thus op- tical depth) has a strong wavelength dependence, also contradicting the KT2009 model, which assumed wave- length independence for optical depth. By increasing the number of monomers from 100 to 500 (mass equivalent to a particle 0.4 µm in radius), the asymmetry param- eter can be made relatively flat over the 0.5 µm to 1 µm range, but the strong wavelength dependence of the extinction efficiency remains, suggesting that it is opti- cal depth dependence on wavelength that offers the best lever for adjusting model I/F spectra, rather than the phase function. It is also clear that no spherical parti- cle can simultaneously reproduce both the fractal phase function and scattering efficiency and their wavelength dependencies. 8.4.5. Photochemical vs. condensation cloud models According to Tomasko et al. (2005), the dominant aerosol in Titan's atmosphere is a deep photochemical haze extending from at least 150 km all the way to the surface, with a smoothly increasing optical depth reach- ing a total vertical optical depth of 4-5 at 531 nm, with no evident layers of significant concentration that might suggest condensation clouds (only a thin layer of 0.001 optical depths was seen at 21 km). KT2009 argued for a similar origin for the dominant aerosols on Uranus. The fact that the main aerosol opacity on Uranus is found somewhat deeper than would be expected for a methane condensation cloud certainly suggests that the aerosols in the 1.2-2 bar region are either H2S, which might con- dense as deep as the 5-bar level or higher, or some pho- tochemical product, or both. And residual haze parti- cles might serve as condensation nuclei for H2S. This pu- tative deeper photochemical haze is apparently not the haze modeled by Rages et al. (1991), which is produced at very high levels of the atmosphere and has UV ab- sorbing properties that do not seem to be characteristic of the deeper haze. In fact, it is not clear that there is enough penetration of UV light to the 1.2-bar level to produce significant photochemical production of any haze material. Ignoring the issue of production rate, the main arguments for a photochemical haze are based on the following expected characteristics of such a haze: (1) a strong north-south asymmetry before the 2007 equinox, with more haze in the south compared to the north; (2) a declining haze near the south pole as solar insolation decreased towards the 2007 equinox (this assumes that the lag between production and insolation is only a few years); (3) an increasing haze near the north pole as it starts to receive sunlight after the 2007 equinox; (4) slow changes because the sub-solar latitude changes by only 4◦/year; (5) a time lag with respect to solar insolation because haze particles accumulate after production but do not exist at the beginning of production (equilibrium would be reached when the fall rate of particles equals the production rate). All five characteristics are indeed observed for Uranus, at least qualitatively, while these changes are not obvious expectations for condensation clouds. Given our preferred explanation for the polar methane depletion, namely that there is a downwelling flow from above the methane condensation level, the mixing ratio of methane would be too low to allow any methane con- densation in the polar region at pressures greater than about 1 bar. Thus it is challenging to explain the in- crease in haze in the polar region after equinox as an increase in the mass of condensed particles in that re- gion. One possibility is that the clouds are formed below the region of downwelling methane, and instead in a re- gion of upwelling H2S. But microwave observations sug- gest that the polar subsidence extends deeper than the deepest aerosol layers that we detect, which would seem to inhibit all cloud formation by condensation. Another possibility is that meridional transport of condensed H2S particles at the observed pressures, if it occurred at a suf- ficiently high rate, could resupply the falling particles. One odd feature of the putative tropospheric photo- chemical haze in the KT2009 model, is the concentration of optical depth within the 1.2-2 bar region, which has about 2 optical depths per bar, which far exceeds the 21 density of any of the other four layers in the KT2009 model. A possible explanation of this effect is that the photochemical aerosols absorb significant quantities of methane, as appears to have occurred in Titan's atmo- sphere (Tomasko et al. 2008), growing larger and also di- luting the UV absorption of the particles originating from the stratosphere. The bottom boundary of this region of enhanced opacity may be where the methane that was adsorbed into the photochemical aerosols is released and evaporated. A problem with this concept is that it is also hard to explain the growth of the haze following equinox in a region of greatly reduced methane abundance. Another mystery is why the methane mixing ratio is so stable over time, if methane is involved in fattening the photochemical particles that have a time varying produc- tion. This might just be due to the fact that it takes very small amounts of condensed material to produce a signif- icant optical depth of particulates. The rate limiting fac- tor might be the arrival rate of UV photons, rather than the amount of methane either as the parent molecule of the photochemical chain of events in the stratosphere, or as the adsorbed material needed to enhance the optical depth of the haze particles in the troposphere. We can hope that some clues can be gleaned from the character- istics of the time dependence and latitude dependence observed in the model parameters. 8.5. Fitting procedures To avoid errors in our approximations of Raman scat- tering and the effects of polarization on reflected inten- sity, we did not fit wavelengths less than 0.54 µm. An upper limit of 0.95 µm was selected because of significant uncertainty in characterization of noise at longer wave- lengths. To increase S/N without obscuring key spectral features, we smoothed the STIS spectra to a FWHM value of 2.88 nm. We chose three spectral samples of the CTL variation, at view and solar zenith angle cosines of 0.3, 0.5, and 0.7, which are fit simultaneously. In its simplest form our multi-layer Mie model has three ad- justable parameters per layer (pressure, particle radius, and optical depth). Each layer is assumed to be a sheet cloud of insignificant vertical thickness. We also fit adjustable gas parameters, illustrated in Fig. 14 and described in Table 4. For the vertically uniform mixing ratio model (up to the CH4 condensa- tion level) we have two adjustable parameters: the deep methane volume mixing ratio and the relative methane humidity above the condensation level (methane relative humidity is the ratio of its partial pressure to its satu- 22 Summary of 2 layer cloud model parameters TABLE 5 Layer Description 1 2 Stratospheric haze of Mie particles with gamma size distribution (m1) Upper tropospheric haze layer of Mie particles (m2) Alternate upper trop. haze of HG particles (hg2) Parameter (function) m1 p (bottom pressure) m1 r (particle radius) m1 b (variance) n1 (refractive index) m1 od (optical depth) Value fixed at 60 mb fixed at 0.06 µm fixed at 0.1 nr=1.4, ni=0 adjustable m2 pt (top pressure) m2 pb (bottom pressure) m2 r (particle radius) m2 b (variance) m2 nr (real refractive index) m2 ni (imag. refractive index) m2 od (optical depth) hg2 pt (top pressure) hg2 pb (bottom pressure) 2(λ) (single-scatt. albedo) g (defines HG phase func.) hg2 od (optical depth) hg2 kod (optical depth slope) hg2 pt (top pressure) hg2 pb (bottom pressure) adjustable adjustable adjustable fixed at 0.1 adjustable adjustable adjustable adjustable adjustable adjustable or fixed adjustable adjustable adjustable adjustable adjustable adjustable or fixed DHG function of KT2009 adjustable Second alternate upper tropospheric haze of double-HG particles 2(λ) (single-scatt. albedo) (hg2) P2(θ, λ) (phase function) hg2 od (optical depth) ration pressure). For the 2-layer Mie-scattering aerosol model, this yields a total of 8-9 adjustable parameters (the top Mie layer has a fixed pressure and often a fixed particle size as well, with optical depth remaining as the only adjustable parameter because the others are so poorly constrained). For the step-function 2-mixing ratio gas model, we use three adjustable gas parameters: the break point pressure, and the upper CH4 mixing ratio, and the relative methane humidity above the conden- sation level, for a total of nine adjustable parameters. The third parameterization of the methane distribution, the descended gas model, also uses three adjustable pa- rameters: the pressure limit of the descent, the methane relative humidity above the condensation level (prior to descent), and the shape exponent vx. We used a modified Levenberg-Marquardt non-linear fitting algorithm (Sromovsky and Fry 2010) to adjust the fitted parameters to minimize χ2 and to estimate uncer- tainties in the fitted parameters. Evaluation of χ2 re- quires an estimate of the expected difference between a model and the observations due to the uncertainties in both. We used a relatively complex noise model following Sromovsky et al. (2011), which combined measurement noise (estimated from comparison of individual measure- ments with smoothed values), modeling errors of 1%, relative calibration errors of 1% (larger absolute calibra- tion errors were treated as scale factors), and effects of methane absorption coefficient errors, taken to be ran- dom with RMS value of 2% plus an offset uncertainty of 5×10−4 (km-amagat)−1. This is referred to in the following as the COMPLX2 error model. 9. FIT RESULTS FOR 2012 AND 2015 STIS OBSERVATIONS Here we first consider conservative fits over a wide 540-980 nm spectral range, which identifies a problem in matching the needed particle properties to fit such a wide range. That problem is then deferred by fitting the critical 730-900 nm wavelength range that provides the strongest constraints on the methane/hydrogen ra- tio, first using Mie scattering particles for all cloud lay- ers, then using an alternative model in which the main two tropospheric layers are characterized by adjustable DHG phase functions. If we assume that the methane mixing ratio is uniform up to the condensation level, we find that it must decrease with latitude by factors of 2-3 from equator to pole with different absolute levels, de- pending on whether particles are modeled as spheres or with DHG phase functions. We then consider two mod- els that restrict methane depletions to an upper tropo- spheric layer, and find that improved fits are obtained with models that restrict depletions to the region above the 5-bar level. 9.1. Initial conservative fits to the 540-980 nm spectrum. Assuming a real refractive index of m2 nr = 1.4, and an imaginary index of zero, we fit our simplified 2-layer model to spectra covering the 540-980 nm range by ad- justing the seven remaining parameters. We obtained a best fit model spectrum with significant flaws that are illustrated in Fig. 16. The parameter values and uncer- tainties are listed in Table 6. The best-fit value for the methane mixing ratio was a remarkably low 1.27±0.05%, but is not credible because the region near 830 nm, which is most sensitive to the CH4/H2 ratio is very poorly fit. Additional flaws are seen near 750 nm, as well as at other continuum features at shorter wavelengths. Almost ex- actly the same fit quality and the same specific flaws were obtained when we replaced the single tropospheric cloud with two sheet clouds with two more adjustable parameters. Better results were obtained by letting the real refrac- tive index be a fitted parameter as well. This is in con- trast to the common procedure of fixing the refractive index, most often at a value of 1.4, as we also did in our initial fit. Irwin et al. (2015), for example, justified their choice of 1.4 by noting that most plausible condensables have real indexes between 1.3 (methane) and 1.4 (ammo- nia). Other simple hydrocarbons are also in this range. However, at the levels where we see significant aerosol op- tical depth, ammonia is not very plausible, and methane is in doubt because most particles are found at pressures exceeding the condensation level. On the other hand, the plausible condensable H2S has a significantly larger real index of 1.55 (Havriliak et al. 1955) at the 80 K – 100 K temperatures characteristic of the main cloud layer on Uranus. Another possible cloud particle is a com- plex photochemical product, one example of which is the tholin material described by Khare et al. (1993), which has a real index near 1.5. Thus, it seems premature to settle on a fixed value at this point. When the initial fit is redone with starting values of m2 r = 1 µm and m2 nr = 1.4, as documented in Ta- ble 6, we obtain a final large particle solution of m2 r = 1.918±0.33 µm and m2 nr = 1.184±0.02. Although this is an improved fit, there are still the same signifi- cant, though slightly smaller, local flaws and the inferred methane mixing ratio is again at a quite low value, this time 1.20±0.15%. A considerably better fit is obtained with the small particle solution, which produced a de- crease in χ2 /N to 0.91. This solution was obtained by using an initial guess of m2 r = 0.5 µm and m2 nr = 1.4. 23 As also shown in Table 6, these parameters adjusted to best-fit values of m2 r = 0.235±0.03 µm and m2 nr = 1.83±0.09. The real index in this case is even larger than the expected value for H2S, and the methane VMR has increased to a more credible 1.90±0.13%. However, even this fit has a few significant local flaws, near 550 nm, 590 nm, and 750 nm. Our interpretation of this situation is that there are wavelength dependent properties to the particle scattering that are not captured by conservative spherical particle models. This suggests that problems in fitting the wavelength dependent I/F over a wide range interfere with attempts to constrain the methane mix- ing ratio. Thus we decided to separate these problems. Leaving wavelength-dependence for the moment, we next focus on a narrower spectral region that provides the best constraint on the methane mixing ratio. 9.2. Fitting the 730–900 nm region Our next step was to concentrate on the spectral re- gion where the ratio of methane to hydrogen is best con- strained, i.e. the 730–900 nm region. As shown if Fig. 1, the short-wavelength side is free of CIA and sensitive to the deep methane mixing ratio, while the middle region from about 810 to 835 nm is strongly affected by hy- drogen CIA, and the long-wavelength side of the region is sensitive to the methane mixing ratio at pressure less than 1 bar. By using this entire region we expect to ob- tain good constraints on both the ratio of methane to hy- drogen as well as on the vertical cloud structure. Results from fitting this region should not be strongly affected by wavelength-dependent particle properties, given the rel- ative modest spectral range we are considering here. If the assumption of Mie scattering over this limited range is seriously flawed, that should show up in an inability to get high quality fits. This relatively narrow spectral range also weakens constraints on particle size, as might be expected. 9.2.1. Effects of different aerosol models We were somewhat surprised to find that the kind of aerosol model chosen to fit the observations has a signif- icant effect on the derived vertical and latitudinal dis- tribution of methane. To investigate these effects we did model fits at two key latitudes: 10◦N and 60◦N. From more detailed latitudinal profiles discussed later, we know that the apparent methane mixing ratio peaks near 10◦ N and is approaching its polar minimum near 60◦ N. These are also two latitudes for which 2012 and 2015 observations provide good samples at the three view angle cosines we selected. 24 A 0.10 F / I stis_fitctl_spec_glats-30.00to87.00year2015mulim0.175Jun10-155158-2016limbcorr2.88nm.unf_lmfit_Dec27-142120-2017.tab Planetographic Latitude = 10o N plot_spec_anal.pro: CASENUM = 13 χ2 = 469.29 Measured Spectrum at mu= 0.3 0.5 0.7 Best fit model with n1 = n2 = 1.4 + 0i at µ= 0.3 µ= 0.5 µ= 0.7 B C 0.6 0.6 0.6 0.7 0.7 0.8 0.8 0.7 Wavelength (µm) 0.8 0.9 0.9 0.9 1.0 0.01 1.2 1.1 1.0 0.9 0.8 0.7 2 0 -2 -4 0.5 . s a e M / l e d o M . t r e c n U / ) . s a e M - l e d o M ( Fig. 16.- Top: Model spectra at three view angle cosines (colored as noted in the legend) compared to the 10◦ N 2015 STIS spectra (black curves). Middle: Ratio of model to measured spectra. Bottom: Difference between model and measured spectra divided by expected uncertainty. The aerosol model used the baseline 2-cloud model, except that m2 nr was fixed at a value of 1.4. Best fit parameter values are given in Table 6. Note the significant discrepancies at short wavelength continuum peaks, near 740 nm, and within the critical region near 830 nm which is most sensitive to the methane to hydrogen ratio. Better fits were obtained with m2 nr allowed to adjust as part of the fitting process. Preliminary fits to the 540-900 nm part of 2015 STIS 10◦ N spectra. TABLE 6 Parameter Value for LP soln. Value for SP soln. Name m2 nr fixed at 1.4 with m2 nr fitted with m2 nr fitted Value m1 od at λ = 0.5 µm m2 od at λ = 0.5 µm m2 pt (bar) m2 pb (bar) m2 r (µm) m2 nr α0 (%) ch4rhc χ2 χ2 /NF 0.046± 0.01 5.155± 0.46 1.149± 0.04 4.137± 0.25 0.382± 0.03 1.400 1.270± 0.05 0.986± 0.13 469.29 1.16 0.050± 0.01 7.536± 1.42 1.054± 0.04 4.102± 0.25 1.918± 0.33 1.184± 0.02 1.380± 0.07 1.200± 0.15 434.53 1.07 0.048± 0.01 2.437± 0.24 0.962± 0.04 3.696± 0.22 0.235± 0.03 1.828± 0.09 1.900± 0.13 1.200± 0.15 371.03 0.91 NOTES: In the last two columns LP soln. denotes large particle solution and SP soln. denotes small particle solution. The χ2 values given here are based on fitting points spaced 3.2 nm apart. 9.2.2. Fitting the spherical particle 2-cloud model assuming a uniform CH4 distribution. We first consider a methane vertical distribution that has a constant mixing ratio from the deep atmosphere to the condensation level. Above that level (at lower pressures) we assume a drop in relative humidity to an adjustable fraction of the saturation vapor pressure, and from there to the tropopause we interpolate from the above cloud value to the tropopause minimum as de- scribed in Section 8.2. The key parameters describing the methane distribution are then the above cloud rela- tive humidity and the deep mixing ratio. We first consider a simple aerosol model in which the tropospheric contribution is characterized by an ad- justable optical depth and a single layer of spherical par- ticles bounded by top and bottom pressures and uni- formly mixed with the gas. We assume initially that these particles scatter light conservatively, but allow the real refractive index to be constrained by the spectral observations. The results of this series of fits for both 2015 and 2012 observations are given in Table 7 where small particle so- lutions are given in the first four rows and large-particle solutions in the remaining four rows. The model spec- tra are compared to the observations in Fig. 17. These fits do achieve their intended result of providing more precise constraints on the above-cloud methane humid- ity, which is high at 10◦N and about 50% of those levels at 60◦ N. The temporal change between 2012 and 2015 in the effective methane mixing ratios is very small and well within uncertainty limits. The low latitude values of 3.14±0.45% and 3.16±0.50% are consistent with no change, as are the 60◦N values, which are 0.99±0.08% and 0.93±0.08%, for 2015 and 2012 respectively. The factors by which the effective methane mixing ratio de- clines with latitude are 3.17 and 3.40 for 2015 and 2012 respectively. For these fits, the refractive index results for the small- particle solution have a weighted average for both lati- tudes and both years of 1.68±0.11, which is much closer to the 1.55 value expected for H2S, although the 60◦N values exceed that value by slightly more than their un- certainties. Perhaps this is an indication of a cloud com- position difference between the two latitudes. A quite different result is obtained for the large-particle fit. In this case the average index is 1.23±0.03, which is lower than that of any of the candidate substances, and the individual values don't vary much from low to high lati- tudes, or between 2012 and 2015. The various determinations of the pressure boundaries of the main tropospheric cloud layer are very similar for both years, both latitudes, and both particle-size so- lutions, extending from a base near 2.5 bars to a top near 1.1 bar. The optical depths do differ substantially between large and small particle solutions because the larger particles are more forward scattering and have a lower refractive index, both differences reducing the back-scattering efficiency of the particles, requiring in- creased optical depth to make up for the losses. These results explain the brightening of the polar re- gion at pseudo-continuum wavelengths between 2012 and 2015. To understand how influential these various pa- rameters are on the observed spectrum, we computed 25 logarithmic spectral derivatives (Fig. 18). These have the useful property of showing the fractional changes in the spectrum produced by fractional changes in the var- ious parameters used to model it. For the small particle solution we see that between 2012 and 2015 m2 od in- creased by 32% at 60◦ N, providing the main driver for the increase. According to Fig. 7, near 750 nm the I/F increased by about 20%, and according to the deriva- tive spectra in Fig. 18, the optical depth increase would account for about 11%, while the increase in refractive index by just 2.8% would increase the I/F by an ad- ditional 10%, accounting for the 20% total. However, these derivatives were computed for a latitude of 10◦N; somewhat different derivatives might be found at 60◦N. A similar increase of 33% is seen in the optical depth de- rived for the large particle solution, although in this case there is also an increase in particle size by 22%, which would also contribute significantly. Weighting these by respective factors of 0.42 and 0.45 (from Fig. 18) we ob- tain from just these parameters an I/F increase of about 24%, which is again close to the entire change observed. The small changes in inferred methane mixing ratios are increases of 6% for the small particle solution and about 10% for the large particle solution, which would yield I/F decreases of 1.2% and 1.5% for the small and large par- ticle solutions respectively, both of which are well below uncertainties. The fitting errors at high latitudes, which are most evident in the 0.75-µm region are highlighted by blue dotted ovals in Fig. 17. 9.2.3. Fitting the 2-cloud non-spherical model assuming a uniform CH4 distribution. We next consider fits in which the main tropospheric layer consists of a single particle type characterized by the simplest possible Henyey-Greenstein function, which is a one-term version of Eq. 5 characterized by a single asymmetry parameter. The vertical structure parame- terization and stratospheric haze layer parameterization are both unchanged from the spherical particle example used in the previous section. Because some wavelength dependence is required, we introduce a wavelength de- pendent optical depth using a simple linear slope, which is a parameter that is adjusted to optimize the fit. Our model is given by τ (λ) = τo × (1 + kOD × (λ − λ0)) (5) where λ0 is taken to be 800 nm. This also makes τo the optical depth at 800 nm. We could also have made the asymmetry parameter wavelength dependent instead of, or in addition to, the optical depth, but found ex- 26 Single tropospheric Mie layer fits to 10◦N and 60◦N STIS 730 - 900 nm spectra. TABLE 7 Lat. m1 od (◦) ×100 m2 od m2 pt (bar) m2 pb (bar) m2 r (µm) m2 nr α0 (%) ch4rhc χ2 YR 10 2.8±0.8 3.07±0.9 1.13±0.04 2.46±0.22 0.34±0.10 1.55±0.16 3.14±0.45 0.68±0.13 148.39 2015 60 0.1±70.7 1.45±0.3 1.02±0.02 2.53±0.13 0.25±0.09 1.86±0.30 0.99±0.08 0.31±0.18 248.62 2015 10 60 10 60 10 60 3.0±0.7 2.52±0.6 1.07±0.04 2.37±0.20 0.25±0.09 1.74±0.26 3.16±0.50 0.95±0.16 192.65 2012 2.2±1.6 1.10±0.2 1.02±0.04 2.22±0.13 0.24±0.07 1.81±0.25 0.93±0.08 0.42±0.20 196.02 2012 2.8±0.8 4.95±1.4 1.11±0.04 2.69±0.19 1.09±0.48 1.28±0.07 2.69±0.28 0.67±0.14 140.80 2015 0.8±4.5 4.28±1.1 1.07±0.03 2.96±0.16 1.75±0.52 1.23±0.05 0.81±0.05 0.39±0.24 256.71 2015 2.8±0.7 6.09±1.9 1.09±0.04 2.67±0.19 1.54±0.58 1.23±0.06 2.56±0.26 0.88±0.16 196.26 2012 3.3±1.5 3.21±0.8 1.08±0.04 2.51±0.14 1.44±0.53 1.22±0.05 0.74±0.05 0.56±0.26 192.54 2012 NOTE: The optical depths are given for a wavelength of 0.5 µm. These fits used 318 points of comparison and fit 8 parameters, for a nominal value of NF=310, for which the normalized χ2 /NF ranged from 0.48 to 0.802. cellent fits without adding any further complexity. We will not be making any claims regarding the true source of wavelength dependence in any case. Our main objec- tive is to find out how this different kind of model affects the methane distribution, and to determine the average asymmetry parameter of these particles. We will also try to infer a single-scattering albedo. Best-fit parameter values and uncertainties for fits at 10◦N and 60◦N for 2012 and 2015 are presented in Ta- ble 8. Best-fit model spectral are compared to observa- tions in the left panel of Fig. 19, while fractional deriva- tive spectra are displayed in the right panel. These fits are comparable in quality to the spherical particle fits presented in the previous section, and have the same problem fitting the high-latitude spectra, most notably in the 750-nm region. This region senses more deeply than other parts of this limited spectral range (see Fig. 1), and thus is most likely to be affected by vertical vari- ations in the methane mixing ratio. According to the derivative spectra, an increase in the methane mixing ratio with depth would reduce the I/F in this region, which we would expect to produce a better fit, and we will later show that this does in fact improve the spectral fit in this region. The methane mixing ratio values for this model average somewhat higher than found for the model using spher- ical particles, although all are within uncertainty limits for a given latitude, and all results indicate an effective mixing ratio decrease by slightly more than a factor of three from 10◦N to 60◦N. The best-fit asymmetry parameter for this model is generally near 0.4, well below the commonly assumed value of 0.6 for near-IR analysis, which is in part based on an analysis of limb-darkening measurements by Sromovsky and Fry (2008). That analysis predates the significant improvement in methane absorption coef- ficients seen in the last decade (Sromovsky et al. 2012a) and may no longer be valid. It seems unlikely that this difference is merely a wavelength dependence. For the sizes inferred for spherical particle solutions, the asym- metries either decrease with wavelength (small particle solution), or remain relatively flat (large particle solu- tion). While the asymmetry parameter is highly nega- tively correlated with the optical depth parameter, these two parameters do have sufficiently different ratios be- tween peaks and valleys to allow them to be indepen- dently determined (shown in Fig. 19F). The asymme- try was determined to within about 10% and the optical depth to within slightly better accuracy. The optical depths for this model appears to be considerably lower than for the spherical particle models, which were at a shorter wavelength of 500 nm. If we convert those Mie scattering optical depths to a wavelength of 800 nm, we find that the 0.3-µm particle optical depth drops from 3.1 to 2.6 and the 1.54 µm particle optical depth increases from 6.1 to 7.5. Thus, the wavelength difference does not explain the low optical depths of the non-spherical model. It is more likely due to the latter's more symmet- ric scattering. The Mie particle models have asymmetries of about 0.68 and 0.87 for the small and large particle 27 Case 22: 2-layer Mie model large r2 fit to 730-900 nm, 2012, 2015 STIS obs. Case 23: 2-layer Mie model small r2 fit to 730-900 nm, 2012, 2015 STIS obs. Orton T(P) Orton T(P) LAT = 10o, χ2 = 140.80, YEAR = 2015 µ= 0.70 µ= 0.50 µ= 0.30 F 0.3 0.2 F / I LAT = 10o, χ2 = 148.39, YEAR = 2015 F µ= 0.70 µ= 0.50 µ= 0.30 0.75 0.80 0.85 LAT = 60o, χ2 = 256.71, YEAR = 2015 µ= 0.70 µ= 0.50 µ= 0.30 0.75 0.80 0.85 LAT = 10o, χ2 = 196.26, YEAR = 2012 µ= 0.70 µ= 0.50 µ= 0.30 0.75 0.80 0.85 LAT = 60o, χ2 = 192.54, YEAR = 2012 µ= 0.70 µ= 0.50 µ= 0.30 0.90 G 0.90 H 0.90 I 0.1 0.0 0.3 0.2 F / I 0.1 0.0 0.3 0.2 F / I 0.1 0.0 0.3 0.2 F / I 0.1 0.0 0.75 0.80 0.85 LAT = 60o, χ2 = 248.62, YEAR = 2015 µ= 0.70 µ= 0.50 µ= 0.30 0.75 0.80 0.85 LAT = 10o, χ2 = 192.65, YEAR = 2012 µ= 0.70 µ= 0.50 µ= 0.30 0.75 0.80 0.85 LAT = 60o, χ2 = 196.02, YEAR = 2012 µ= 0.70 µ= 0.50 µ= 0.30 0.90 G 0.90 H 0.90 I 0.3 0.2 F / I 0.1 0.0 0.3 0.2 F / I 0.1 0.0 0.3 0.2 F / I 0.1 0.0 0.3 0.2 F / I 0.1 0.0 0.75 0.80 Wavelength (µm) 0.85 0.90 0.75 0.80 Wavelength (µm) 0.85 0.90 Fig. 17.- Comparison of observed spectra (curves) with model fits (points) for the large particle solutions (left) and the small particle solutions (right), both using the model parameterization defined in Table 7. Fits to 2015 STIS observations are shown in the top pair of panels and fits to 2012 observations in the bottom pair of panels. Blue dotted ovals identify regions of high-latitude fitting errors, which can be greatly reduced by using a non-uniform vertical distribution of methane. solutions respectively, both adjusted to a reference wave- length of 800 nm. The particles of KT2009, which use a DHG phase function, with their adopted values of g1 = 0.7 and g2= -0.3, yield an asymmetry of 0.6, which is much closer to that of our small particle solution for spherical particles, and much larger than our HG par- ticle solutions. We tried to find an HG solution with larger asymmetry by using a first guess with g = 0.63, but the regression again converged on g = 0.43. It is apparently the case that very different scattering proper- ties can lead to very nearly the same fit quality, but very different optical depths and asymmetry parameters. At phase angles near zero, there is a considerable ambiguity between more forward scattering particles with larger op- tical depths and more backward scattering particles with smaller optical depths. The best-fit optical depth slope parameter hg2 kod is negative, as generally expected, and is around -2/µm at 10◦N but -3.2 /µm to -4/µm at 60◦N. A spherical par- ticle of radius 0.3 µm and real index 1.4 would have a slope of about -2.4/µm. For spherical particles, some of the wavelength dependence in scattering is provided by wavelength dependence in the phase function. This sug- gests a possible decrease in particle size at high latitudes. There is better agreement between the HG solution and the small particle Mie solutions regarding other pa- rameters, including pressure boundaries, methane mixing ratios and above cloud humidities. Thus, the preponder- ance of evidence suggests that the cloud particles can be roughly approximated by the small particle Mie solu- tions, which is the solution type we will use to investigate the latitude dependent characteristics in more detail. 9.2.4. Latitude-dependent fits To illustrate the latitude dependence of the effective methane mixing ratio and the inferred aerosol distribu- 0.3 0.2 0.1 0.0 -0.1 -0.2 0.5 0.0 -0.5 -1.0 1.5 1.0 0.5 0.0 -0.5 1.0 0.5 0.0 -0.5 -1.0 E dln(I/F)/dlnm2_od 0.75 0.80 0.85 0.90 G dln(I/F)/dlnm2_n 0.01 1.0 0.5 0.0 -0.5 -1.0 1.5 1.0 0.5 0.0 -0.5 8 6 4 2 0 1.0 0.5 0.0 -0.5 -1.0 1.0 0.5 0.0 -0.5 -1.0 1.5 1.0 0.5 0.0 -0.5 G 8 6 4 2 0 0.90 0.75 I 1.0 0.5 0.0 -0.5 -1.0 1.5 1.0 0.5 0.0 -0.5 1.0 0.5 0.0 -0.5 -1.0 28 1.00 A 0.10 F / I µ=0.7 glat= 10o B dln(I/F)/dlnm1_od 1.00 A 0.10 F / I 0.01 0.75 0.80 0.85 0.90 C dln(I/F)/dlnm2_pb 0.75 0.80 D dln(I/F)/dlnm2_pt 0.85 0.90 0.75 C B dln(I/F)/dlnm1_od 0.75 0.80 0.85 0.90 D dln(I/F)/dlnm2_pt µ=0.7 glat= 10o 0.80 0.85 0.90 dln(I/F)/dlnm2_pb 0.3 0.2 0.1 0.0 -0.1 -0.2 0.5 0.0 -0.5 -1.0 0.75 0.80 0.85 0.90 0.75 0.80 0.85 0.90 dln(I/F)/dlnm2_r F 0.85 0.90 0.75 0.75 0.80 E dln(I/F)/dlnm2_od F 0.80 0.85 0.90 dln(I/F)/dlnm2_r 0.75 0.80 0.85 0.90 0.75 H dln(I/F)/dlnch4v0 0.80 0.85 0.90 dln(I/F)/dlnm2_n 0.75 0.80 0.85 0.90 H dln(I/F)/dlnch4v0 0.75 0.80 0.85 0.90 0.75 0.80 0.85 Wavelength (µm) I dln(I/F)/dlnch4rhc Derivs. taken at: m1_od= 0.028 m2_pb= 2.686 m2_pt= 1.111 m2_od= 4.954 m2_r= 1.093 m2_n= 1.275 ch4v0= 0.027 ch4rhc= 0.671 0.80 0.85 0.90 0.75 0.80 0.85 Wavelength (µm) 0.90 dln(I/F)/dlnch4rhc Derivs. taken at: m1_od= 0.028 m2_pb= 2.458 m2_pt= 1.125 m2_od= 3.070 m2_r= 0.343 m2_n= 1.553 ch4v0= 0.031 ch4rhc= 0.684 0.75 0.80 0.85 Wavelength (µm) 0.90 0.75 0.80 0.85 Wavelength (µm) 0.90 Fig. 18.- Derivative spectra for uniform mixing ratio models evaluated for the large-particle solution (Left group) and small particle solution (Right group). In each group we show I/F model spectrum (A) and derivatives of fractional changes in I/F with respect to fractional changes in parameters m1 od (B), m2 pb (C), m2 pt (D), m2 od (E), m2 r (F), m2 nr (G), ch4v0 ≡ α0 (H) and ch4rhc (I). All the derivative panels are scaled the same, except for panel B, which has been expanded by a factor of 4, and panel G, which has been compressed by a factor of 5 because of their unusually small and large effects, respectively, on the I/F spectrum. In panels F and G, the dotted curve represents a version of the m2 od derivative spectrum scaled to match the lower features of the m2 r and m2 nr derivative spectra respectively, to illustrate their strong correlations but resolvable differences. Single tropospheric HG layer fits to 10◦N and 60◦N STIS spectra. TABLE 8 Lat. m1 od (◦) ×100 hg2 od hg2 pt (bar) hg2 pb (bar) hg2 g hg2 kod (/µm ) α0 (%) ch4rhc χ2 YR 10 2.6±0.5 1.58±0.13 1.13±0.03 2.33±0.15 0.43±0.04 -2.23±0.4 3.48±0.45 0.65±0.08 151.51 2015 60 0.0±0.0 0.97±0.05 1.01±0.02 2.53±0.12 0.26±0.02 -3.18±0.3 0.97±0.06 0.31±0.03 252.65 2015 10 2.7±0.5 1.57±0.14 1.07±0.03 2.47±0.17 0.42±0.04 -1.91±0.4 2.85±0.32 0.87±0.12 192.22 2012 60 2.0±1.2 0.71±0.06 1.01±0.02 2.04±0.08 0.39±0.04 -3.95±0.3 1.04±0.07 0.39±0.15 193.30 2012 NOTE: The optical depth is for a wavelength of 0.8 microns for hg2 od, and for 0.5 µm for the stratospheric haze (m1 od). These fits used 318 points of comparison and fit 8 parameters, for a nominal value of NF=310, for which the normalized χ2 /NF ranged from 0.48 to 0.802. tion we selected the simple 2-layer model using a compact stratospheric haze and an extended diffuse layer of spher- ical tropospheric particles, characterized by Mie scatter- ing parameters of radius and refractive index. We also µ=0.7 glat= 10o 1.00 A 0.10 F / I 0.01 0.75 0.80 0.85 0.90 C dln(I/F)/dlnhg2_pd 0.75 0.80 0.85 0.90 E dln(I/F)/dlnhg2_od 0.75 0.80 0.85 0.90 G dln(I/F)/dlnhg_kod 29 B dln(I/F)/dlnm1_od 0.75 0.80 0.85 0.90 D dln(I/F)/dlnhg2_pt 0.75 0.80 0.85 0.90 F dln(I/F)/dlnhg_g 0.75 0.80 0.85 0.90 H dln(I/F)/dlnch4v0 0.3 0.2 0.1 0.0 -0.1 -0.2 0.5 0.0 -0.5 -1.0 -1.5 1.0 0.5 0.0 -0.5 -1.0 1.0 0.5 0.0 -0.5 -1.0 0.3 0.2 F / I 0.1 0.0 0.3 0.2 F / I 0.1 0.0 0.3 0.2 F / I 0.1 0.0 0.3 0.2 F / I 0.1 0.0 G H I J HG Case 7: Mie stratospheric haze, ONE HG layer, 730-900 nm, Orton T(P), uniform deep CH4vmr LAT = 10o, χ2 = 151.51, STIS 2015 µ = 0.70 µ = 0.50 µ = 0.30 0.75 0.80 0.85 0.90 LAT = 60o, χ2 = 252.65, STIS 2015 µ = 0.70 µ = 0.50 µ = 0.30 0.75 0.80 0.85 0.90 LAT = 10o, χ2 = 192.22, STIS 2012 µ = 0.70 µ = 0.50 µ = 0.30 0.75 0.80 0.85 0.90 LAT = 60o, χ2 = 193.30, STIS 2012 µ = 0.70 µ = 0.50 µ = 0.30 0.75 0.80 Wavelength (µm) 0.85 0.90 1.0 0.5 0.0 -0.5 -1.0 1.5 1.0 0.5 0.0 -0.5 0.3 0.2 0.1 0.0 -0.1 -0.2 1.0 0.5 0.0 -0.5 -1.0 0.75 0.80 0.85 0.90 0.75 0.80 0.85 Wavelength (µm) 0.90 I dln(I/F)/dlnch4rhc 0.75 0.80 0.85 Wavelength (µm) 0.90 Derivs. taken at: m1_od= 0.026 hg2_pd= 2.325 hg2_pt= 1.126 hg2_od= 1.580 hg_g= 0.426 hg_kod=-2.230 ch4v0= 0.035 ch4rhc= 0.651 Fig. 19.- Left: HG Model spectra compared to observations at 10◦N and 60◦N for 2012 (bottom pair) and 2015 observations (top pair), with models plotted as points with error bars. Right: Derivative spectra showing the ratio of a fractional change in I/F to the fractional change in the parameter producing the change (here ch4v0 ≡ α0). The dotted curve in panel F of the derivative group is an inverted plot of the curve in panel E, with minima scaled to match the solid curves. Note that the maxima do not match, making them distinguishable. chose the small-radius solution set because of their high quality fits and relative consistency between 2012 and 2015, as well as their better agreement with HG fits as noted in the previous section. Other models show similar characteristics, except that they contain more variation between years, as can be surmised from the table of fit parameters from fits at 10◦N and 60◦N, shown in Ta- ble 7 for large Mie particle fits and in Table 8 for HG model fits. We assumed a methane profile that has a vertically uniform fitted deep mixing ratio, a fitted rela- tive humidity immediately above the condensation level, a minimum relative humidity of 30%, with linear inter- polation filling in values between the condensation level and the tropopause. Above the tropopause we assumed a mixing ratio equal to the tropopause value. From fitting spectra every 10◦ of latitude for both 2012 and 2015 ob- servations we obtained the best-fit parameters and their formal uncertainties given in Table 9. The parameters are also plotted in Fig. 20, where panels A-E display the fit parameter values and their estimate errors, and pan- els F-I display samples of model and observed spectra for 10◦ N and 60◦ N for 2015 (F and G) and 2012 (H and I). Most of the model parameters are found to have only weak variations with latitude. The top pressure of the sole tropospheric cloud layer is surprisingly invariant from low to high latitudes as well as from 2012 to 2013, even though there are substantial variations in optical depth between years as well as with latitude. This boundary pressure is also very well constrained by the observations. The bottom pressure of this cloud is more variable, but its variation is not much more than its un- certainty which is much larger than that of the cloud top pressure. The larger uncertainty is consistent with the derivative spectra given in Fig. 18, which shows that, compared to the top pressure, the bottom pressure has a smaller fractional effect on the I/F spectrum for a given fractional change in pressure. 30 Single tropospheric Mie layer fits to the 730-900 nm spectra as a function of latitude assuming vertically uniform CH4 below the condensation level. [!t] TABLE 9 Lat. m1 od (◦) ×100 m2 od m2 pt (bar) m2 pb (bar) m2 r (µm ) m2 nr α0 (%) ch4rhc χ2 YR -10 2.4±0.8 3.60±1.37 1.09±0.04 2.66±0.22 0.22±0.09 1.65±0.31 2.93±0.37 0.75±0.14 180.61 2015 0 10 20 30 40 50 4.5±0.8 2.52±0.67 1.07±0.04 2.55±0.20 0.25±0.08 1.72±0.24 2.69±0.38 0.61±0.14 137.93 2015 2.8±0.8 3.07±0.88 1.13±0.04 2.46±0.22 0.34±0.10 1.55±0.16 3.14±0.45 0.68±0.13 148.39 2015 2.8±0.7 1.99±0.58 1.08±0.05 2.55±0.22 0.28±0.11 1.75±0.29 2.85±0.39 0.97±0.18 170.32 2015 3.8±0.8 1.48±0.49 1.06±0.05 2.60±0.20 0.27±0.13 1.81±0.36 2.10±0.24 0.88±0.19 170.86 2015 2.8±0.9 1.41±0.36 1.01±0.04 2.65±0.17 0.27±0.10 1.79±0.28 1.41±0.12 0.75±0.20 205.54 2015 2.4±1.1 1.25±0.42 1.01±0.04 2.51±0.14 0.26±0.16 1.88±0.46 1.13±0.09 0.76±0.22 266.49 2015 60 0.1±70.7 1.45±0.29 1.02±0.02 2.53±0.13 0.25±0.09 1.86±0.30 0.99±0.08 0.31±0.18 248.62 2015 70 0.4±13.4 1.49±0.26 1.01±0.03 2.71±0.14 0.23±0.09 1.90±0.33 0.88±0.07 0.36±0.21 278.56 2015 -20 1.4±0.7 3.14±1.19 1.11±0.04 2.71±0.22 0.24±0.09 1.66±0.29 2.87±0.36 0.77±0.13 137.66 2012 -10 4.4±0.7 3.28±1.22 1.04±0.05 2.77±0.23 0.25±0.09 1.64±0.27 2.63±0.32 1.17±0.21 147.88 2012 0 10 20 30 40 4.7±0.8 2.51±0.64 1.08±0.04 2.56±0.21 0.25±0.08 1.73±0.24 2.69±0.37 0.58±0.12 150.60 2012 3.0±0.7 2.52±0.64 1.07±0.04 2.37±0.20 0.25±0.09 1.74±0.26 3.16±0.50 0.95±0.16 192.65 2012 2.6±0.7 3.42±1.14 1.06±0.05 2.63±0.21 0.27±0.08 1.62±0.22 2.65±0.33 0.95±0.17 197.68 2012 2.7±0.7 1.98±0.51 1.02±0.05 2.69±0.19 0.26±0.08 1.74±0.24 1.99±0.21 0.89±0.18 149.64 2012 2.8±1.0 2.07±0.51 1.06±0.04 2.42±0.15 0.32±0.08 1.57±0.14 1.29±0.12 0.62±0.18 255.27 2012 50 0.4±1.55 1.94±0.39 1.07±0.03 2.47±0.13 0.32±0.06 1.59±0.12 1.03±0.08 0.30±0.17 191.32 2012 60 70 2.2±1.6 1.10±0.23 1.02±0.04 2.22±0.13 0.24±0.07 1.81±0.25 0.93±0.08 0.42±0.20 196.02 2012 1.8±2.1 1.00±0.12 1.01±0.03 2.23±0.12 0.19±0.09 1.97±0.36 0.97±0.08 0.29±0.19 235.18 2012 NOTE: The optical depths are for a wavelength of 0.5 µm. These fits used 318 points of comparison and fit 8 parameters, for a nominal value of NF=310, for which the normalized χ2 /NF ranged from 0.44 to 0.90. The most prominent latitudinally varying parameter is the effective deep methane mixing ratio, which attains a low-latitude maximum of about 3.15%, dropping to about 2% by 30◦ N, reaching a high-latitude value of about 1% at between 50◦N and 60◦N. Close behind, is the variation in methane humidity above the condensation level, which was found to be 60-100% at low latitudes, declining to about 30-40% for regions poleward of 50◦ N. This decline towards the north pole is also seen in other model types as well. There is also close agreement, for this model, between between 2012 and 2015 results for both the extremes in the methane mixing ratio and in its latitudinal variation. The slight dip at the equator is also present in results for both years, as is the peak at 10◦N. The agreement of the 2012 and 2015 methane profiles (on both the deep mix- ing ratio and the above cloud humidity) is close enough that we must look elsewhere to explain the brightening of the polar region between 2012 and 2015. The most likely aerosol change responsible for the polar brightening is the increase in the bottom cloud layer optical depth (m2 od) by about 60% at latitudes north of 50◦, a factor already discussed in Section 9.2.2. However, because multiple aerosol parameters differ between 2012 and 2015, it is useful to show that the combined effect of layer m2 pa- rameter changes does indeed result in the increased scat- tering that produced the observed brightness increase. This was done by starting with the model spectrum for 2012 and computed a new model spectrum in which only the layer-m2 parameters were changed to match those of 2015, leaving other parameters unchanged. We also computed the spectrum change when only the optical depth of the m2 layer was changed to the 2015 value. We did this at latitudes of 50◦N, 60◦N, and 70◦N. The Case 25: 2-layer Mie model small r2 fit to 730-900 nm, 2012, 2015 STIS obs. Case 25: 2-layer Mie model small r2 fit to 730-900 nm, 2012, 2015 STIS obs. Orton T(P) Orton T(P) 31 0.5 1.0 1.5 2.0 2.5 3.0 100 10-1 ) r a b ( e r u s s e r P t h p e d l a c i t p O r o , ) m µ ( s u d a R i H R H C 4 10-2 1.4 1.2 1.0 0.8 0.6 0.4 0.2 0.0 0.04 R M V e n a h t e M 0.03 0.02 0.01 ) F N ( / 2 χ 1.4 1.2 1.0 0.8 0.6 0.4 A m2_pt m2_n m2_pb -20 B m2_od m2_r m1_od -20 C ch4rhc -20 D ch4v0 STIS 2015 STIS 2012 -20 E Nfree= 310 Nfree= 310 -20 m2_pt m2_n m2_pb 0.3 0.2 F / I LAT = 10o, χ2 = 148.39, YEAR = 2015 F µ= 0.70 µ= 0.50 µ= 0.30 0 0 0 20 40 60 80 m2_od 0.1 0.0 0.3 m2_rm2_r 0.2 F / I 0.1 0.0 0.3 20 40 60 80 m1_od m1_od 0.2 F / I ch4rhc ch4rhc 20 40 60 80 0 complx2 error model 20 40 60 80 ch4v0 0.1 0.0 0.3 0.2 F / I 0.1 0.0 0.75 0.80 0.85 LAT = 60o, χ2 = 248.62, YEAR = 2015 µ= 0.70 µ= 0.50 µ= 0.30 0.75 0.80 0.85 LAT = 10o, χ2 = 192.65, YEAR = 2012 µ= 0.70 µ= 0.50 µ= 0.30 0.75 0.80 0.85 LAT = 60o, χ2 = 196.02, YEAR = 2012 µ= 0.70 µ= 0.50 µ= 0.30 0.90 G 0.90 H 0.90 I 0 Planetographic Latitude (deg) 20 40 60 80 0.75 0.80 Wavelength (µm) 0.85 0.90 Fig. 20.- Left: Single tropospheric Mie model fits as a function of latitude under the assumption that the methane VMR is constant for pressures exceeding the condensation level. Parameter values are also given in Table 9. Right: sample spectra, with blue dotted ovals identifying regions of larger I/F errors. results are summarized in the following figures. The left- hand figure provides a sample spectral view at 60◦ N. It shows the measured spectral difference between 2012 and 2015 as a shaded curve, with shading range indicat- ing uncertainties. Also shown are the difference in model fits (+), the difference due only to layer m2 differences (×), and the difference due only to the change in optical depth (o). The right hand plot displays the latitude de- pendence for two pseudo-continuum wavelengths. Again are shown the measured differences (shaded curves), the model difference (×), and the brightness change due only to layer m2 (+). This figure shows that layer m2 is clearly responsible for the vast majority of the brightness increase between 2012 and 2015, but changes in the m2 layer optical depth are only responsible for about half of the total scattering increases of that layer (as in the left hand plot), except at 50◦N, where even though the opti- cal depth decreased, the layer still brightened because of changes in particle size and refractive index). At low latitudes, the fit quality for both years is better than expected from our uncertainty estimates, but fit quality decreases significantly at high northern latitudes, especially for the 2015 fits, which have increased aerosol scattering. The high latitude fitting problem is most obvious just short of 750 nm, as shown in panels G and I of Fig. 20, where the model values exceed the measured values (note the encircled regions). This problem can be greatly reduced by using an altered vertical profile of methane, as discussed in Section 9.3. 9.2.5. Summary of uniform methane results Both spherical particle and HG models for the upper tropospheric layer lead to declining effective methane vol- ume mixing ratios with latitude by similar factors, but are in some disagreement with respect to magnitudes, as shown in greater detail in Fig. 22. The more detailed latitudinal fit results in Fig. 20 for the small-particle solu- tion, show that the effective methane mixing ratio peaks near 10◦ N in both years, has a local minimum at the equator and declines with latitude by more than a factor 32 0.08 0.06 0.04 0.02 0.00 ) 2 1 0 2 ( F / I - ) 5 1 0 2 ( F / I 60o N shaded = measured difference for µ = 0.7 RED x = model difference BLK + = difference due to m2 model changes GRN o = difference due to m2_od change only shaded = measured difference for µ = 0.7 RED x = model difference BLK + = difference due to m2 model changes 750 nm 830 nm I/F - 0.01 ) 2 1 0 2 - 5 1 0 2 ( e c n e r e f f i d F / I 0.08 0.06 0.04 0.02 0.00 0.70 0.75 0.80 Wavelength (µm) 0.85 0.90 50 65 55 Planetographic latitude (o) 60 70 Fig. 21.- Left: spectral difference at 60◦N between 2012 and 2015 observations at a zenith angle cosine of 0.7 (shaded curve) compared to all model differences (×), to those contributed only by layer m2 (+), and to those due only to the m2 optical depth change (o). Right: latitudinal variation of observed temporal differences at 750 nm (upper shaded curve) and 830 nm (lower shaded curve offset by 0.01), compared to total model differences (×) and differences due to all changes in layer m2 only (+). This shows that increased scattering by layer m2 is primarily responsible for the observed brightening of the polar region between 2012 and 2015. of two by 50-60◦ N. For each year, the two aerosol models lead to similar shapes, and in the 50-70◦ range the two models agree that there is a crossover in which the 2015 vmr declines from 50◦ to 70◦, while the 2012 vmr rises slightly over the same interval. The fitted values of the methane relative humidity just above the condensation level, shown in Fig. 22B, have considerable uncertainty. But both results indicate a peak near 20◦ N, a clear local minimum near the equa- tor, and a strong decline towards the north pole. This is suggestive of rising motions near 20◦ and descending motions near the equator and poles, with the latter being more significant. 9.2.6. The effects of particle absorption on derived methane amounts The modeling results presented so far are for conser- vative particles ( = 1.0). Particles that absorb some fraction of the incident light will act to darken the at- mosphere and reduce the amount of methane needed to fit the spectrum. This is true even if the parti- cles are not distributed vertically in the same fashion as methane, and even though they lack the band struc- ture of methane. The aerosol optical depths and derived pressure locations of the layers are also altered. To in- vestigate the magnitude of these effects we did fits of the 2 Mie layer model to the 2015 STIS observations, under the assumption of vertically uniform methane, but with the imaginary index of the tropospheric layer increased from zero to 0.0049, which, for a 0.3-µm radius particle with a real refractive index of 1.8 corresponds to a de- crease in single scattering albedo at 0.8 µm from = 1.0 to ≈ 0.979. This amount of absorption in the 730 to 900 nm part of the spectrum, makes it possible to fit the entire spectrum (down to 540 nm) if the particles are assumed to be conservative at the shorter wavelengths (see Sec. 9.4 for more information). Table 10 shows that adding this amount of absorption changes the layer-2 top pressures by just fractions of a percent, while increasing the bottom pressures by 18-22%. The optical depth of the layer changes in less consistent directions. If the par- ticle's refractive index and size did not change much, then an increase in optical depth would be required to make up for the lower single-scattering albedo produced by absorption. However, the optical depth is seen to de- crease at 10◦N, where m2 nr has increased by almost 10% (and m2 nr - 1 by 28%), increasing the scattering efficiency substantially. For 60◦N, the changes in m2 r, m2 nr, and m2 od are all substantially smaller. Most importantly, the effective deep methane mixing ratio is decreased by 3% at 10◦N and 7% at 60◦N, which sug- gests that a fair approximation of the latitudinal profile for absorbing cloud particles can be obtained by scal- ing the profile we derived from conservative scattering. Whether the cloud particles are actually absorbing in the 730-900 nm region remains to be determined. For the large-particle solution, we made a similar com- parison, but just for 10◦ N and for 2012. In this case the increase of imaginary index needed to adjust the wavelength dependence (as described above for the small- particle solution) is only from 0 to 6.2×10−4, which de- creases the single-scattering albedo for a 1.535 µm par- ticle with real index 1.225 to = 0.990 at 0.8 µm. Al- 2 compact DHG layers (w KT2009 fn) 2 compact DHG layers (w KT2009 fn) 2 compact Mie layers (small r, n=1.4) 2 compact Mie layers (small r, n=1.4) 1 diffuse HG layer (k_od fit) 1 diffuse HG layer (k_od fit) 1 diffuse Mie layer (large r, n fit) 1 diffuse Mie layer (large r, n fit) 1 diffuse Mie layer (small r, n fit) 1 diffuse Mie layer (small r, n fit) 153.41 (2015) 195.79 (2012) 175.05 (2015) 199.77 (2012) 151.51 (2015) 192.22 (2012) 140.80 (2015) 196.26 (2012) 148.39 (2015) 192.65 (2012) 0.00 0.01 0.04 Methane Volume Mixing Ratio at 10o N 0.03 0.02 33 A 0.05 2 compact DHG layers (w KT2009 fn) 2 compact DHG layers (w KT2009 fn) 2 compact Mie layers (small r, n=1.4) 2 compact Mie layers (small r, n=1.4) 1 diffuse HG layer (k_od fit) 1 diffuse HG layer (k_od fit) 1 diffuse Mie layer (large r, n fit) 1 diffuse Mie layer (large r, n fit) 1 diffuse Mie layer (small r, n fit) 1 diffuse Mie layer (small r, n fit) 253.40 (2015) 187.22 (2012) 275.59 (2015) 227.90 (2012) 252.65 (2015) 193.30 (2012) 256.71 (2015) 192.54 (2012) 248.62 (2015) 256.71 (2012) 0.000 0.005 Effective Methane VMR at 60o N 0.010 B 0.015 Fig. 22.- Effective deep methane VMR for different aerosol model parameterizations at 10◦ N (A) and 60◦ N (B). Vertical lines show unweighted mean values for 2015 (dashed) and 2012 (dotted). though this seems like a small change, it produces a 50% increase in optical depth, a 56% increase in the cloud bottom pressure, and a 10.6% decrease in the best-fit methane mixing ratio, as shown in Table 11. For non- spherical particles in which wavelength-dependent op- tical depths or wavelength dependent phase functions might be used to adjust the wavelength dependent I/F spectrum, there may be no need for absorbing particles, in which case the somewhat higher methane mixing ra- tios may apply. 9.3. Fitting latitude-dependent vertically non-uniform methane depletion models 9.3.1. Alternative models of vertically varying methane The fits discussed in previous sections have assumed that the methane profile is vertically uniform from the bottom of our model atmosphere all the way up to the methane condensation level. We have already noted problems with those fits in the 750 nm region of the spectrum, which suggest that the methane mixing ra- tio likely increases with depth at high latitudes. There are also independent physical arguments suggesting the same characteristic. Sromovsky et al. (2011) pointed out that extending the very low high latitude mixing ratios to great depths would result in horizontal density gradi- ents over great depths. As a consequence of geostrophic and hydrostatic balance, these gradients would lead to vertical wind shears (Sun et al. 1991). This would re- sult in an enormous wind difference with latitude at the visible cloud level, which would be inconsistent with the observed winds of Uranus. Thus, we would ex- pect that the polar depletion would be a relatively shal- low effect, as we have inferred from our previous work (Karkoschka and Tomasko 2009; Sromovsky et al. 2011, 2014). As indicated by KT2009, the 2002 spectral obser- vations did not require that methane depletions extend to great depths, and Sromovsky et al. (2011) showed that shallow depletions were preferred by the 2002 spectra. This was further supported by de Kleer et al. (2015), who used our descended profile parameterization, fixed the shape parameter at vx = 2, and constrained the depth parameter vs latitude using H band spectra. They found a clear latitude trend, with a low-latitude value of 1.7±0.2 bars, increasing to 3.2±1 bars in the 40–50◦N band, and as deep as 26+11 −18 bars in the 60–70◦N band, although at that extreme value the depth parameter is constrained more by the shape of the profile at much lower pressures than by any direct sensing of sunlight 34 Changes in small-particle best-fit parameter values derived from the STIS 2015 observations, as a result of adding absorption to aerosol layer 2 by increasing m2 ni from 0.0 to 0.005. TABLE 10 Parameter Value Value Value Value 10◦N Latitude 60◦N Latitude Name m2 od m2 pt (bar) m2 pb (bar) m2 r (µm) m2 nr α0 ×100 ch4rhc χ2 m2 ni = 0 m2 ni = 0.005 Difference m2 ni = 0 m2 ni = 0.005 Difference 3.084 1.126 2.450 0.342 1.554 3.160 0.687 148.03 2.864 1.127 2.993 0.307 1.706 3.060 0.701 148.85 -7.15% 0.14% 22.12% -10.16% 9.77% -3.16% 2.04% 0.55% 1.445 1.023 2.519 0.248 1.862 0.989 0.318 248.29 1.482 1.032 2.968 0.256 1.900 0.916 0.355 246.28 2.54% 0.87% 17.79% 2.89% 2.05% -7.38% 11.64% -0.81% TABLE 11 Changes in large-particle best-fit parameter values derived from the STIS 2015 observations, as a result of adding absorption to aerosol layer 2 by increasing m2 ni from zero to 6.2×10−4. Parameter Value Value 10◦N Latitude Name m2 od m2 pt (bar) m2 pb (bar) m2 r (µm) m2 nr α0 ×100 ch4rhc χ2 m2 ni = 0 m2 ni = 6.2×10−4 Difference 6.088 1.094 2.675 1.535 1.225 2.560 0.879 196.26 9.124 1.112 4.183 1.597 1.243 2.290 0.877 193.13 49.87% 1.64% 56.39% 4.01% 1.42% -10.55% -0.23% -1.59% reflected from the 26-bar level. From the previous discussion, we expect a reasonable physical model has some pressure value Pd for which the methane mixing ratio is independent of latitude for P > Pd, but allows a decline in mixing ratio with lat- itude for P < Pd. We assume that the highest mixing ratio we observe at low latitudes (which turns out to be at 10◦N) is representative of the deep mixing ratio and assume all of the variation with latitude is a depletion relative to that level. Here we describe the results of fit- ting two alternative vertically varying depletion models: the descended profile model described in Fig. 14A and Eq. 4, and the step function depletion described in Fig. 14B. Both options result in improved fit quality at high latitudes, with depletions confined to the upper tropo- sphere. We first consider the stepped depletion model shown in Fig. 14B because it is easier to constrain its bottom boundary at all latitudes. A more detailed look at the 60◦N spectrum from 2012 in comparison with a model fit using a vertically uniform methane mixing ratio is shown in Fig. 23A-C, while our best fit model for the stepped methane profile is displayed in Fig. 23D-F. Here we as- sume that the deep mixing ratio is equal to the 10◦N best fit uniform VMR value of 3.14%, and optimize the depleted mixing ratio α1 and the depth of the depletion Pd to minimize χ2. The result is seen to be a substan- tial improvement of the fit in the 750-nm region, with minor improvements in other areas, with an overall sig- nificant reduction in χ2 for the entire fit from 196.02 to 160.72. The fact that the difference plots show strong features in the vicinity of large slopes in the spectrum, particularly at 0.88 µm, suggests that there may be a slight error in the STIS wavelength scale. If we move the observed spectrum just 0.24 nm towards shorter wave- lengths, these χ2 values can be reduced to 170.02 and 137.93 respectively. (Although the STIS wavelengths are very accurate up to 653 nm because of the availability of numerous Fraunhofer calibration lines, longer wave- lengths require extrapolation that allows errors of this size.) The best fit values for the methane profile parame- ters are ch4vx = 0.73±0.08% and Pd = 3.0+3.5 −1.5 bars. The methane value is a little below the 0.93±0.08% for the uniform mixing ratio model, as expected. The methane relative humidity above the condensation level was found to be 95±16% for the uniform case and 67±32% for the upper tropospheric depletion case. The uncertainty in the depth of depletion (Pd) is much larger on the high side because the sensitivity to that parameter decreases with depth. We also tried fits with the descended depletion function described in Fig. 14A and Eq. 4, which is defined by a shape parameter vx and a depth parameter Pd. We found the depth parameter difficult to constrain because the rate of change of mixing ratio with depth can be come quite small for large depths due to the shape of the function. However, fixing Pd at 5 bars, and using just the 35 A B C D E F STIS OBS from 60oN 2012 UNIFORM METHANE MODEL CH4 VMR = 0.93 % for P > Pcond 0.75 χ2 = 196.02 0.80 Wavelength (µm) 0.85 STIS OBS from 60oN 2012 UPPER TROPOSPHERIC METHANE DEPLETION MODEL CH4 VMR = 0.73% from Pcond to 3 bars, CH4 VMR = 3.15% for P > 3 bars. 0.75 χ2 = 160.72 0.80 Wavelength (µm) 0.85 0.90 0.90 0.10 F / I s a e M - l e d o M 0.01 0.02 0.01 0.00 -0.01 2 1 0 -1 -2 c n U / f f i D 0.10 F / I s a e M - l e d o M 0.01 0.02 0.01 0.00 -0.01 2 1 0 -1 -2 c n U / f f i D Fig. 23.- Detailed comparison of 60◦N 2012 STIS observations with best-fit model spectra assuming vertically uniform methane VMR (A-C) and with model calculations assuming a step-function change in methane VMR (D-F), where observations are plotted as continuous curves and models as colored points, using red, green, and blue for µ values of 0.3, 0.5, and 0.7 respectively. Below each spectral plot are plots of model minus observation (B and E) and the same difference divided the expected uncertainty (C and F). Methane profiles are described in the legends. shape parameter to control the depletion, we obtained a χ2 value of 167.34 and a shape parameter of vx = 1.22±0.54, fitting the same 2012 60◦N observation as in Fig. 23. Thus, a descended depletion fit is also viable, and probably a more realistic vertical variation than the step function. The advantage of the step function is that both parameters can be fit without too much difficulty. 9.3.2. Latitude dependent fits with a stepped depletion of methane Here we describe the results of assuming a stepped de- pletion of methane, parameterized by one fixed parame- ter (α0, the deep methane VMR, which is set to 0.0315) and two adjustable parameters (Pd, the pressure at which the step occurs and α1, the decreased mixing ratio be- 36 tween that level and the condensation level (which is a function of the decreased methane VMR). In addition to fitting these two parameters, we fit the usual aerosol pa- rameters and the methane relative humidity above the condensation level, resulting in a net increase of one ad- justable parameter, for a new total of nine. The best-fit parameter values and their uncertainties are given at 10◦ latitude intervals for both 2012 and 2015 in Table 12. These are plotted versus latitude in the left column of Fig. 24 and comparisons of model and observed spectra are displayed in the right column. The best-fit methane depletion depth parameter values are shown by dashed lines in panel B of Fig. 24 for Pd and by dotted lines in panel D for α1. At high latitudes the latter is near 0.8%, and increases somewhat at low latitudes, but becomes very uncertain at low latitudes, which is a result of having less and less influence on the spectrum as the depth of the depletion decreases towards the condensation level. As shown in panel D, the deple- tion depth is in the 3-5 bar range from 70◦N down to about 50◦N, and then declines to nearly the condensa- tion level by 20◦N, and at low latitudes there is almost no depletion. The improvement in fit quality is significant at high latitudes. In comparison with the uniform methane fit results, we see only minor changes in most of the other parameters. The top pressure of the tropospheric cloud layer is nearly the same for both models, although the stepped depletion model results show a little more variability. The retrieved bottom pressure shows more significant changes. The new results show much closer agreement between years, but more change with respect to latitude, increasing from about 2.5 bars at low latitude to 3 bars at high latitude. The prior results showed no consistent trend with lati- tude, averaging about 2.7 bars. The optical depth for that layer shows about the same trend with latitude and the same increase at high latitudes between 2012 and 2015. The particle size generally remains between 0.2 and 0.4 µm for both models, but the descended model fits indicate that particles in the northern hemisphere are about 40% larger in 2015 than in 2012, while the uni- form model showed much less difference between years. All these particle size differences are within uncertain- ties, however. The relative humidity results for methane are roughly similar for the two model types, with higher, near saturation levels at low latitudes and a factor of two decline in the polar region. Both find the methane hu- midity depressed at the equator, with a slightly sharper decline seen in the descended profile results. The refractive index results differ a little. For the de- scended depletion model fits for 2012 and 2013 are in somewhat better agreement than for the uniform model, and do not show as much trending towards slightly higher values at high latitudes. 9.3.3. Latitude dependent fits with descended depletion of methane Because the descended depletion function approaches the deep mixing ratio on a tangent, it is hard to constrain the depth parameter for this model at most latitudes. Thus, from preliminary fits we found a Pd value that worked well at high latitudes (Pd = 5 bars) and kept that constant, while using just the shape parameter (vx) as the additional adjustable parameter in maximizing fit quality as a function of latitude. The results for best fit parameter values and uncertainties are given in Table 13 and plotted in Fig. 25. These two depletion model fits are compared in Fig. 26, with descended model fits in panel A and the stepped depletion model fits in panel B. The descended model fits yield slightly lower χ2 values, especially at 70◦ N, although even there the difference is smaller than the expected uncertainty of p2χ2, which is 22 in this case. Both models imply that the high latitude depletion is of limited depth, and both imply that the methane hu- midity above the 1 bar level is near saturation at low latitudes and decreases poleward. Not only can we ob- tain good fits with a shallow depletion of methane, they are preferred on the basis of fit quality. Not only does the high latitude fit near 745 nm improve significantly when the vertically varying depletion models are used, but the overall χ2 at high latitudes is also significantly improved, as illustrated in Fig. 27. This is especially ap- parent at the higher latitudes and in comparing averages over the 50◦ – 70◦ latitude range. Although the stepped depletion model is seen to yield slightly better χ2 val- ues than the descended depletion models, the difference is less than the expected uncertainty. The virtue of the stepped depletion model is that it can be well constrained at all latitudes, while the virtue of the descended deple- tion model is that it makes more sense physically. We were able to extend the latitude range of the descended model fits by fixing the value of the depth parameter Pd to 5.0 bars. We then found that both depletion models do not quite yield zero depletion at low latitudes, which one might interpret to mean that we should have cho- sen a slightly lower deep methane VMR value. However, the χ2 values for the vertically uniform values are just as Case 24: 2-layer Mie model small r2 fit to 730-900 nm, 2012, 2015 STIS obs. Case 24: 2-layer Mie model small r2 fit to 730-900 nm, 2012, 2015 STIS obs. Orton T(P) Orton T(P) A m2_pt m2_n m2_pb m2_pt m2_n m2_pb 0.3 0.2 F / I LAT = 10o, χ2 = 144.54, YEAR = 2015 F µ= 0.70 µ= 0.50 µ= 0.30 37 0.1 0.0 0.3 0.2 F / I 0.1 0.0 0.3 0.1 0.0 0.3 0.2 F / I 0.1 0.0 0.75 0.80 0.85 LAT = 60o, χ2 = 239.38, YEAR = 2015 µ= 0.70 µ= 0.50 µ= 0.30 0.75 0.80 0.85 LAT = 10o, χ2 = 192.40, YEAR = 2012 µ= 0.70 µ= 0.50 µ= 0.30 0.75 0.80 0.85 LAT = 60o, χ2 = 160.72, YEAR = 2012 µ= 0.70 µ= 0.50 µ= 0.30 0.90 G 0.90 H 0.90 I ) r a b ( P r o x e d n i l a e R 0.5 1.0 1.5 2.0 2.5 3.0 3.5 101 ) r a b ( p r o , h t p e d l a c i t p o , ) m µ ( s u d a R i 10-1 2.0 1.5 1.0 0.5 0.0 H R H C 4 r m v 4 H C d e t l e p e D 0.020 0.015 0.010 0.005 0.000 1.4 1.2 1.0 0.8 0.6 0.4 ) F N ( / 2 χ m2_r -20 C ch4rhc -20 D α1 STIS 2015 STIS 2012 -20 E Nfree= 309 Nfree= 309 -20 -20 0 20 B m2_od 100 pd 40 60 80 pdpd m2_od m2_rm2_r 0 0 20 40 60 80 0.2 F / I ch4rhc ch4rhc 20 40 60 80 α1 40 60 80 0 complx2 error model 20 0 Planetographic Latitude (deg) 20 40 60 80 0.75 0.80 Wavelength (µm) 0.85 0.90 Fig. 24.- Stepped depletion model of vertical methane distribution fit to STIS spectra from 2012 and 2015. Conservative cloud model and gas profile parameters for a Mie-scattering haze above a single diffuse Mie-scattering tropospheric layer, assuming a deep mixing ratio of α0 = 0.0315, and a methane profile characterized by a pressure depth parameter Pd and a depleted mixing ratio α1 (defined in Fig. 14) and constrained by spectral observations from 730 nm to 900 nm. The parameter values are in panels A-E, with red (open circle) points displaying results of fitting 2012 STIS observations and black (filled circle) points displaying the results of fitting 2015 observations. Sample comparisons between measured and large-particle model spectra are in panels F-I. Note the great improvement in the high latitude fits near 745 nm, compared to results given in Fig. 20. good or slightly better than the depleted models at low latitudes. 9.4. Wavelength dependence issues Although the best-fit parameters given in Table 7 pro- vide great spectral matches over the fitted range (730– 900 nm), they do not provide good matches over the entire range. As expected, and as illustrated in Fig. 28, the corresponding model spectra fit even worse over the rest of the wavelength range than the initial fit shown in Fig. 16. The problem with both the small-particle and large-particle models is that they do not produce a large enough I/F at the short wavelength side of the spectrum (from 0.54 µm to 0.68 µm) for the two largest zenith an- gle cosines, and produce too high an I/F in the deeply penetrating region near 0.94 µm for all three zenith an- gles. The problem is less extreme for the small-particle solution because it produces a larger increase in I/F at shorter wavelengths. One way to solve the short wavelength deficit prob- to abandon spherical particles and use a lem is wavelength-dependent phase function that provides in- creased backscatter at short wavelengths, which is the approach followed by KT2009, and one which we will return to in a later section. An alternative approach considered here is to use a wavelength-dependent imagi- nary index that is small at short wavelengths and larger at long wavelengths, an approach used by Irwin et al. (2015) to solve a similar problem in fitting near-IR spec- tra. The utility of this approach is that the increased optical depth required to compensate for the small ab- sorption at long wavelengths leads to a needed increase 38 Single tropospheric Mie layer fits to the 730-900 nm spectra as a function of latitude assuming stepped depletion of CH4 below the condensation level. TABLE 12 Lat. m1 od (◦) ×100 m2 od m2 pt (bar) m2 pb (bar) m2 r (µm ) m2 nr ch4rhc α1 (%) Pd (bar) χ2 YR -10 3.5±1.3 3.25±1.10 1.18±0.10 2.55±0.15 0.21±0.07 1.71±0.27 1.07±0.43 0.79±0.41 1.12±0.61 178.88 2015 0 3.1±2.0 3.20±1.14 1.19±0.08 2.53±0.17 0.27±0.06 1.62±0.19 0.36±0.31 0.77±0.67 1.13±0.37 132.16 2015 10 4.7±1.3 3.37±1.04 1.20±0.08 2.41±0.12 0.33±0.10 1.54±0.15 1.33±0.57 0.60±0.20 1.11±0.47 144.54 2015 20 3.8±1.3 2.38±0.78 1.21±0.10 2.52±0.17 0.30±0.08 1.67±0.20 1.34±0.51 1.08±0.53 1.12±0.01 164.33 2015 30 4.2±1.0 1.95±0.57 1.17±0.12 2.93±0.38 0.31±0.10 1.69±0.22 0.98±0.25 1.45±0.33 1.61±0.12 167.31 2015 40 3.3±1.0 1.75±0.42 1.02±0.06 3.25±0.28 0.30±0.08 1.71±0.19 0.92±0.26 1.18±0.13 2.76±1.49 194.88 2015 50 3.1±1.0 1.33±0.51 0.96±0.05 2.96±0.36 0.26±0.19 1.88±0.52 1.06±0.33 1.05±0.12 3.90±4.58 258.44 2015 60 0.8±3.8 1.74±0.35 1.03±0.03 3.06±0.37 0.28±0.07 1.76±0.19 0.43±0.28 0.85±0.08 3.44±1.87 239.38 2015 70 1.4±1.9 1.85±0.33 1.00±0.04 3.10±0.32 0.29±0.06 1.71±0.15 0.55±0.29 0.81±0.08 4.89±4.36 268.77 2015 -20 2.0±1.1 4.41±2.06 1.25±0.09 2.67±0.19 0.19±0.05 1.58±0.38 0.96±0.30 1.09±0.45 1.15±0.21 134.90 2012 -10 5.2±1.1 3.46±1.53 1.17±0.11 2.63±0.21 0.20±0.07 1.75±0.27 1.52±0.51 1.17±0.73 1.12±0.01 142.97 2012 0 5.1±1.5 3.10±1.04 1.22±0.09 2.48±0.13 0.25±0.07 1.67±0.22 0.66±0.35 0.67±0.27 1.13±0.15 147.11 2012 10 3.7±1.1 2.62±0.53 1.09±0.08 2.42±0.15 0.22±0.07 1.81±0.24 1.17±0.38 1.39±1.32 1.12±1.05 192.40 2012 20 3.1±1.1 3.23±0.92 1.14±0.09 2.53±0.20 0.21±0.06 1.78±0.24 1.12±0.40 1.29±0.82 1.12±0.01 197.25 2012 30 3.2±1.0 2.44±0.60 1.21±0.12 2.83±0.37 0.22±0.08 1.82±0.28 1.03±0.32 1.13±0.31 1.41±0.03 141.96 2012 40 4.8±1.1 3.24±1.36 1.33±0.16 2.71±0.33 0.23±0.08 1.62±0.25 1.27±0.42 0.72±0.12 1.67±0.14 234.70 2012 50 0.9±2.1 1.71±0.31 1.04±0.03 2.92±0.28 0.22±0.09 1.86±0.30 0.48±0.23 0.93±0.10 3.12±4.50 166.19 2012 60 3.4±1.4 1.21±0.28 1.03±0.06 2.99±0.27 0.22±0.09 1.86±0.33 0.67±0.32 0.73±0.08 3.02±2.78 160.72 2012 70 3.4±1.7 1.26±0.34 1.02±0.06 3.16±0.36 0.22±0.09 1.84±0.33 0.54±0.33 0.72±0.08 2.87±0.17 202.39 2012 NOTE: The optical depth is for a wavelength of 0.5 µm. These fits used 318 points of comparison and fit 8 parameters, for a nominal value of NF=310, for which the normalized χ2 /NF ranged from 0.426 to 0.87. in the I/F at short wavelengths where the absorption is absent. To follow up on this approach we added an adjustable imaginary index to cloud particles in the m2 Mie layer, and then optimized model parameters to fit both the 730–900 nm region and the 540-580 nm region simultaneously, as described in the following section. 9.4.1. Controlling wavelength dependence with particulate absorption The first example of controlling wavelength depen- dence over a larger spectral range is based on adjustment of particulate absorption. For this example, we assume two Mie scattering clouds, with the top layer (m1) lo- cated at an arbitrary pressure of 50 mbar and containing conservative particles with an assumed effective radius of 0.06 µm, and an adjustable optical depth. The top layer has a very small optical depth and its particle size is not very well constrained by our observations. We chose a somewhat arbitrarily value based on preliminary fitted values. The Rages et al. (1991) haze model estimates a particle size closer to 0.1 µm at 50 mbar. The other Mie layer is assumed to be composed of a non-conservative material, characterized by a refractive index of m2 nr + 0 × i for λ <700 nm and n = m2 nr + m2 ni × i for λ > 710 nm. The tropospheric Mie layers (m2) is charac- terized by three additional fitted parameters: pressure, particle size, and optical depth. We then simultaneously fit just two sub regions of the spectrum: the 540-580 nm region, where we assume the particles are conser- vative, and the 730–900 nm region, where we assume a locally wavelength-independent imaginary index that is adjusted to minimize χ2. We also allowed m2 nr to be adjustable. This process produced a best-fit value of (4.9±1.3)×10−3 for the imaginary index and 2.7±0.3% for the deep methane mixing ratio. However, this pro- Case 26: 2-layer Mie model small r2 fit to 730-900 nm, 2012, 2015 STIS obs. Case 26: 2-layer Mie model small r2 fit to 730-900 nm, 2012, 2015 STIS obs. Orton T(P), descended CH4 with CH4v0=3.15% and Pd=5 bars Orton T(P), descended CH4 with CH4v0=3.15% and Pd=5 bars 39 0.5 1.0 1.5 2.0 2.5 3.0 3.5 100 10-1 10-2 1.5 1.0 0.5 0.0 10 1 1.4 1.2 1.0 0.8 0.6 0.4 ) r a b ( e r u s s e r P h t p e d l a c i t p O r o , ) m µ ( s u d a R i H R H C 4 x v r e t e m a r a p e p a h S ) F N ( / 2 χ A m2_pt m2_n m2_pb -20 m2_od B m2_r m1_od -20 C ch4rhc D -20 vx STIS 2015 STIS 2012 -20 E Nfree= 310 Nfree= 310 -20 20 40 60 m2_pt m2_n m2_pb 80 m2_od m2_rm2_r m1_od m1_od 20 40 60 80 0.3 0.2 F / I 0.1 0.0 0.3 0.2 F / I 0.1 0.0 0.3 ch4rhc ch4rhc 0.2 F / I 20 40 60 80 0 0 0 vx 40 60 80 0 complx2 error model 20 0.1 0.0 0.3 0.2 F / I 0.1 0.0 LAT = 10o, χ2 = 153.03, YEAR = 2015 F µ= 0.70 µ= 0.50 µ= 0.30 0.75 0.80 0.85 LAT = 60o, χ2 = 242.69, YEAR = 2015 µ= 0.70 µ= 0.50 µ= 0.30 0.75 0.80 0.85 LAT = 10o, χ2 = 193.20, YEAR = 2012 µ= 0.70 µ= 0.50 µ= 0.30 0.75 0.80 0.85 LAT = 60o, χ2 = 168.19, YEAR = 2012 µ= 0.70 µ= 0.50 µ= 0.30 0.90 G 0.90 H 0.90 I 0 Planetographic Latitude (deg) 20 40 60 80 0.75 0.80 Wavelength (µm) 0.85 0.90 Fig. 25.- Descended depletion model of vertical methane distribution fit to STIS spectra from 2012 and 2015. Conservative cloud model and gas profile parameters for a Mie-scattering haze above a single diffuse Mie-scattering tropospheric layer, assuming a deep mixing ratio of 0.0315, and a methane profile characterized by a pressure depth parameter Pd and a shape parameter vx (defined in Eq. 4 and illustrated in Fig. 14) and constrained by spectral observations from 730 nm to 900 nm. The parameter values are in panels A-E, with red (open circle) points displaying results of fitting 2012 STIS observations and black (filled circle) points displaying the results of fitting 2015 observations. Sample comparisons between measured and large-particle model spectra are in panels F-I. Note the great improvement in the high latitude fits near 745 nm, compared to results given in Fig. 20. cess slightly degraded the fit in the 730-900 nm region. To better constrain the methane mixing ratio for the case with absorbing aerosols we adopted the imaginary index obtained from the dual fit, then refit the remaining pa- rameters using the 730-900 nm region for our spectral constraints, yielding the results given in Table 10. Ap- plying these parameters over the entire spectral range from 540 nm to 960 nm, we then obtained a much im- proved match to the observations, with a χ2 of 724.50. This was further improved to 586.32 by optimizing values of m2 pt (1.09±0.01 bar), m2 pb (3.35±0.13 bar), m1 od (0.030±0.002), m2 od (3.91±0.34), m2 r (0.30±0.02), m2 nr (1.69±0.04), and m2 nilw (0.0051±0.0003), af- ter adding an intermediate imaginary index of 0.0011 to the tropospheric aerosol particles in the spectral interval from 670 nm to 730 nm, yielding the fit displayed in Fig. 29. Although the fit is good, it is not known whether any plausible cloud material has this absorption characteris- tic. Complex hydrocarbons, such as tholins (Khare et al. 1993), absorb more at shorter wavelength and have de- clining absorption over the range where our example model shows increased absorption. Judging from frost re- flection spectra obtained by Lebofsky and Fegley (1976), H2S does not appear to exhibit such a trend either. Thus we have some motivation to consider other ways to gen- erate wavelength dependence. 9.4.2. Controlling λ dependence with optical depth variations Although Mie scattering calculations for spherical par- ticles produce wavelength dependent optical depth and scattering phase functions, if these do not yield needed 40 Two-cloud spherical particle fits as a function of latitude assuming descended depletion of CH4 as a function of latitude. TABLE 13 Lat. m1 od (◦) ×100 m2 od m2 pt (bar) m2 pb (bar) m2 r (µm ) m2 nr ch4rhc vx χ2 YR -10 3.3±1.0 4.06±2.15 1.06±0.09 2.94±0.23 0.22±0.09 1.59±0.32 0.96±0.28 16.90±31.60 181.84 2015 0 4.6±0.7 3.14±1.09 1.13±0.06 2.70±0.25 0.26±0.07 1.63±0.21 0.61±0.11 32.70±12.30 134.79 2015 10 4.2±1.1 3.12±1.02 1.07±0.08 2.76±0.22 0.30±0.08 1.61±0.19 1.00±0.32 19.10±30.30 153.03 2015 20 3.0±0.9 2.42±0.69 1.13±0.06 2.70±0.20 0.33±0.09 1.62±0.17 1.02±0.23 14.40±11.20 168.49 2015 30 4.8±1.1 1.99±0.54 1.17±0.07 2.89±0.23 0.33±0.09 1.64±0.17 1.19±0.37 3.81± 1.38 163.38 2015 40 4.2±1.9 1.83±0.48 1.18±0.08 2.86±0.17 0.28±0.08 1.74±0.21 1.39±0.93 1.98± 0.76 192.65 2015 50 3.4±2.4 1.47±0.42 1.21±0.07 2.58±0.11 0.26±0.12 1.87±0.38 1.30±1.09 1.70± 0.74 256.78 2015 60 2.7±2.8 1.75±0.24 1.28±0.07 2.60±0.09 0.21±0.07 1.95±0.27 1.04±1.14 1.36± 0.61 242.69 2015 70 2.5±2.7 1.90±0.22 1.33±0.08 2.66±0.09 0.18±0.04 2.05±0.19 1.04±1.18 1.23± 0.54 263.79 2015 -20 2.1±0.9 3.73±1.77 1.11±0.08 2.97±0.23 0.25±0.09 1.58±0.28 0.95±0.26 15.70±29.40 138.18 2012 -10 4.4±0.6 3.82±1.64 1.08±0.03 2.86±0.30 0.25±0.09 1.61±0.28 1.19±0.17 30.70±28.60 145.93 2012 0 4.8±0.7 2.89±0.87 1.15±0.05 2.62±0.23 0.26±0.07 1.68±0.21 0.58±0.10 34.30± 9.49 148.60 2012 10 3.6±0.7 2.61±0.71 1.04±0.05 2.51±0.18 0.25±0.08 1.73±0.24 1.13±0.23 35.40±23.00 193.20 2012 20 2.7±0.9 3.62±1.30 1.07±0.06 2.70±0.18 0.27±0.08 1.60±0.21 0.98±0.23 17.00±12.60 197.04 2012 30 3.2±1.0 2.44±0.75 1.11±0.06 2.89±0.18 0.27±0.07 1.69±0.21 1.06±0.31 4.29± 1.45 142.42 2012 40 3.9±1.4 2.35±0.74 1.20±0.07 2.55±0.13 0.27±0.06 1.65±0.19 1.04±0.50 2.21± 0.61 245.82 2012 50 2.2±2.1 2.06±0.35 1.28±0.07 2.49±0.10 0.20±0.06 1.85±0.22 0.84±0.61 1.67± 0.58 170.15 2012 60 4.3±1.9 1.54±0.52 1.38±0.11 2.36±0.10 0.15±0.04 2.05±0.16 1.19±1.05 1.21± 0.41 168.19 2012 70 3.8±1.9 1.65±0.47 1.41±0.15 2.47±0.10 0.18±0.07 1.88±0.24 0.83±0.73 1.31± 0.44 206.35 2012 NOTE: The optical depth is for a wavelength of 0.5 µm. These fits used a fixed value of P d = 5 bars, α0 = 3.15%. There were 318 points of comparison and 8 fitted parameters, for a nominal value of NF=310, for which the normalized χ2 /NF ranged from 0.42 to 0.83. dependencies, and if particle size is constrained, and wavelength-dependent absorption is not acceptable, then non-spherical particles need to be considered. The simplest option is to use a single HG scattering phase function and simply adjust the wavelength depen- dence of the optical depth to match the observed spectral variation. An increase in optical depth with size param- eter (2πr/λ) is certainly a characteristic shared by most particles and by aggregates in our trial calculations. It also is plausible that a non-spherical particle might ex- hibit a greater λ dependence in optical depth than a spherical particle for the case in which both particles satisfy the other constraints in the 730-900 nm region. To define the needed τ (λ) function we began by tak- ing our best fit vertical structure and asymmetry fit for the 730-900 nm region, then computed a series of model spectra with optical depths increasing until we could find an optical depth at any wavelength that would match the observed I/F at that wavelength. But we found a problem with this approach. At short wavelengths, the optical depth needed to match two successive continuum regions (e.g. at 560 nm and 585 nm) was about 4-5 times the value at 800 nm that was derived from fits to the 730- 900 nm region. But to match the intervening absorption feature at 576 nm would require about half of that opti- cal depth. Thus a smoothly varying optical depth func- tion could not be created in this fashion, and a function that included wiggles at all the methane features was completely implausible. The fix to this problem was to distribute the cloud particles over a greater atmospheric depth. This would not change the continuum I/F values very much, but in the weakly absorbing regions, there would be more absorption. At longer wavelengths this required an increase in the cloud's optical depth, which 10-1 10-1 C H 4 Saturation Profile 100 100 Pcond= 1.13 bars DESCENDED PROFILE MODELS ) ) r r a a b b ( ( e e r r u u s s s s e e r r P P c c i i r r e e h h p p s s o o m m A A t t 41 10-1 10-1 100 100 ) ) r r a a b b ( ( e e r r u u s s s s e e r r P P c c i i r r e e h h p p s s o o m m A A t t C H 4 Saturation Profile STEPPED DEPLETION MODELS A 101 101 10-5 10-5 10oN χ2 = 173.11 30oN χ2 = 152.90 50oN χ2 = 213.46 70oN χ2 = 235.07 10-4 10-4 10-3 10-3 Methane Volume Mixing Ratio Methane Volume Mixing Ratio 10-2 10-2 10-1 10-1 B 101 101 10-5 10-5 10oN χ2 = 168.47 30oN χ2 = 154.64 50oN χ2 = 212.32 70oN χ2 = 235.58 10-4 10-4 10-3 10-3 Methane Volume Mixing Ratio Methane Volume Mixing Ratio 10-2 10-2 10-1 10-1 Fig. 26.- A: Best-fit descended profiles at 4 latitudes, using average parameter fits for 2012 and 2015. B: Best fit stepped depletion profiles, using average parameter fits for 2012 and 2015. The profiles in A are overlain in light gray in B for reference. Both sets of fits show decreasing methane humidity with latitude above the 1 bar level, and both indicate that the depletion is of limited depth (∼5 bars or less). In both cases average χ2 values for 2012 and 2015 are given in the legend. 300 250 2 χ 200 2012 Vertically uniform CH4 Descended depletion Stepped depletion 300 250 2 χ 200 2015 Vertically uniform CH4 Descended depletion Stepped depletion 150 100 -20 0 20 40 60 Planetographic Latitude (deg) . g v A o 0 7 - o 0 5 80 150 100 -20 0 20 40 60 Planetographic Latitude (deg) . g v A o 0 7 - o 0 5 80 Fig. 27.- Comparison of 2012 (left) and 2015 (right) χ2 versus latitude values for for three different methane vertical distribution models: uniform (solid line), descended depletion (dotted line), and stepped depletion (dashed line). Corresponding averages over the 50◦ – 70◦ latitude range are also shown near 80◦N in each panel. The depleted profile values are slightly shifted in latitude to avoid error bar overlaps. This shows that the overall fit quality is improved by use of the descended profile, in addition to the more obvious improvement near 750 nm. The small overall improvement seen in fits to the 2015 observations is likely due to the increased noise level at high latitudes and low signal levels for this data set. in turn required readjustment of the optical depth ratio between 800 nm and 540 nm. The result of this process applied to our model of the 2015 STIS spectrum at 10◦N is shown in Fig. 30. tion can be written as f1(λ) = a − b × sinα[ π 2 (λo − λ)/(λ0 − λ1)], (6) λ1 ≤ λ ≤ λo 9.4.3. Controlling λ dependence with phase function variations KT2009 assumed that the main tropospheric cloud layer had a wavelength-independent optical depth, which is a plausible assumption for large particles, and used a wavelength dependent phase function to match the ob- served spectral variation. A general form of their func- in which KT2009 assumed α = 4, a = 0.94, b = 0.427, λ0 = 1 µm, and λ1 = 0.3 µm, which makes f1 reach a maximum of 0.94 at a wavelength of 1 µm and a mini- mum of 0.513 at 0.3 µm. They applied this to a double HG function with adopted values of g1 = 0.7 and g2 = -0.3. (Note that there is no basis for applying this func- tion to wavelengths greater than 1 µm or less than 0.3 µm.) We found that this function was able to fit low 42 Params_from_Mie_conservative_730-900nm_fit_Dec10-191250-2017.tab Planetographic Latitude = 10o N plot_spec_anal.pro: CASENUM = 2 0.10 F / I χ2 = 1732.43 (for total range) Measured spectrum displayed as solid lines Best fit model displayed as colored points: µ= 0.3 µ= 0.5 µ= 0.7 . s a e M / l e d o M . t r e c n U 0.01 1.2 1.1 1.0 0.9 0.8 0.7 2 0 -2 -4 -6 -8 0.5 / ) . s a e M - l e d o M ( 0.6 0.6 0.6 0.7 0.7 0.8 0.8 fitted range 0.7 Wavelength (µm) 0.8 0.9 0.9 0.9 1.0 Fig. 28.- As in Fig. 16, except 10◦ N spectral comparisons are shown for a conservative cloud model that provide the best match to the 10◦ N 2015 STIS spectra from 730 nm to 850 nm, using the small particle solution. Note the significant model falloff at shorter wavelengths. See text for implications. stis_fitctl_spec_glats-30.00to87.00year2015mulim0.175Jun10-155158-2016limbcorr2.88nm.unf_lmfit_May14-184627-2018.tab plot_spec_comparisons.pro CASENUM = 10 0.10 F / I STIS obs. from 10oN 2015 s a e M / l e d o M 0.01 1.2 1.1 1.0 0.9 0.8 0.7 4 2 0 -2 -4 -6 0.5 c n U / f f i D χ2 = 586.32 χ2 = 159.10 (0.54 - 0.67 µm) 0.6 0.7 Wavelength (µm) 0.8 0.9 1.0 Fig. 29.- Extended range wavelength dependent model, using imaginary index variations to adjust the wavelength dependence. The imaginary index m2 ni, multiplied by a factor of 100, is shown by the dashed curve. latitude spectra over the 730 nm to 900 nm range quite well, but that the a and b constants needed to vary with latitude and that we needed to increase g1, leading us to adopt a new value of 0.8. When applied to the extended spectral range, we needed to increase λ1 to about 0.45 and α to 5. The resulting spectral match was intermedi- ate between those shown in Figs. 29 and 30. A problem with this formulation is that extending the idea to longer wavelengths would require the particles to become more and more forward scattering at longer wavelengths (in stis_fitctl_spec_glats-30.00to87.00year2015mulim0.175Jun10-155158-2016limbcorr2.88nm.unf_lmfit_Dec14-111419-2017.tab plot_spec_comparisons.pro CASENUM = 8 43 0.10 F / I STIS OBS from 10oN 2015 χ2 = 889.49 s a e M / l e d o M 0.01 1.2 1.1 1.0 0.9 0.8 0.7 4 2 0 -2 -4 -6 0.5 c n U / f f i D 0.6 0.7 Wavelength (µm) 0.8 0.9 1.0 Fig. 30.- Extended range wavelength dependent model for HG particles, using an optical depth variation with wavelength adjusted to match the 2015 STIS observations at 10◦N. The dashed curve displays the wavelength dependent optical depth normalized by its value at 800 nm, then scaled downward by a factor of 10. order to produce the same effect that absorbing Mie par- ticles produce, as discussed later). This is not a plausible trend. For large enough wavelengths the particles must become less forward scattering. We also considered whether the HG model could use a wavelength-dependent asymmetry parameter instead of a wavelength dependent optical depth to match the observed spectrum over a wider spectral range. How- ever, matching the shorter wavelengths required a nega- tive asymmetry parameter, which is an implausible con- dition, and thus not an acceptable solution. Thus over the longer spectral ranges it is most likely that optical depth variation and possible particulate ab- sorption will be needed to model reflected spectra. Phase function variations will also be present, but cannot be the sole way to produce the needed wavelength dependence in scattering properties. 9.5. Two-layer Mie model applied to Near-IR spectra To test whether our 2-layer Mie models would be capa- ble of fitting near-IR spectra, we extended model calcula- tions to 1.6 µm and compared them to a central meridian SpeX spectrum covering the 0.8–1.65 µm range. [We ob- tained this spectrum from the Infrared Telescope Facility on 18 August 2013, using the cross-dispersed mode of the SpeX spectrometer. The spectrum was spatially aver- aged over the central 0.4 arcseconds of the central merid- ian covered by the 0.15-arcsecond slit, corresponding to an average latitude of 24◦N. It was spectrally smoothed to the same spectral resolution as the smoothed STIS spectrum (a FWHM of 2.88 nm). The spectrum was scaled to match the 1.09×10−2 I/F center-of-disk H-band I/F from Sromovsky and Fry (2007).] The initial small- particle model parameters we used were from the 20◦N spectrum and used an imaginary index of 0.0046 at all wavelengths longer than 730 nm. For the initial large- particle model, we used a fit to the 10◦N spectrum and used an imaginary index of 6.2×10−4 for λ > 730 nm. The smaller index for the large particle solution is a result of the lower real refractive index for the best-fit larger particles. Fig. 31 shows that the extended models agree well in the dark regions of the spectrum, indicating that little change in stratospheric haze properties is needed, but is far too bright in the longer wavelength continuum re- gions, indicating that the real cloud particles have, at longer wavelengths, a lower optical depth or greater ab- sorption than the model particles. The large particle solution is the worst offender because its scattering effi- ciency is a relatively weak function of wavelength, while the scattering efficiency of the smaller particles declines substantially, though not enough to match the falloff in 44 pseudo continuum I/F values with wavelength. The excess model I/F at these wavelengths can be re- duced by increasing the imaginary index as indicated in the bottom panel of Fig. 31. Our procedure for devel- oping these solutions was to start with a conservative solution constrained by the 730-900 nm spectrum. We then used that as an initial guess for a split fit of the 540-580 nm plus 730-900 nm region, assuming that the imaginary index was zero for λ ≤ 580 nm and had an adjustable value of m2 nilw for for λ ≥ 730 nm. From that we obtained an estimate for the imaginary index in the 730-900 nm region. We then fixed that imaginary index and did a new fit within the 730-900 nm region to get a revised estimate of the methane profile. We then fixed the methane profile and used a second split fit to improve the optical depth and vertical aerosol distribu- tions, as well as particle size and real index. We then adjusted the imaginary index in the 670-730 nm range to optimize the fit to that part of the spectrum. That provided the parameters used for the initial near-IR cal- culations. To match the near-IR spectrum we did a suite of forward calculations with different constant imaginary index values to find in each wavelength region the imag- inary index that provided the best model match to the observations. This was not done at a fine wavelength resolution in an attempt to match every detail because the solid materials making up the cloud particles would not likely have such fine-scale absorption features. This figure shows that the STIS-based model with two layers of small spherical particles can match the observed infrared spectrum out to 1.65 µm by increasing the imag- inary index with wavelength as shown in Fig. 31, reach- ing a maximum of 0.1 for the H-band region. Our in- dex is generally larger than the imaginary index esti- mated by Irwin et al. (2015) although of roughly sim- ilar shape. Our mean value in the H band is similar to the adopted value of de Kleer et al. (2015). Our fit- ted real index of 1.72±0.2 for the extended-wavelength small-particle model is significantly larger than the value of 1.4 assumed by Irwin et al. (2015). Since our par- ticles are thus inherently brighter, it is not surprising that we might need more absorption than Irwin et al. to match the observations. Our large-particle model, with a lower real index, has an imaginary index profile of similar shape but lower amplitude. The imaginary index value for de Kleer et al. (2015) we derived from their assumed single-scattering albedo of 0.75, which corresponds to an imaginary index of 0.06 for 1-µm particles. Irwin et al. suggested that the refractive index spectrum would allow us to determine the composition of the cloud particles. However, the most likely cloud material (H2S) does not have well characterized (quantitative) absorption proper- ties, and frost reflection spectra between 1.2 and 1.6 µm (Fink and Sill 1982) provide little qualitative evidence for significant absorption of the type we seem to need to match the observed spectrum. We could also have modeled the drop in I/F at longer wavelengths using a HG particle scattering model, either by varying the single-scattering albedo with wavelength, or by varying the optical depth as a function of wave- length. It is left for future work to evaluate which sort of variation provides the best overall compatibility with the observations. Although our modified imaginary index allows our two cloud model to closely reproduce the observed spectrum in most regions, there are some problems that need fur- ther work to address. First, note that at the 1.08-µm con- tinuum peak, the model contains modulations that are not observed in the measured spectrum. This is also the case for model calculations shown by Tice et al. (2013), and is an indication of a possible flaw in our commonly used absorption coefficients in this region. There is also a relatively sharp feature at 1.1 µm that is much larger in the model than in the observations. Fur- ther, the detailed shape of the pseudo continuum peak near 1.27 µm is not fit very well. 10. DISCUSSION 10.1. Why occultation constrained fits produced larger methane VMR values Given the previous discussion of methane depletion profiles, this might be a good point at which to com- pare the methane profiles in Fig. 26 with those obtained from the occultation analysis of Lindal et al. (1987) or Sromovsky et al. (2014). This is provided in Fig. 32, where we also show the results of Orton et al. (2014b) and Lellouch et al. (2015). The main regions of sen- sitivity to the methane VMR values are indicated by thicker lines for our current STIS results and those of Lellouch et al. (2015). Note that our current STIS re- sults at 30◦ N are in very good agreement with the Lel- louch et al. results where they have overlapping sen- sitivity (roughly the 200–700 mbar range). Both have relatively high methane relative humidities compared to the saturation vapor pressure profile computed for the Orton et al. (2014a) thermal profile. The occultation re- sults for methane are at much lower levels at pressures less than the putative methane condensation pressure A 0.100 45 plot_spec_anal.pro: CASENUM = 10 Planetographic Latitude = 24o N 0.010 F / I 0.001 . s a e M / l e d o M 5 4 3 2 1 0 x e d n i y r a n g a m i I 0.100 0.010 0.001 0.8 B 0.8 C This work (small r) This work (large r) 2013 SpeX spectrum STIS 2012 models (+ = large particle, line = small particle) STIS 2012 models extended to near-ir by adjusting ni(λ) as in C 1.0 1.2 1.4 1.6 1.0 1.2 1.4 1.6 de Kleer et al. (2015) ) 5 1 0 2 I r w i n e t a l. ( 0.8 1.0 1.2 Wavelength (µm) 1.4 1.6 Fig. 31.- A: Our 2013 SpeX near-IR spectrum of Uranus from latitude 24◦ N (black) compared to model spectra for the same observing geometry but using the gas and aerosol parameters from the 20◦ N two-layer Mie scattering model for two particle size solutions: small- particle (line only) and large-particle (lines with points). The same models, extended to the near-IR by adjusting the imaginary index of the cloud particles, are shown in red. In the first model set of models, a vertically uniform methane mixing ratio of 2.65% was assumed up to the methane condensation level. In the second set (red curves) a deep mixing ratio of 3.15% was assumed and a descended depletion profile shape was used, with vx = 7.34 an Pd = 5 bars. B: ratio of model spectra to the SpeX observed spectrum. C: Imaginary index spectra assumed in the second set of models (red), compared to imaginary index values derived by Irwin et al. (2015) and a value inferred from a single scattering albedo used by de Kleer et al. (2015). (about 1.2 bars). In the occultation analysis, temper- ature and methane profiles are linked. Both tempera- ture and composition affect density, which in turn af- fect refractivity versus altitude, which is the main result produced from the radio measurements. The refractiv- ity profile can be matched by a family of thermal and corresponding methane profiles. A hotter atmosphere is less dense, and thus allows more methane to produce the same refractivity. Because the occultation profiles have such low relative methane humidities above the cloud level compared to what the STIS spectra require to ob- tain the best fits, the hottest occultation profile is fa- vored. If the only allowed adjustment of methane is se- lection of the optimum occultation profile, as was the case for our previous analyses (Sromovsky et al. 2011, 2014), then we obtain a deep mixing ratio that is rela- tively high (4%) so that the methane mixing ratio near and above the cloud level can approach closer to the level needed to provide the best spectral match. As an exam- ple of this behavior, we carried out fits of STIS spectra at 10◦ N, using STIS spectral fit quality as the only con- straint, and compared that to the best fits obtained for profiles with fixed occultation consistent methane verti- cal profiles. The results, tabulated in the legend of Fig. 32, show that all the occultation fits are much worse than the STIS-only constrained fits, and that the best of the occultation constrained fits (for the F profile) is for the hottest profile, which provides the most upper tropo- 46 spheric methane, even though that has a deep methane VMR that is much higher than is needed if one does not force the methane to fit an occultation profile. Just be- low the cloud level, the methane VMR at low latitudes is closer to 2% at least in the region above the lower tropospheric clouds (near 2.5 bars) and perhaps deeper, although the STIS spectra are not sensitive to values deeper than that. 10-1 ) r a b ( e r u s s e r P c i r e h p s o m A t 100 Occultation profiles F1 D1 F D 10-4 10-5 Lellouch et al. (2015) STIS fit at 30oN Orton et al. (2014) χ2 ratio to STIS fit 3.42 9.42 1.72 5.51 10-3 Methane Volume Mixing Ratio 10-2 10-1 Fig. 32.- Comparison of methane profiles derived from STIS- constrained and other spectral observations by Lellouch et al. (2015) and Orton et al. (2014b) compared to those derived from oc- cultation observations by Lindal et al. (1987) and Sromovsky et al. (2011). The 4% deep methane VMR occultation profiles provide better agreement with STIS-constrained results in the upper tro- posphere. But without occultation constraints, the preferred deep mixing ratio is closer to 3% for most aerosol models fit to the 730- 900 nm spectrum. 10.2. Evidence for a deep cloud layer In our previous paper dealing with earlier STIS obser- vations (Sromovsky et al. 2014), we found that the fit quality at short wavelengths was improved by adding a deep cloud layer, which we fixed at the 5 bar level and as- sumed had the same tropospheric scattering parameters as KT2009. The only adjustable parameter for that layer was its wavelength-independent optical depth, which we found to vary from about 4 at low latitudes to about half that at high latitudes. It is possible in our current mod- eling that the more extended vertical extent of our upper tropospheric cloud layer serves to reduce the need for the contribution of a deeper cloud. The main function of the deeper cloud is to improve the fit in the 540 to 600 nm range where matching weak methane band depth is eas- ier if some of the aerosol scattering is moved to higher pressures. To provide a better test of the existence of a deeper cloud, we looked at spectra with deeper penetration. Choosing a spectrum with a nearly vertical view (µ = 0.9 at 10◦N), we computed simultaneous model spectra for view angle cosines of µ = 0.3, 0.5, and 0.9, based on the fit we obtained using the standard set of view angle cosines (µ = 0.3, 0.5, and 0.7). That model did not fit the weak methane bands very well even with the standard view angles and was even worse for this more deeply penetrating set. The χ2 values rose from 586.32 to 714.59, with an expected χ2 uncertainty of 35-40. This χ2 increase by 128.3 is about three times its uncertainty. However, by refitting the same model (still without a deep cloud) to the new set of view angles, we reduced the χ2 value for this new set of angles to 705.84, and thus reducing the difference to 119.5, which is still about three times the expected uncertainty in χ2. After in- serting an optically thick deep cloud with an adjustable pressure, a new fit further reduced χ2 from 705.84 to 645.62, a decrease of 59.52, which is about 1.6 times its uncertainty. A comparison of the latter and initial fits to the measured spectra is displayed in Fig. 33. The χ2 improvement is even more dramatic when computed just for the region from 540 nm through 670 nm. In that case the χ2 change is from 180.53 to 125.17, a decrease of 75.46, which is over three times the expected uncer- tainty of about 22 for this more limited range that has 243 comparison points. The model with a deep cloud also improved fits at the original set of view (and zenith) angles. Adding that layer to the model plotted in Fig. 29 and refitting, decreased χ2 from 586.32 to 529.95, a decrease of 56.37, with most of this change taking place in the 540-670 nm region where χ2 dropped from 159.10 to 102.47, a decrease by 56.63, which is about 2.6 times the expected uncertainty. Thus both sets of view angles lead to significant local fit improvements, with derived effective pressures of 10.6±0.4 bars for the more deeply penetrating view angles and 9.5±0.5 bars for our stan- dard set. A better estimate for the effective pressure of an optically thick deep cloud is probably 10±0.5 bars. A lower pressure is likely if the cloud is not optically thick. When we fixed the deep cloud pressure at 5 bars the best-fit optical depth of the cloud was 4.2±0.7 (us- ing the more deeply penetrating spectral constraints). This is quite consistent with the optical depth of the 5- bar deep cloud fits of Sromovsky et al. (2014). Further investigation of the nature of this deep cloud layer, in- 0.8 stis_fitctl_spec_glats-30.00to87.00year2015mulim0.175Jun10-155158-2016limbcorr2.88nm.unf_lmfit_May10-163352-2018.tab stis_fitctl_spec_glats-30.00to87.00year2015mulim0.175Jun10-155158-2016limbcorr2.88nm.unf_lmfit_May10-174307-2018.tab plot_spec_comparisons.pro CASENUM = 16 STIS obs. from 10oN 2015 (shaded) Model w/o deep cloud χ2 = 705.84 (0.54 - 0.96 µm) χ2 = 180.53 (0.54 - 0.67 µm) Model with deep cloud χ2 = 645.62 (0.54 - 0.96 µm) χ2 = 125.17 (0.54 - 0.67 µm) 0.6 F / I 0.4 0.2 µ = 0.3, 0.5, 0.9 0.55 0.60 Wavelength (µm) 0.65 0.70 Fig. 33.- STIS 2015 observations at 10◦N (green shading indi- cating uncertainties), compared to fitted model results without a deep cloud layer (red) and with a deep cloud layer (blue). These are for non-standard, more deeply penetrating, zenith angle cosines of 0.3, 0.5, and 0.7, with largest cosines corresponding to largest I/F values at continuum wavelengths. The legend gives χ2 values for the entire spectral range that was fitted (0.54 - 0.96 µm) and for the region most influenced by the deep cloud (0.54 - 0.70 µm). cluding its latitude dependence, is left for future work. A plausible composition for such a cloud is NH4SH. 10.3. Comparison with other models of gas and aerosol structure on Uranus. Models of 0.8-1.8 µm SpeX spectra of Uranus by Tice et al. (2013) and more recently by Irwin et al. (2015) and recent models of H-band (1.47-1.8 µm) spec- tra by de Kleer et al. (2015) present what appear to be different views of the cloud structure from that derived from our STIS observations. Some fraction of the dif- ferences are due to different constraining assumptions. The other authors typically constrain the upper cloud boundary pressure and fit the scale height ratio, while we have here mainly assumed a unit scale height ratio (particles uniformly mixed with gas) and treated the up- per boundary pressure as adjustable. The differences are probably not due to very different conditions on Uranus, as the spectral observations are generally very similar, as illustrated in Fig. 34. These spectra are all obtained near the center of the disk, and all near latitude 20◦ N. In most of the spectral range they are all within 10% of each other. The main exception is the de Kleer et al. (2015) spectrum, which is much brighter than the other two spectra in the 1.63-1.8 µm region. This would pre- sumably lead to a model with much greater stratospheric haze contributions than would be needed to match the other spectra. To better characterize these differences and better understand their origin, we attempted to re- 47 produce results from these near-IR analyses. The first attempt was to match the Irwin et al. (2015) retrieval of a two-cloud structure from the Tice et al. (2013) 2009 SpeX central meridian data (0.8-1.8 µm range). They used 1.6% deep CH4 with 30% RH above condensation level and the Lindal et al. (1987) Model D T(P) profile. They retrieved self-consistent refractive indexes for their Tropospheric Cloud (TC) and Tropo- spheric Haze (TH), similar to Tice et al. 2-cloud model. They retrieved particle sizes, but used "combined H-G" phase function fits in the forward modeling, rather than Mie calculations. An additional complication was that their plots of optical depth vs pressure were incorrect in the paper (estimated to be about an order of mag- nitude too large, P.G.J. Irwin private communication). That and the uncertain way double HG phase functions were obtained from the Mie phase functions, led us to not attempt detailed quantitative comparisons. We decided to make our quantitative comparisons with 2-cloud results of Tice et al. (2013). This was more tractable, as the phase functions for both TC (tropo- spheric cloud) and UH (Upper Haze, called TH in Ir- win et al.) were simply H-G phase functions with an assumed asymmetry parameter g = 0.7. They also utilized wavelength-dependent optical depths based on Mie calculations of extinction efficiency, but they did not use the wavelength-dependent phase functions or wavelength-dependent asymmetry parameters for either particle mode. Although, for the larger particles in the TC, the asymmetry parameter obtained from Mie calcu- lations is close to their chosen value, the 0.1-µm parti- cle model has a very small asymmetry, which leads to a backscatter phase function value about ten times that for an HG function with g = 0.7. Since their UH (or TH) particles have such small optical depths, their con- tribution can be well approximated by single-scattering, in which case the observed I/F contribution is given by I/F = 1 4 P (θ)τ /µ (7) where θ is the scattering angle (about 180◦ in this case), τ is the vertical optical depth, and µ is the cosine of the ob- server zenith angle. This makes the modeled I/F strongly dependent on the assumed phase function, specifically its backscatter amplitude. While there is substantial varia- tion in scattering efficiency with wavelength for a 0.1-µm particle, such a particle would not have such a strongly forward peaked phase function, and would probably re- quire roughly a factor of ten lower optical depth than Tice et al. (2013) found for their UH layer. However, 48 0.1000 0.0100 F / I 0.0010 0.0001 5 X e p S 3 1 0 2 / m u r t c e p S 4 3 2 1 0 plot_spec_anal.pro: CASENUM = 7 2013 SpeX 20oN spectrum 2009 SpeX 8oN from Tice et al. (2015) 2010 OSIRIS 10oS-10oN spectrum from de Kleer et al. (2015) 0.8 1.0 1.2 1.4 1.6 1.8 0.8 1.0 1.2 Wavelength (µm) 1.4 1.6 1.8 Fig. 34.- Comparison of near-IR spectra of Uranus. Our 2013 central-disk spectrum is shown in black. The 2009 SpeX central-disk spectrum of Tice et al. (2013) is shown in blue, and the 2010 OSIRIS 10◦S - 10◦ N spectrum of de Kleer et al. (2015) in red. The bottom panel plots the ratio of each spectrum to our 2013 SpeX spectrum. using their peculiar scattering characterization for this layer, and using their more plausible characterization for the lower layer, and their chosen single-scattering albedos, we were able to roughly match our own 2013 SpeX center-of-disk spectra (which are quite similar to the spectra shown in the Tice paper). Thus we have two different vertical structures that can match the spectra. Ours has a single tropospheric layer uniformly mixed be- tween 1.06 and 3.3 bars (small particle solution), while theirs has a very strongly varying optical depth per bar between their assumed cloud top of 1 bar and their fit- ted bottom at 2.3 bars. We did not attempt to reproduce the more complex structures based on Sromovsky et al. (2011) three- and four-cloud models. We also tried to reproduce de Kleer et al. (2015) re- sults for a 2-cloud model. Their retrievals were for a more limited H-band wavelength range (H-band spec- tra). Their spectra were also similar to our 2013 SpeX re- sults, except their dark regions were as much as 3-4 times brighter (see Fig. 34). They used a two-stream radiative transfer model, with wavelength dependent H-G param- eters based on Mie calculations. Using their retrieved optical depths, we roughly matched their window I/F. However, we used correlated-k coefficients for Hartmann type line-shape wings, while de Kleer et al. used the hybrid wing shape from Sromovsky et al. (2012a) that produces more absorption in the H-band window. Using these c-k coefficients, our I/F values in the methane win- dow were lower than those of de Kleer et al. by a factor of 2 or so. The origin of these differences remain to be determined. It is likely that it is not entirely a result of very different numbers of streams, as de Kleer et al. did trial calculations showing that their approximation was good to within ∼10%. A comparison of the characteristics of the tropospheric cloud models from aforementioned references is displayed in Fig. 35. Although all of these models provide good fits to the spectra (ignoring the fact that we could not repro- duce all these results), they have very different vertical structures and total optical depths and column masses. In fact the widest variation in total cloud mass is be- tween our own small-particle and large-particle solutions. In the log-log plot in Fig. 35A, the various cloud struc- tures seem more similar than in the linear plot in panel B, where the huge differences in optical depths and mass loading are more accurately conveyed. The column num- ber density in particles per unit area is computed as n = τ /(πr2Qext), where r is the particle radius and Qext is the extinction efficiency (extinction cross section divided by geometric cross section). From n the mass loading (mass per unit area) is computed as m = nρπr3, assuming that the particle density ρ is 1 g/cm3. Our small-particle tropospheric cloud is one of very low main- tenance. It needs very little material to form, the parti- cles fall slowly because they are small, and thus probably a low level of mixing is needed to sustain it. It also has the virtue of having a refractive index similar to that of its potential main component, H2S. The large particle cloud is thirty times more massive, with larger particles that fall much more quickly, needing much more vertical transport to be sustained. If these clouds are to be made of H2S, it is worth con- sidering whether there is enough H2S available to make them. For a mixing ratio αH2S, the mass per unit area of H2S between two pressures separated by ∆P would be (MH2S/M)αH2S∆P/g, where g is gravity (9.748 m/s2), and the ratio of molecular weights of H2S to the total is given by 34/2.3 = 14.78. For H2S to condense at the tropospheric (layer-2) cloud base its mixing ratio must have a minimum value that depends on base pressure as shown in Fig. 36. To condense at the 3.3 bar level would require the H2S VMR to be equal to its the solar mixing ratio of 3.1×10−5 (Lodders 2003). About 10 times that VMR would be needed to condense as deep as the 5 bar level and about ten times less would lead to condensa- tion no deeper than the 2.4 bars. Microwave observations by de Pater et al. (1991) suggest H2S is at least a factor of ten above solar. Even for just a 10 ppm mixing ratio, this yields an H2S mass loading of 169 mg/cm2 per bar of pressure difference. Thus, condensing all the H2S in just a 1-bar interval would make 170 times the cloud mass that is inferred for the large-particle solution and more than 5000 times the mass needed for the small-particle cloud. Thus, none of these clouds is immediately ruled out by lack of condensable supply. A more sophisticated microphysical analysis would be needed to evaluate them, accounting for eddy mixing, coagulation, sedimentation, and other effects. Another test would be to compare model spectra for these various distributions with STIS spectra at CCD wavelengths. We have verified that our STIS-based models can fit near-IR spectra, but the re- verse has not yet been demonstrated for near-IR based models. 11. SUMMARY AND CONCLUSIONS We observed Uranus with the HST/STIS instrument in 2015, following the same approach as in 2012 and 2002. We aligned the instrument's slit parallel to the spin axis of Uranus and stepped the slit across the face of Uranus from the limb to the center of the planet, building up an 49 image of half the disk with each of 1800 wavelengths from 300.4 to 1020 nm. The main purpose was to constrain the distribution of methane in the atmosphere of Uranus, taking advantage of the wavelength region near 825 nm where hydrogen absorption competes with methane ab- sorption and displays a clear spectral signature. Our revised analysis approach used a considerably simplified cloud structure, relaxed the restriction that methane and thermal profiles should be consistent with radio occul- tation results, considered the new Uranus global mean profile of Orton et al. (2014a) that was inconsistent with radio occultation results, and included parameters defin- ing the methane profile as part of the adjusted parame- ter sets in fitting observed spectra. This revised analysis applied to STIS observations of Uranus from 2015 and comparisons with similar 2002 and 2012 observations, as well as analysis of HST and Keck/NIRC2 imaging ob- servations from 2007 and 2015, and IRTF SpeX spectra from 2013, have led us to the following conclusions. 1. TEMPORAL CHANGES 1.1 A direct comparison of 2012 STIS spectra with 2015 STIS spectra reveals no statistically significant difference at low latitudes. At 10◦N and a zenith cosine of 0.7, the spectra from the two years are within the noise level of the measurements. 1.2 A different result is obtained by comparing 2012 and 2015 STIS spectra at high latitudes. There we find significant differences at pseudo-continuum wavelengths beyond 500 nm, where weaker methane bands are present, and where the 2015 I/F exceeds 2012 I/F values by up to 0.04 I/F units (about 15- 20%). However, no difference is seen in the strong methane bands that would be sensitive to changes in stratospheric aerosols. 1.3 The brightening of high latitudes at pseudo con- tinuum wavelengths between 2012 and 2015 is a re- sult of increased scattering by tropospheric aerosols, and not due to a change in the effective methane mixing ratio. This is shown by radiation transfer modeling as well as by direct comparisons of imag- ing at wavelengths with different fractions of hydro- gen and methane absorption. 1.4 The polar brightening from 2012 to 2015 that we found in comparisons of STIS spectra is part of a long-term trend evident from comparisons of H-band images from the 2007 equinox and on- ward, including recent images obtained in 2017 (Fry and Sromovsky 2017). 50 0.1 A ) r a b ( e r u s s e r P 1.0 B 10.0 10-4 ) r a b ( e r u s s e r P 0 1 2 3 4 5 0 de Kleer et al. (2015) Irwin et al. (2015) Tice et al. (2013) This work (large particle) This work (small particle) r (µm) 1.00 0.89 1.35 1.52 0.26 τTOT 2.00 5.48 7.00 13.38 0.83 N (/cm2) 2.27E+07 8.09E+07 3.84E+07 6.95E+07 4.42E+08 mg/cm2 0.0952 0.2390 0.3962 1.0229 0.0329 10-3 10-2 10-1 Optical depth/bar at 1.6 µm 100 101 102 5 10 Optical depth/bar at 1.6 µm 15 20 Fig. 35.- Comparison of tropospheric cloud density vertical profiles on log scales (A) and linear scales (B). Our small-particle fit is shown with solid lines in both panels, while results from other investigators are shown using lines defined in the legend. The Irwin et al. (2015) result has been scaled downward by a factor of 10, which is a rough correction from what is shown in the left panel their Fig. 2, suggested by P.G.J. Irwin (personal communication). Their profile was derived for a deep methane mixing ratio of 1.6% and would move upward by several hundred mbar for double that mixing ratio. The Tice et al. (2013) profile was derived using a deep methane mixing ratio of 2.2%, which is also the case for the de Kleer et al. (2015) profile. As noted in the legend, our small-particle model has much less optical depth at 1.6 µm and much less total mass than the other results shown. The estimated total column cloud mass per unit area assumes a density of 1 g/cm3. 10-3 10-4 10-5 R M V n o i t a s n e d n o c S 2 H H2S solar abundance 10-6 2.0 2.5 3.0 3.5 H2S cloud base pressure (bar) 4.0 4.5 5.0 Fig. 36.- Minimum H2S VMR required to condense at the cloud base versus cloud base pressure. 2. METHANE DISTRIBUTION: 2.1 While the increased brightness of the polar region between 2012 and 2015 is due to increased aerosol scattering, the fact that the polar region is much brighter than low latitudes in 2015 is due to the lower mixing ratio of upper tropospheric methane at high latitudes. 2.2 We found that the STIS spectra from 2015 and 2012 can be well fit by relatively simple aerosol structures. We used a two-layer cloud structure with an optically thin stratospheric haze, and one tropospheric cloud, the latter extending from near 1 bar to several bars. This is similar to the 2-cloud model of Tice et al. (2013) except that we fit the upper boundary instead of fixing the upper bound- ary and fitting the scale height ratio. The parti- cles in the tropospheric cloud were modeled either as spherical particles uniformly mixed with the gas and with a fitted real index, or as non-spherical particles using an HG phase function with a fitted asymmetry parameter. 2.3 Our initial fits to the 2015 STIS spectra over the entire range from 540 nm to 980 nm using either a two-cloud or three-cloud model using spherical par- ticles of real refractive index of n = 1.4 produced good overall fits that were especially bad near 830 nm, just where the spectrum is especially sensitive to the methane to hydrogen ratio. Much better fits were obtained by allowing the refractive index of the tropospheric aerosols to be adjusted, which yielded two solutions, one a large-particle low-index solu- tion and second small-particle high-index solution, the latter providing the better fit and somewhat closer match to the refractive index of H2S. 2.4 Our preliminary 2-cloud models using spherical particles found little variation as a result of using different temperature profiles, as long as we did not force the deep methane mixing ratio or the methane humidity above the condensation level to be constrained either by occultation results or by a prohibition against supersaturation. We chose to use the Orton et al. (2014a) profile, even though it is inconsistent with occultation results, because its higher upper tropospheric temperatures allowed more methane without supersaturation. 2.5 For subsequent models containing a stratospheric haze and just a single tropospheric conservative Mie-scattering layer mixed uniformly with the gas, we did preliminary fits to spectra at 10◦ N and 60◦ N over the 730 nm to 900 nm range, and for both 2012 and 2015, assuming that methane was uniformly mixed below the condensation level. We found two classes of solutions, one with large par- ticles of 1.1-1.75 µm in radius and a real index of 1.22±0.05 to 1.28±0.07, and a second solution set with small particles about 0.24±0.07 µm to 0.34±0.1 µm in radius with much larger real in- dex values from 1.55±0.16 to 1.86±0.30. The small particle index values are much closer to that of H2S, a prime candidate for the cloud's main constituent. 2.6 The above preliminary fits with uniform methane mixing ratios found those ratios ranged from 2.56%±0.26% to 3.16%±0.5% at 10◦N, and from 0.74%±0.05% to 0.99%±0.08% at 60◦N, with lower values in both cases obtained from the large par- ticle solutions, but good agreement between 2012 and 2015 in both cases. 2.7 Preliminary fits using non-spherical HG parti- cles for the single tropospheric cloud layer pro- duced similar results, with a methane mixing ra- tio from 2.85%±0.3% to 3.48%±0.5% at 10◦N and from 0.97%±0.06% to 1.04%±0.07% at 60◦N, and 51 in this case differences between 2012 and 2015 are within estimated uncertainties. 2.8 All the above preliminary fits found methane hu- midities in the 68% to 95% at 10◦N, and 30% to 56% at 60◦N, generally with uncertainties of 12-16% and 18-26% respectively. 2.9 STIS results in the upper troposphere are in good agreement with the Lellouch et al. (2015) results based on Herschel observations. For 2015, the rela- tive methane humidity above the nominal conden- sation level, which is roughly at the 1-bar level, for the Orton et al. thermal profile is roughly 50% north of 30◦ N but near saturation from 20◦ N and southward, but becomes supersaturated for the F1 and F0 profiles. 2.10 Latitude dependent fits assuming a uniform methane mixing ratio below the condensation level show that a local maximum value of about 3% is attained near 10◦N latitude. From that point the effective mixing ratio smoothly declines by a factor of 2 by 45◦N, and by a factor of three by 60◦N, attaining a value of about 1% from 60◦ to 70◦N. However, if particle absorption is present, the de- rived mixing ratios are lowered by up to 10% of their values, or possibly more, depending on mod- els. Thus, it is not possible to give a firm value of the mixing ratio without a deeper understanding to the aerosols within the atmosphere. 2.11 For a vertically uniform methane mixing ratio, the high-latitude model fits failed to accurately fol- low the observed spectra in the 750 nm region, sug- gesting that the upper tropospheric methane mix- ing ratio increased with depth. This was espe- cially obvious for the 2012 observations, probably because of reduced aerosol scattering in 2012. A model profile containing a vertical gradient above the 5-bar level, using either what Sromovsky et al. (2011) called a descended depletion profile or a step decrease at the 3 bar level made a substantial im- provement in the fit quality. 2.12 When the methane depletion with latitude is modeled as a stepped depletion, we find that the step change occurs at pressures between 3 and 5 bars, although the uncertainty is typically 2 bars. This level applies between about 50◦ and 70◦N, but moves to lower pressures between 50◦ N and 20◦ N, and remains near the condensation level from that point to 20◦S. The mixing ratio above the break 52 point pressure is near 0.75% in the 60◦N to 70◦N range, increasing to about 1.2% at low latitudes, al- though by that point the depleted layer is so thin that it is hard to distinguish from the uniformly mixed case with a single mixing ratio up to the con- densation level. 2.13 Because the shape of the descended profile makes the depth parameter of that profile difficult to constrain with the spectral observations, we were guided by the stepped depletion results to choose a fixed depth parameter of 5 bars, and fit just the shape parameter vx as a function of latitude. The results show a relatively smooth variation from slightly greater than 1 at high latitudes, increasing to about 4 by 30◦N, then rising to very high val- ues at low latitudes, which yields a nearly vertical profile that produces negligible depletion. 3. AEROSOL PROPERTIES: 3.1 Preliminary fits with non-spherical particles with a simple HG phase function yielded asymmetry pa- rameters that ranged from 0.43±0.04 at 10◦N to 0.26-0.39 at 60◦N. These are smaller values than the commonly used value of g = 0.7, e.g. by Tice et al. (2013). It is also smaller than the asymmetry pa- rameters of even the small-particle solutions for the tropospheric aerosols, which ranged from about 0.4 at 1.6 µm to 0.6 at 0.8 µm. The large-particle asym- metry values were near 0.86 at 0.8 µm and 0.9 at 1.6 µm. 3.2 The cloud pressure boundaries varied with model structure. When a vertically uniform methane pro- file is assumed, the top boundary of the cloud is precisely constrained and nearly invariant with lat- itude, moving from slightly greater than 1 bar at low latitudes to almost exactly 1 bar at high lati- tudes. The lower boundary is more uncertain vary- ing about a mean near 2.6 bars. The optical depth of the tropospheric cloud declines by roughly a fac- tor of two from low to high latitudes, when 2012 and 2015 results are averaged. For the stepped deple- tion models, the top boundary behavior is similar to that of the uniform model, but the bottom bound- ary moves from 2.5 bars at low latitude to 3 bars at high latitude. 3.3 A very different characteristic is seen for the de- scended methane fits as a function of latitude. In this case the upper boundary of the tropospheric cloud moves significantly downward with latitude instead of slightly upward, with the pressure in- creasing from about 1.1 bar to 1.3 bar. We also found that the refractive index increased with lat- itude, from 1.6 to about 2.0, perhaps a result of a low-index coating evaporating from a high index core as the cloud descends to warmer temperatures. The particle radius also decreases somewhat with latitude, which would be consistent with that spec- ulation. The tropospheric cloud optical depth is also seen to decline somewhat at high latitudes, as seen for other models. 3.4 The real refractive index of the main cloud has a relatively flat latitude dependence for the stepped depletion model, but significant increases with lat- itude are seen for uniform and descended depletion models. Better agreement is obtained at low lati- tudes, where weighted averages over 2012 and 2015 from 20◦S to 20◦N are 1.65±0.08, 1.66±0.07, and 1.63±0.07 for uniform, stepped depletion, and de- scended depletion models respectively, which are all above the expected value of 1.55 for H2S by amounts that are not much greater than combined uncertainties. 3.5 The way aerosol contributions produce the in- creased polar brightness between 2012 and 2015 is simplest to understand within the context of the models assuming vertically uniform methane. In these cases an increased amount of scattering in the main cloud layer produces the brightness increase. And at 60◦ N and 70◦ N this is due to a combina- tion of increased optical depth and increased par- ticle size. Similar effects are seen in the stepped depletion model model (small particle solution). In the descended depletion model it appears that an increase in the cloud top pressure over time may be a significant factor. For the simple non-spherical HG particle cloud we found a 37% increase in opti- cal depth coupled with a 33% decrease in the asym- metry parameter from 0.39 to 0.26. These effects would produce a combined rise in pseudo continuum I/F of about 32% (= 0.45×(0.37+0.33)), which is comparable to the observed change. 3.6 The association of high-latitude methane deple- tions with descending motions of an equator-to-pole deep Hadley cell does not seem to be consistent with the behavior of the detected aerosol layers, at least if one ignores other cloud generation mechanisms such as sparse local convection. Both on Uranus and Neptune (de Pater et al. 2014), aerosol layers seem to form in what are thought to be downwelling regions on the basis of the effective methane mixing ratio determinations. 3.7 Models using conservative spherical particles in the tropospheric cloud layer have significant flaws when fit to the wider spectral range from 540 nm to 980 nm and assuming a real index of refraction of n = 1.4. Much smaller flaws are seen with small particles with a larger refractive index, but more accurate fits require additional wavelength depen- dent scattering characteristics. This can be done by adding absorption in the longer wavelength regions, which allows increasing optical depths enough to brighten the shorter wavelength regions. For small particles with high real index values we needed to increase the imaginary index from zero at short wavelengths to 1.09×10−3 between 670 to 730 nm and to 4.9×10−2 from 730 nm to 1 µm. For non-spherical HG particles, we were able to match the same spectral region by creating an appropri- ate variation in optical depth with wavelength. It is also possible to produce a similar fit for DHG particles by appropriate wavelength dependence in the phase function, following an approach used by KT2009. 3.8 We were able to extend Mie model fits to the near-IR spectral range by further adjustments of the imaginary index with wavelength. For small particles the imaginary index had to be elevated to 0.1 at 1.6 µm, where its single-scattering albedo de- scends to 0.64. For large particles, the needed imag- inary index increase was to a level eight times less than for small particles, and the single-scattering albedo was decreased to a more modest value of 0.90. 3.9 Our two solutions for cloud structures that can match spectra from visible to near-IR wavelengths to at least 1.6 µm, require vast differences in the total optical depth and cloud mass. These so- lutions bound solutions from other investigators, which have different vertical structures that in most cases match spectra from 0.8 to 1.6 µm. The col- umn masses of particles in these clouds range from 500 to 17 times smaller than the total mass of H2S in a 1-bar pressure interval, and thus, even the most massive of these clouds cannot be ruled out on the basis of insufficient parent condensate. 53 3.10 We found evidence for a deep cloud layer in the 9 bar to 11 bar range if optically thick and possi- bly composed of NH4SH. Including this layer in our models has the main effect of improving our fits to the weak methane band structure at wavelengths from 540 nm to 600 nm. Placing the deep cloud at 5 bars yields an optical depth near 4 but a worse fit to the spectra. Further work is needed to better constrain the properties of this cloud. 4. CONSTRAINTS ON H2S: 4.1 If the tropospheric cloud is a condensation cloud, H2S is the likely main component. This conclusion is based on the fact that virtually all of the model cloud mass is below the level at which methane can condense, but likely within the pressure range at which H2S can condense. It is also the case that our preferred small-particle solutions are in rough agreement with the refractive index of H2S at low latitudes, although that agreement worsens at high latitudes and thus does not provide com- pelling support. More compelling support for H2S as the main constituent of this cloud is the recent detection of H2S vapor at saturation levels above this cloud (Irwin et al. 2018). What remains un- clear is whether the spectrally varying imaginary index that seems to be required for this cloud is compatible with the absorbing properties of con- densed H2S. 4.2 Based on the estimated bottom boundary of the tropospheric aerosol layer, if the small particle solu- tion is to be consistent with a composition of H2S, the mixing ratio of H2S at the 3.3-bar level and im- mediately below must be at least ∼30 ppm. To be consistent with the large particle solution would re- quire around an order of magnitude higher VMR near the 5-bar level. Advancing our understanding of the distribution and composition of Uranus' aerosols would be helped by good measurements of the optical properties of H2S, the most likely primary constituent of the most visible tropo- spheric cloud layer. Another helpful undertaking would be microphysical modeling of photochemical haze for- mation and seasonal evolution as well as microphysical modeling of condensation clouds. The variety of vertical aerosol structures and mass loadings that can produce model spectra matching the observations is surprisingly large and it seems likely that not all of these options 54 would satisfy microphysical constraints. A better under- standing of the aerosols is also the key to better con- straints on the distribution of methane because differ- ent aerosol models yield different methane mixing ratios, with deep VMR values mostly falling between 2% and 4%. As Uranus seems to be continuing to change, espe- cially the continued brightening of the north polar region through at least 2017, and many uncertainties remain, continued observations are also warranted. ACKNOWLEDGMENTS This research was supported primarily by grants from the Space Telescope Science Institute, managed by AURA. GO-14113.001-A supported LAS and PMF. Partial support was provided by NASA Solar System Observations Grant NNXA16AH99G (LAS and PMF). EK also acknowledges support by an STScI grant un- der GO-14113. I.dP was supported by NASA grant NNX16AK14G. We thank staff at the W. M. Keck Obser- vatory, which is made possible by the generous financial support of the W. M. Keck Foundation. We thank those of Hawaiian ancestry on whose sacred mountain we are privileged to be guests. Without their generous hospi- tality none of our groundbased observations would have been possible. Acton, C. H., 1996. Ancillary data services of NASA's Navigation Hernandez, S., Aloisi, A., Bohlin, R., Bostroem, A., Diaz, R., REFERENCES and Ancillary Information Facility. Planet. and Space Sci. 44, 65–70. Borysow, A., Borysow, J., Fu, Y., 2000. Semi-empirical model of collision-induced absorption spectra of H2-H2 complexes in the second overtone band of hydrogen at temperatures from 50 to 500 K. Icarus 145, 601–608. Colina, L., Bohlin, R. C., Castelli, F., 1996. The 0.12-2.5 micron Absolute Flux Distribution of the Sun for Comparison With Solar Analog Stars. Astron. J. 112, 307–315. Conrath, B., Hanel, R., Gautier, D., Marten, A., Lindal, G., 1987. The helium abundance of Uranus from Voyager measurements. J. Geophys. Res. 92 (11), 15003–15010. Conrath, B. J., Gierasch, P. J., Leroy, S. S., 1990. Temperature and circulation in the stratosphere of the outer planets. Icarus 83, 255–281. Conrath, B. J., Pearl, J. C., Appleby, J. F., Lindal, G. F., Orton, G. S., B´ezard, B., 1991. Thermal structure and energy balance of Uranus. Uranus, pp. 204–252. de Kleer, K., Luszcz-Cook, S., de Pater, I., ´Ad´amkovics, M., Hammel, H. B., 2015. Clouds and aerosols on Uranus: Radiative transfer modeling of spatially-resolved near-infrared Keck spectra. Icarus 256, 120–137. de Pater, I., Fletcher, L. N., Luszcz-Cook, S., DeBoer, D., Butler, B., Hammel, H. B., Sitko, M. L., Orton, G., Marcus, P. S., 2014. Neptune's global circulation deduced from multi-wavelength observations. Icarus 237, 211–238. de Pater, I., Romani, P. N., Atreya, S. K., 1991. Possible microwave absorption by H2S gas in Uranus' and Neptune's atmospheres. Icarus 91, 220–233. Fink, U., Sill, G. T., 1982. The infrared spectral properties of frozen volatiles. In: Wilkening, L. L. (Ed.), IAU Colloq. 61: Comet Discoveries, Statistics, and Observational Selection. pp. 164–202. Friedson, J., Ingersoll, A. P., 1987. Seasonal meridional energy balance and thermal structure of the atmosphere of Uranus - A radiative-convective-dynamical model. Icarus 69, 135–156. Fry, P. M., Sromovsky, L. A., 2017. Uranus' post-equinox north polar brightening characterized with 2013 and 2016 IRTF SpeX observation. In: AAS/Division for Planetary Sciences Meeting Abstracts. Vol. 49 of AAS/Division for Planetary Sciences Meeting Abstracts. p. 115.17. Fry, P. M., Sromovsky, L. A., de Pater, I., Hammel, H. B., Rages, K. A., 2012. Detection and Tracking of Subtle Cloud Features on Uranus. Astron. J. 143, 150–161. Hanel, R., Conrath, B., Flasar, F. M., Kunde, V., Maquire, W., Pearl, J., Pirraglia, J., Samuelson, R., Horn, L., Schulte, P., 1986. Infrared observations of the Uranian system. Science 233, 70–74. Hansen, J. E., 1971. Circular polarization of sunlight reflected by clouds. Journal of Atmospheric Sciences 28, 1515–1516. Havriliak, S., Swenson, R. W., Cole, R. H., 1955. Dielectric Constants of Liquid and Solid Hydrogen Sulfide. J. Chem. Phys. 23, 134–135. Dixon, V., Ely, J., Goudfrooij, P., Hodge, P., Lennon, D., Long, C., Niemi, S., Osten, R., Proffitt, C., Walborn, N., Wheeler, T., York, B., Zheng, W., 2012. STIS Instrument Handbook, Version 12.0, (Baltimore: STScI). Space Telescope Science Institute, Baltimore, Maryland. Irwin, P. G. J., Teanby, N. A., Davis, G. R., 2010. Revised vertical cloud structure of Uranus from UKIRT/UIST observations and changes seen during Uranus' Northern Spring Equinox from 2006 to 2008: Application of new methane absorption data and comparison with Neptune. Icarus 208, 913–926. Irwin, P. G. J., Tice, D. S., Fletcher, L. N., Barstow, J. K., Teanby, N. A., Orton, G. S., Davis, G. R., 2015. Reanalysis of Uranus' cloud scattering properties from IRTF/SpeX observations using a self-consistent scattering cloud retrieval scheme. Icarus 250, 462–476. Irwin, P. G. J., Toledo, D., Garland, R., Teanby, N. A., Fletcher, L. N., Orton, G. A., B´ezard, B., 2018. Detection of hydrogen sulfide above the clouds in Uranus's atmosphere. Nature Astronomy 2, 420–427. Karkoschka, E., Tomasko, M., 2009. The haze and methane distributions on Uranus from HST-STIS spectroscopy. Icarus 202, 287–309. Karkoschka, E., Tomasko, M. G., 2010. Methane absorption coefficients for the jovian planets from laboratory, Huygens, and HST data. Icarus 205, 674–694. Karkoschka, E., Tomasko, M. G., 2011. The haze and methane distributions on Neptune from HST-STIS spectroscopy. Icarus 211, 780–797. Khare, B. N., Thompson, W. R., Cheng, L., Chyba, C., Sagan, C., Arakawa, E. T., Meisse, C., Tuminello, P. S., 1993. Production and optical constraints of ice tholin from charged particle irradiation of (1:6) C2H6/H2O at 77 K. Icarus 103, 290–300. Krist, J., 1995. Simulation of HST PSFs using Tiny Tim. In: Shaw, R. A., Payne, H. E., Hayes, J. J. E. (Eds.), Astronomical Data Analysis Software and Systems IV. Vol. 77 of Astronomical Society of the Pacific Conference Series. pp. 349–352. Lebofsky, L. A., Fegley, Jr., M. B., 1976. Laboratory reflection spectra for the determination of chemical composition of ice bodies. Icarus 28, 379–387. Lellouch, E., Moreno, R., Orton, G. S., Feuchtgruber, H., Cavali´e, T., Moses, J. I., Hartogh, P., Jarchow, C., Sagawa, H., 2015. New constraints on the CH4 vertical profile in Uranus and Neptune from Herschel observations. Astron. & Astrophys. 579, A121. Lindal, G. F., Lyons, J. R., Sweetnam, D. N., Eshleman, V. R., Hinson, D. P., 1987. The atmosphere of Uranus - Results of radio occultation measurements with Voyager 2. J. Geophys. Res. 92, 14987–15001. Lodders, K., 2003. Solar System Abundances and Condensation Temperatures of the Elements. Astrophys. J. 591, 1220–1247. 55 Orton, G. S., Aitken, D. K., Smith, C., Roche, P. F., Caldwell, J., Sromovsky, L. A., Fry, P. M., Boudon, V., Campargue, A., Snyder, R., 1987. The spectra of Uranus and Neptune at 8-14 and 17-23 microns. Icarus 70, 1–12. Orton, G. S., Fletcher, L. N., Moses, J. I., Mainzer, A. K., Hines, D., Hammel, H. B., Martin-Torres, F. J., Burgdorf, M., Merlet, C., Line, M. R., 2014a. Mid-infrared spectroscopy of Uranus from the Spitzer Infrared Spectrometer: 1. Determination of the mean temperature structure of the upper troposphere and stratosphere. Icarus 243, 494–513. Orton, G. S., Moses, J. I., Fletcher, L. N., Mainzer, A. K., Hines, D., Hammel, H. B., Martin-Torres, J., Burgdorf, M., Merlet, C., Line, M. R., 2014b. Mid-infrared spectroscopy of Uranus from the Spitzer infrared spectrometer: 2. Determination of the mean composition of the upper troposphere and stratosphere. Icarus 243, 471–493. Rages, K., Pollack, J. B., Tomasko, M. G., Doose, L. R., 1991. Properties of scatterers in the troposphere and lower stratosphere of Uranus based on Voyager imaging data. Icarus 89, 359–376. Rannou, P., McKay, C. P., Botet, R., Cabane, M., 1999. Semi-empirical model of absorption and scattering by isotropic fractal aggregates of spheres. Plan. & Sp. Sci. 47, 385–396. Sromovsky, L. A., 2005a. Accurate and approximate calculations of Raman scattering in the atmosphere of Neptune. Icarus 173, 254–283. Sromovsky, L. A., 2005b. Effects of Rayleigh-scattering polarization on reflected intensity: a fast and accurate approximation method for atmospheres with aerosols. Icarus 173, 284–294. Sromovsky, L. A., Fry, P. M., 2007. Spatially resolved cloud structure on Uranus: Implications of near-IR adaptive optics imaging. Icarus 192, 527–557. Sromovsky, L. A., Fry, P. M., 2008. The methane abundance and structure of Uranus' cloud bands inferred from spatially resolved 2006 Keck grism spectra. Icarus 193, 252–266. Sromovsky, L. A., Fry, P. M., 2010. The source of 3-µm absorption in Jupiter's clouds: Reanalysis of ISO observations using new NH3 absorption models. Icarus 210, 211–229. Nikitin, A., 2012a. Comparison of line-by-line and band models of near-IR methane absorption applied to outer planet atmospheres. Icarus 218, 1–23. Sromovsky, L. A., Fry, P. M., Hammel, H. B., Ahue, W. M., de Pater, I., Rages, K. A., Showalter, M. R., van Dam, M. A., 2009. Uranus at equinox: Cloud morphology and dynamics. Icarus 203, 265–286. Sromovsky, L. A., Fry, P. M., Hammel, H. B., de Pater, I., Rages, K. A., 2012b. Post-equinox dynamics and polar cloud structure on Uranus. Icarus 220, 694–712. Sromovsky, L. A., Fry, P. M., Kim, J. H., 2011. Methane on Uranus: The case for a compact CH4 cloud layer at low latitudes and a severe CH4 depletion at high latitudes based on re-analysis of Voyager occultation measurements and STIS spectroscopy. Icarus 215, 292–312. Sromovsky, L. A., Karkoschka, E., Fry, P. M., Hammel, H. B., de Pater, I., Rages, K. A., 2014. Methane depletions in both polar regions of Uranus inferred from HST/STIS and Keck/NIRC2. Icarus 238, 137–155. Sun, Z., Schubert, G., Stoker, C. R., 1991. Thermal and humidity winds in outer planet atmospheres. Icarus 91, 154–160. Tice, D. S., Irwin, P. G. J., Fletcher, L. N., Teanby, N. A., Hurley, J., Orton, G. S., Davis, G. R., 2013. Uranus' cloud particle properties and latitudinal methane variation from IRTF SpeX observations. Icarus 223, 684–698. Tomasko, M. G., Archinal, B., Becker, T., B´ezard, B., Bushroe, M., Combes, M., Cook, D., Coustenis, A., de Bergh, C., Dafoe, L. E., Doose, L., Dout´e, S., Eibl, A., Engel, S., Gliem, F., Grieger, B., Holso, K., Howington-Kraus, E., Karkoschka, E., Keller, H. U., Kirk, R., Kramm, R., Kuppers, M., Lanagan, P., Lellouch, E., Lemmon, M., Lunine, J., McFarlane, E., Moores, J., Prout, G. M., Rizk, B., Rosiek, M., Rueffer, P., Schroder, S. E., Schmitt, B., See, C., Smith, P., Soderblom, L., Thomas, N., West, R., 2005. Rain, winds and haze during the Huygens probe's descent to Titan's surface. Nature 438, 765–778. Tomasko, M. G., Doose, L., Engel, S., Dafoe, L. E., West, R., Lemmon, M., Karkoschka, E., See, C., 2008. A model of Titan's aerosols based on measurements made inside the atmosphere. Plan. & Space Sci. 56, 669–707. SUPPLEMENTAL MATERIAL. The calibrated hyperspectral STIS cubes are archived at the Mikulski Archive for Space Telescopes (MAST) as High Level Science Products (HLSPs). They can be found at https://archive.stsci.edu/prepds/uranus-stis/; https://dx.doi.org/10.17909/T9KQ4N. The 2015 cube is named hlsp uranus-stis hst stis uranus-2015 g430l-g750l v1 cube.fits; 2002 and 2012 cubes are named analagously. The hy- perspectral cubes contain calibrated I/F values as a function of wavelength and location, with navigation backplanes that provide viewing geometry and latitude-longitude coordinates for each pixel. A detailed explanation of the file contents is provided in the file README SUPPLEMENTAL.TXT. A sample IDL program that reads a cube file, plots a monochromatic image, extracts data from a particular location on the disc, and plots a spectrum, is provided in the file stis cube example.pro. The IDL astronomy library will be needed to run the sample program. NOTE: Until the MAST archive submission is finalized, which is underway at this writing and expected to be complete by the end of June 2018, the materials will be available for review at http://www.ssec.wisc.edu/planetary/ uranus/ onlinedata/ura2015stis/.
1303.2928
2
1303
2013-03-21T20:26:17
Resonant Behavior of Comet Halley and the Orionid Stream
[ "astro-ph.EP" ]
Comet 1P/Halley has the unique distinction of having a very comprehensive set of observational records for almost every perihelion passage from 240 B.C. This has helped to constrain theoretical models pertaining to its orbital evolution. Many previous works have shown the active role of mean motion resonances in the evolution of various meteoroid streams. Here we look at how various resonances, especially the 1:6 and 2:13 mean motion resonances with Jupiter, affect comet 1P/Halley and thereby enhance the chances of meteoroid particles getting trapped in resonance, leading to meteor outbursts in some particular years. Comet Halley itself librated in the 2:13 resonance from 240 B.C. to 1700 A.D. and in the 1:6 resonance from 1404 B.C. to 690 B.C., while stream particles can survive for timescales of the order of 10,000 years and 1,000 years in the 1:6 and 2:13 resonances respectively. This determines the long term dynamical evolution and stream structure, influencing the occurrence of Orionid outbursts. Specifically we are able to correlate the occurrence of enhanced meteor phenomena seen between 1436-1440, 1933-1938 & 2006-2010 with the 1:6 resonance and meteor outbursts in 1916 & 1993 with the 2:13 resonance. Ancient as well as modern observational records agree with these theoretical simulations to a very good degree.
astro-ph.EP
astro-ph
Resonant Behavior of Comet Halley and the Orionid Stream A. Sekhar1 , 2 and D. J. Asher1 1Armagh Observatory, College Hill, Armagh BT61 9DG, United Kingdom 2Queen's University of Belfast, University Road, Belfast BT7 1NN , United Kingdom E-mail: [email protected] , [email protected] Received: 24 Aug 2012; Accepted: 6 Mar 2013; Meteoritics & Planetary Science Abstract- Comet 1P/Halley has the unique distinction of having a very comprehensive set of observational records for almost every perihelion passage from 240 B.C. This has helped to constrain theoretical models pertaining to its orbital evolution. Many previous works have shown the active role of mean motion resonances in the evolution of various meteoroid streams. Here we look at how various resonances, especially the 1:6 and 2:13 mean motion resonances with Jupiter, affect comet 1P/Halley and thereby enhance the chances of meteoroid particles getting trapped in resonance, leading to meteor outbursts in some particular years. Comet Halley itself librated in the 2:13 resonance from 240 B.C. to 1700 A.D. and in the 1:6 resonance from 1404 B.C. to 690 B.C., while stream particles can survive for timescales of the order of 10,000 years and 1,000 years in the 1:6 and 2:13 resonances respectively. This determines the long term dynamical evolution and stream structure, influencing the occurrence of Orionid outbursts. Specifically we are able to correlate the occurrence of enhanced meteor phenomena seen between 1436-1440, 1933-1938 & 2006-2010 with the 1:6 resonance and meteor outbursts in 1916 & 1993 with the 2:13 resonance. Ancient as well as modern observational records agree with these theoretical simulations to a very good degree. Keywords: comet, meteor, orbit 1 Introduction Various contributions from ancient civilizations have helped in making a detailed observational record (Yeomans & Kiang 1981) of comet 1P/Halley for almost every perihelion passage right from 240 B.C. There are no credible observations relating to this comet before 240 B.C. Further- more comet Halley has reliably determined (Yeomans & Kiang 1981) perihelion passage times (the first calculations done by Halley 1705 using Newton 1687) and orbital elements back till 1404 B.C., beyond which the uncertainty in the orbit starts to increase because of a significant close encounter with Earth at a distance of about 0.04 AU. 1 Historical confirmations of the annual nature of the Orionid meteor shower date back to as early as Edward Herrick's observations in 1839 (Lindblad & Porubcan 1999) and Alexander Her- schel's radiant determination (Denning 1899) in 1864 (Herschel 1866). Many ancient records of meteors seen in October from the Chinese, Japanese and Korean civilisations (Imoto & Hasegawa 1958, Zhuang 1977) could also correspond to the Orionid shower. Nevertheless the association of the stream with comet Halley and explaining the differences of the Orionid shower compared to the Eta Aquariids (which have the same parent body) has been a very challenging task (McIn- tosh & Hajduk 1983, McIntosh & Jones 1988) which interested many theoreticians for decades. Coincidentally it is widely believed that Sir Edmond Halley was the first (by 1688) to suggest that meteors were of cosmic origin (Williams 2011). Comet Halley might lose approximately 0.5% of its mass during every perihelion passage (Whipple 1951, Kresak 1987) which would predict its physical lifetime to be a couple of hundred revolutions or ∼15 kyr. The dynamical lifetime (time scale to remain on any kind of Halley type comet orbit i.e. orbital period from 20 to 200 years) of 1P/Halley is estimated to be of the order of 100,000 years (Hughes 1985, Hadjuk 1986, Steel 1987, Bailey & Emel'yanenko 1996). Bailey & Emel'yanenko (1996) showed that Halley undergoes Kozai resonance (Kozai 1979) during its long term evolution. It is reasonable to believe (see also Section 4 below) that Halley has been on an orbit comparable to its present one (with perihelion distance q < ∼ 1 AU), and outgassing particles thereby populating the Orionid stream, for a couple of tens of kyr. It is interesting to note that the zenithal hourly rates (ZHR) of Orionids are non-uniform (Miskotte 1993, Rendtel & Betlem 1993, Rendtel 2007, Trigo-Rodriguez et al. 2007, Arlt et al. 2008, Rendtel 2008, Kero et al. 2011, IMO database) with respect to each year. Many previous works have discussed the active role of mean motion resonances (MMR) in the dynamical evolution of various meteoroid streams (e.g. Asher et al. 1999, Emel'yanenko 2001, Asher & Emel'yanenko 2002, Ryabova 2003, Ryabova 2006, Jenniskens 2006, Vaubaillon et al. 2006, Jenniskens et al. 2007, Christou et al. 2008, Soja et al. 2011), and consequent year to year variations in shower activity. Our work aims to study the long term evolution of Halley and its associated stream focusing especially on past resonant behavior. We model particles ejected from the comet and try to correlate these with ancient as well as present observational records of meteor showers. 2 Resonant Motion of Comet Halley Over the time frame during which 1P/Halley's orbit is reliably known, i.e. since 1404 B.C., our calculations show that the comet was resonant in the past: it was trapped in the 1:6 and 2:13 MMR with Jupiter from 1404 B.C. to 690 B.C. and 240 B.C. to 1700 A.D. respectively. Integrations were repeated for different values of non-gravitational parameters (Marsden et al. 1973, Marsden & Williams 2008) to ensure that this resonant pattern is not sensitive to small changes in non-gravitational forces. Fig. 1 shows the 1:6 resonant argument librating from 1404 B.C. to about 690 B.C., and Fig. 2 shows the 2:13 resonant argument librating from 240 B.C. to 1700 A.D. All the orbit integrations in this work were done using the MERCURY package (Chambers 1999) incorporating the RADAU algorithm (Everhart 1985), and including the sun and eight planets, whose orbital elements were taken from JPL Horizons (Giorgini et al. 1996). Elements for the comet were taken from Yeomans & Kiang (1981). 2 Figure 1: Evolution of 1:6 resonant argument of 1P/Halley over 6000 years from 2404 B.C. Since the comet has a retrograde orbit, the change in the definition of longitude of pericentre (Saha & Tremaine 1993, Whipple & Shelus 1993) was incorporated while computing the resonant arguments: = Ω − ω (1) where Ω and ω are longitude of ascending node and argument of pericentre respectively. In order to absolutely confirm the librating versus circulating behavior of the resonant ar- gument during the time frames mentioned above, various combinations of terms to define the resonant argument (Murray & Dermott 1999, Sections 6.7 and 8.2) according to the D'Alembert rules were verified. Mathematically the D'Alembert rule is given by Equation 2, and Equations 3 and 4 should be satisfied for Equation 2 to be valid. In the case of the p:(p+q) mean motion resonance σ = pλj − (p + q)λc + k1c + k2j + k3Ωc + k4Ωj (2) where q is the order of resonance, σ and λ denote resonant argument and mean longitude respectively, and subscripts c and j stand for the comet and Jupiter. k1 + k2 + k3 + k4 = q k3 + k4 = 0, 2, 4, ...... (3) (4) For the 1:6 MMR (q=5) there are 28 combinations and each of them was checked, verifying the interval 1404 to 690 B.C. shown in Fig. 1. For the 2:13 MMR (q=11) there are 182 combinations 3 Figure 2: Evolution of 2:13 resonant argument of 1P/Halley over 6000 years from 2404 B.C. of which 50 were checked, all of them verifying the result of Fig. 2. In Figures 1 and 2, σ is plotted for the combinations shown in Equation 5 and 6 respectively. k1 = 5, k2 = k3 = k4 = 0 k1 = 5, k2 = 4, k3 = 1, k4 = 1 (5) (6) When the comet itself is resonant, there are more chances for the ejected meteoroid particles to get trapped in resonance which in turn would enhance the chances for meteor outbursts in future years. That is an important motivation for looking into the resonant behavior of the parent body. 3 Resonant Structures in the Orionid Stream 3.1 General Schematic Figures 3 and 4 shows the general schematic of the geometry of resonant zones in the case of 1:6 (an=17.17 AU) and 2:13 (an=18.11 AU) resonances respectively. Here we quote a = an = the 'nominal resonance location' (Murray & Dermott 1999, Section 8.4), which is the value of semi-major axis corresponding to a resonant orbital period assumes the most simple case where (d/dt) (which denotes orbital precession) is zero, i.e. as implied by Kepler's third law (Kepler 1609, 1619) In a real case (d/dt) is never exactly zero, e.g. with the 1:6 resonance (resonant argument as per Equation 5): 4 Figure 3: (a,M) space for 1:6 resonance showing regions where particles undergo resonant li- brations, as a function of initial semi-major axis and mean anomaly at the initial epoch JD 1208880.5. σ = λj − 6λc + 5c If we assume, as a time average, for resonant libration: dσ/dt = 0 then λj − 6(c + Mc) + 5c = σ where M denotes the mean anomaly. Simplifying the expression we get λj − c − 6Mc = σ Differentiating Equation 10 on both sides with respect to time and using Equation 8, (dλj /dt) − (dc/dt) − 6(dMc/dt) = 0 (7) (8) (9) (10) (11) From our numerical integrations we find that (dc/dt) is always positive for these resonant particles. If (dc/dt) is positive, then we require the rate of change of mean anomaly to be smaller compared to the 'nominal' case, which in turn means an increase in the value of semi- major axis. 5 Figure 4: (a,M) space for 2:13 resonance (cf. Fig. 3); initial epoch JD 1633920.5 areal = an + ∆a (12) Therefore the actual resonant value of semi-major axis would be slightly greater than the ones given in this section. A much more comprehensive and general description about this subject and its application to meteor streams can be found in Emel'yanenko (2001). Integrations were done by taking 7200 particles, varying the initial a from 16.5 to 17.9 in steps of 0.014 AU for 1:6 MMR and from 17.6 AU to 18.6 AU in steps of 0.01 AU for 2:13 MMR, and initial M from 0 to 360 degrees in steps of 5 degrees, keeping all other orbital elements (namely e, i, ω and Ω) constant. The starting epochs for Fig 3 and Fig 4 are 1P/Halley's perihelion return times in 1404 B.C. and 240 B.C. respectively. All the particles were integrated for 2,000 years using the RADAU algorithm with accuracy parameter set to 10−12. Output data were generated for every 10 years. Resonant particles were identified by employing a simple technique which looks at the overall range of the resonant arguments (for various combinations allowed by D'Alembert rules, see Section 2) of all particles during the whole integration time to check when there is no circulation. Checks on an extensive set of representative particles covering many different libration amplitudes confirm that those particles filtered by this algorithm will have librating resonant arguments which thereby confirm the presence of respective mean motion resonances. Figures 3 and 4 give a general picture of these resonances: we can visualize 6 or 13 resonant zones spaced in mean anomaly along the whole orbit, each zone consisting of individual librating particles (cf. Emel'yanenko 1988). These zones, or clouds of resonant particles, are preserved for as long as substantial numbers of particles continue to librate. Our test integrations showed that for some particular ejection epochs the 1:6 MMR is exceptionally effective in retaining the compact dust trail structures for as long as 30,000 years. However particles disperse in mean anomaly much faster (in a few thousands of years) in the case of the 2:13 resonance and do not show such high stability. Typically the rule of thumb is that the higher the order of resonance 6 Figure 5: Ascending Nodal Distance in 2007 vs Initial Semi-major Axis of Meteoroids in -910. (denoted by q, see equation 3), the lower the strength of the resonance. A necessary condition for a resonant meteor outburst is that the Earth should encounter one of these clusters of resonant particles, i.e. when the Earth misses these clusters, there is no enhancement (which is the common case in most years) in meteor activity, at least due to the MMR mechanism. Of course there are various other factors like nodal distance, solar longitude, date and time of intersection of the meteoroid with Earth, geocentric velocity etc. which play a key role in confirming the occurrence and characteristics of a meteor outburst or storm (see Section 3.2). It is possible in reality firstly that many resonant zones would only be partially filled (unlike the uniform pattern shown in Figures 3 and 4), and secondly that within a given resonant zone there is significant fine structure which could lead to enhanced meteor phenomena if Earth happens to pass exactly through the densest parts. Similar plots to Figures 3 and 4 reveal mean anomaly distributions of resonant particles in the long term (a few millennia). The cluster of particles in a single 1:6 resonant zone occupies a much longer part of the orbit (covering 5-6 years) than the equivalent for a 2:13 resonant zone (only 1-2 years). This means that for the 1:6 MMR there can be 5-6 consecutive years of enhanced meteor activity (depending on the exact parts of the resonant stream's fine structure encountered by the Earth), compared to just 1-2 years of outburst possibilities from 2:13. The 1:6 resonance is also more effective in trapping considerably larger numbers of particles compared to 2:13. Hence meteor outbursts from the 1:6 resonance would have a higher intensity than those due to the 2:13 resonance in most cases, though dependent again on the fine structure within the respective resonances. Since there is a long record of observations for Orionids (Rendtel 2008) the same pattern could be compared and justified. The maximum ZHR in 2007 was about 80 (Arlt et al. 2008) whereas in 1993 it was about 35 (Rendtel & Betlem 1993); cf. normal rates of 20-25. When the comet is resonant, it would remain in a single resonant zone and populate that particular zone. When the comet goes out of resonance, it would keep traversing between zones 7 Figure 6: Solar Longitude in 2007 vs Initial Semi-major Axis of Meteoroids in -910. and thereby populate different resonant zones gradually, implying that meteor outbursts could come from different zones. One of the main aims of this work is to correlate particular res- onant zones to past and present meteor outbursts. Our calculations show that the outbursts during 1436-1440, 1933-1938 and 2006-2010 (Section 3.2) are from the same 1:6 resonant zone, specifically the same one in which the comet librated from 1404 B.C. to 690 B.C. Also a future meteor outburst in 2070 would occur due to the particles in the same 2:13 resonant zone which caused increased meteor activity in 1916 and 1993. Halley presumably released many meteoroids into the 2:13 resonant zone in which it librated from 240 B.C. to 1700 A.D. but most of these meteoroids do not have the precession rate required for producing Earth-intersecting orbits at the present time. Comparison with resonant zones are an excellent way to match observations with theoretical simulations. The numbers of particles trapped in other resonances close to this range of semi-major axis (16.5 AU to 19 AU) were also checked. For example the percentage of particles in 1:7 (an=19.03 AU), 3:17 (an=16.53 AU), 3:19 (an=17.80 AU) and 3:20 (an=18.42 AU) resonances with Jupiter are very low compared to the 1:6 (an=17.17 AU) and 2:13 (an=18.11 AU). It should be pointed out that if particle ejection were centred around the resonant value of 1:7 MMR, then there would be substantial amounts of resonant particles which can cause outbursts, but in reality the comet is never near this resonant semi-major axis in the time frames which we consider in this work. Also the ratio of particles trapped in the Saturnian resonances of 2:5 (an=17.56 AU), 3:8 (an=18.34 AU), 5:12 (an=17.09 AU), 5:13 (an=18.03 AU), 7:17 (an=17.23 AU), 7:18 (an=17.90 AU) and 8:19 (an=16.97 AU) are extremely small owing to the overpowering effect of Jupiter's gravity. Hence significant measurable enhancements in meteor activity can be ruled out from these obscure Jovian and Saturnian resonances. Even if such resonant particles encounter Earth it will be almost impossible to distinguish them because of the lack of any sizeable increase in ZHR in any year. Hence it should be understood that just the mere fact of having some resonant particles intersecting Earth does not mean an increase from normal meteor rates. The sole criterion depends on how effective that resonance mechanism is in trapping very large numbers 8 Figure 7: Difference in Nodal Passage Times in 2007 vs Initial Semi-major Axis of Meteoroids in -910. of ejected particles and subsequently avoiding close encounters with other planets. 3.2 Specific Calculations Past observations (Millman 1936, Lovell 1954, Imoto & Hasegawa 1958, Rendtel & Betlem 1993, Dubietis 2003, Rendtel 2007, Trigo-Rodriguez et al. 2007, Arlt et al. 2008, Spurny & Shrbeny 2008, Kero et al. 2011) of Orionids have shown enhanced meteor activity in some particular years. Previous interesting works (Rendtel 2007, Sato & Watanabe 2007) have highlighted the significance of 1:6 MMR in explaining the outburst in 2006. According to Sato & Watanabe (2007), the meteor outburst in 2006 was caused by 1:6 resonant particles ejected from the comet in -1265 (1266 B.C.), -1197 (1198 B.C.) and -910 (911 B.C.). All these ejection years can be directly linked to the time frame in which the comet itself was 1:6 resonant (Section 2). Hence more meteoroids became trapped in this resonance during this time frame compared to other years when the comet was not resonant. Calculations were done on similar lines to Sato & Watanabe (2007). Ejection epochs were set between 1404 B.C. and 1986 A.D. All the ejections were done by keeping the perihelion distance and other elements as constant and by varying the semi-major axis and eccentricity. In this simple model, ejection was done at perihelion (M=0) in the tangential direction. Ejection velocities were set in the range -50 to +50 m/s, i.e. both behind and ahead of the comet, meaning that over all epochs collectively the initial orbital periods range from 60 to 88 years, encompassing all possible 1:6 and 2:13 resonant particles, positive ejection velocity corresponding to larger periods. Radiation pressure and Poynting-Robertson effects were not incorporated in these calculations. Because they span a range of orbital periods, test particles ejected tangentially at perihelion and moving only under gravitational perturbations are able, over the time frames we consider here, to represent the motion of all meteoroids released over the comet's perihelion arc with different 9 Figure 8: Ascending Nodal Distance in 2007 vs Initial Orbital Period of Meteoroids in -910. velocities and subject to different radiation pressures (Kondrat'eva & Reznikov 1985, Asher & Emel'yanenko 2002). In order to confirm the correlation between theory and observations, it is vital to match the time (second half of October) when the meteoroids reach their ascending node, solar longitude λ⊙ (approx. 204-210 degrees) at the node and heliocentric distance of ascending node (by analyzing the difference ∆r between heliocentric distances of Earth and ascending node of meteoroid; Earth diameter is about 0.0001 AU). These essential parameters from our simulations (Table 1) can be matched with real observations (listed in past meteor records). Each entry in Table 1 is a carefully chosen meteoroid which has the average value out of many candidate particles covering the small range of orbital periods that favors an outburst in the given year. For example, Figures 5, 6 and 7 are plots of heliocentric distance of ascending node, solar longitude and difference in time of nodal crossing of the particles from the time of observed outburst (all three parameters computed at 2007 Oct 22) versus initial semi-major axis of meteoroids (ejected at -910 return). Heliocentric distance of Earth on 2007 Oct 22 was 0.995 AU. These results show that the conditions for an outburst to occur (from particles ejected in -910) at the observed time in 2007 is satisfied if initial semi-major axis is around 17.22 AU, but the total suitable range in initial semi-major axis and ejection velocities for meteoroids are 17.20 AU <a <17.25 AU and -16.53 m/s <v <-15.05 m/s respectively. The plots are similar for other ejection epoch / outburst year pairs as well. Our simulations indicate meteor outbursts from 1436-1440 A.D. due to 1:6 resonant mete- oroids which were ejected around Halley's -1265, -1197, -985, -910 and -836 returns (details in Table 1). The initial orbital periods of ejected meteoroids which lead to all five outbursts show that most of them had almost 6 times the Jovian period. There are also historical observa- tional records which show heightened activity, indicating hundreds of bright meteors, in 1436 and 1439 (Imoto & Hasegawa 1958) and match our theoretical simulations well. We converted the dates from Julian calendar to proleptic Gregorian calendar in order that all dates in Table 10 Figure 9: Evolution of 1P/Halley's semi-major axis in an integration going back in time from 240 B.C. 1 are referred to a single calendar. No observational records could be traced or identified for 1437, 1438 and 1440 though. Either there were no observations done in those years (unfavorable lunar phase is a possible explanation only in 1438) or the meteor outbursts would have been insignificant in 1437, 1438 and 1440 compared to the ones in 1436 and 1439. In our simulations we find that resonant meteoroids ejected with positive ejection velocity (higher orbital period) encountered Earth in 1436 and 1439. The ones with negative ejection velocity (smaller period) encountered Earth in 1437, 1438 and 1440. Radiation pressure (not included in our integrations) would always act in the direction which would increase the orbital period of meteoroids, i.e. affects the period in the same sense as positive ejection velocities. In general we expect the peak of the ejection velocity distribution is close to zero and so the largest number of particles, if affected by radiation pressure, is represented by particles having positive ejection velocites in our gravitational integration model. Radiation pressure is having a detrimental effect (with regard to causing meteor outbursts) when we calculate that negative ejection velocities are required to produce a meteor outburst in a particular year. This can explain why we find this trend in ejection velocities for resonant meteoroids reaching Earth in 1436 and 1439 which caused meteor outbursts (agreeing with past observations as shown in Imoto & Hasegawa 1958) and possibly no (or very low) activity in 1437, 1438 and 1440. We calculate that the meteor outburst in 1993 was due to 2:13 resonant meteoroids ejected around Halley's -1333, -985, -910 and -835 returns. The outburst (Miskotte 1993, Rendtel & Betlem 1993) occurred when solar longitude was between 204.7 to 204.9 degrees, a notably different time compared to other known outbursts. Our theoretical calculations match this unusually early peaking on 1993 Oct 18. Our simulations also indicate a meteor outburst in 1916 from the 2:13 resonance and there is a hint of enhanced meteor rates from past observations in 1916 (Olivier 1921) compared to the adjacent years of 1915 and 1917. For the future we predict a similar outburst (like in 1993 because favorable ejection velocities are similar in both cases) from the 2:13 resonance mechanism in 2070. 11 Figure 10: Heliocentric distance of descending node of 1P/Halley in an integration going back in time from 240 B.C. The ejection epochs for 1:6 resonant meteoroids which caused continuous enhanced activity in 2007, 2008, 2009 and 2010 (Trigo-Rodriguez et al. 2007, Arlt et al. 2008, Kero et al. 2011, International Meteor Organization database) are also given in Table 1. These ejection years correspond to the time when the comet itself was 1:6 resonant. Hence it is obvious that a large number of meteoroids would have been trapped into this resonance during those time frames which would clearly indicate the reason for high ZHR apart from the contribution due to the inherent geometry (see Section 3.1) of these zones. Our simulations match the observed ranges (207-210 degrees) of solar longitude and outburst times (Trigo-Rodriguez et al. 2007, Arlt et al. 2008, Kero et al. 2011, IMO database) for these outburst years very well. Even though the uncertainties in semi-major axis (to directly compare with theoretical values in our calculations) of observed meteoroids from these highly successful observations are quite high (which is the typical case for all meteor observations, especially when the semi-major axis itself is high), the matching of outburst time frames and solar longitudes from these papers itself is a very effective way of comparing the orbital evolution of resonant meteoroids with real observations. Our results show that 1:6 resonant meteoroids ejected from the resonant comet also caused enhanced activity from 1933-1938 which match old observational records (Millman 1936, Lovell 1954). Most of these meteoroids had positive ejection velocities which is more favorable for stronger outbursts as discussed before. Fig. 8 clearly shows that meteoroids with initial orbital periods corresponding to 1:6 (around 71 years) and 2:13 (around 77 years) resonances have their ascending nodes near the orbit of the Earth. Hence it can be concluded that resonance mechanisms (specifically 1:6 and 2:13 MMR in this case) aid these particles to come near the Earth at the present epoch while the non-resonant ones precess away from the Earth's orbit. This is typical of other ejection epochs (as shown in Table 1) as well. The ascending node of Halley during its last apparition (in 1986) was 1.8 AU. One could clearly see that the number of particles trapped in 1:6 resonance is considerably larger than the number trapped in 2:13 resonance. Moreover in Fig. 5 we notice that particles having orbital periods of almost 5Pj(59.2 years), 6Pj(71.1 years) and 7Pj(83.0 years) come near 12 the Earth's orbit which agrees with earlier calculations done by Sato & Watanabe (2007). The typical ZHR for Orionids during non-outburst years is about 20 (Rendtel & Betlem 1993, Rendtel 2008, IMO records). From the recorded previous observations it is seen that the ZHR is about 60 (Rendtel 2007, Kero et al. 2011, IMO records) due to 1:6 MMR during 2006- 2010 and about 35 (Rendtel & Betlem 1993) due to 2:13 MMR in 1993. Using these previous observations and flux, one could actually make a simplistic estimation of the mass delivered to Earth from the Orionid stream during these outburst years. According to the detailed work of Hughes and McBride 1989, the typical influx rate (at shower maximum perpendicular to the radiant) of Orionids is 1.8 × 10−18gcm−2s−1 which in turn (after multiplying with the incident area of Earth) predicts 7 g/s during 2006-2010 and 4 g/s during 1993. It should be made clear that enhanced ZHR could be quite different in other outburst years (in past as well as future) as it depends on the exact cross section and density distribution of resonant trails intersecting the earth. Moreover, given the high speed of Orionid meteors and the strong dependence of meteor brightness on velocity, the meteoroidal mass influx to Earth is not dominated by this level (ZHR = a few tens) of Orionid outburst. 4 Orbit of Halley before 1404 B.C. Our calculations show that the orbit of Halley was substantially different from the present orbit at about 12,000 years in the past. Fig. 9 shows the time evolution of semi-major axis, indicating a drastic change in the semi-major axis near this time frame. A similar sudden change occurred in eccentricity, inclination and longitude of pericentre. Fig. 10 plots the time evolution of heliocentric distance of descending node, showing that close encounters with Jupiter are the reason for this drastic variation in the comet's orbit. 100 clones with orbits very similar (varying semi-major axis and eccentricity minutely while keeping the perihelion distance as constant) to the comet were integrated 30,000 years backwards in time from 240 B.C. and this behavior is typical for about 95% of the clones. From these orbital integrations it is clear that any meteoroid ejection before 12,000 years in the past would not correspond to the present day Orionid meteor shower. Hence this particular time constraint can be used as a starting epoch for ejection to simulate the present day Orionid stream. It is also interesting to note that this timescale is close to the physical lifetime of the comet itself. In our test simulations, almost 80% of the clones get trapped into 1:6 and 2:13 resonances for at least a few thousand years between -12,000 and -1403. Hence it is confirmed that the phenomenon of resonance plays a vital role in the long term dynamical evolution of Halley itself which further stresses the motivation in looking into more resonant structures in the present day Orionid stream. This gives good scope for a lot of interesting further work. 5 Main Results We find that dust trails formed by 2:13 resonant meteoroids caused the unusual meteor outbursts on 1993 Oct 18 (Miskotte 1993, Rendtel & Betlem 1993) and 1916 Oct 17 (Olivier 1921). Meteor outbursts from 1436-1440 and 1933-1938 were due to the 1:6 resonance mechanism which matches historical observations in 1436 and 1439 (Imoto & Hasegawa 1958) and 1933-1938 (Millman 1936, Lovell 1954). Furthermore we are able to correlate the recent observations of outbursts from 13 Table 1: Data of dust trails which caused various Orionid outbursts Ejection Expected peak time λ⊙ Ejection Period at velocity (m/s) +13.16 -12.79 -13.60 -22.08 +11.88 -14.74 -18.07 -20.37 +10.64 +8.79 +15.69 -12.18 +9.13 -11.16 -16.42 -14.35 +17.35 +10.77 +17.27 -13.68 +7.00 +14.31 +31.88 +12.68 +8.98 +10.45 +11.03 -17.73 +8.60 +12.95 +11.25 -15.22 -15.81 -15.76 +6.61 +14.65 +9.30 -18.72 -17.34 -17.88 +6.30 +15.26 +9.24 +9.80 ejection (years) 71.76 71.64 71.95 70.12 70.67 71.23 71.01 70.54 77.22 77.29 71.54 71.77 71.82 71.99 71.88 71.79 71.93 71.21 71.91 71.85 71.37 71.22 77.70 77.77 77.34 77.64 70.48 71.08 71.70 70.91 71.32 71.13 71.48 71.41 71.25 71.30 70.88 70.41 71.16 70.97 71.18 71.44 77.41 77.56 year -1197 -985 -910 -836 -1265 -985 -910 -910 -985 -910 -1265 -985 -1333 -985 -985 -910 -1265 -1197 -1265 -836 -1197 -1265 -1333 -985 -910 -835 -1265 -910 -1333 -1265 -1197 -985 -910 -836 -1333 -1265 -1197 -985 -910 -836 -1333 -1265 -910 -910 (UT) 1436 Oct 13 01:44 1436 Oct 14 17:40 1437 Oct 14 03:00 1438 Oct 14 13:30 1439 Oct 14 23:54 1439 Oct 15 00:00 1439 Oct 16 15:03 1440 Oct 13 19:17 1916 Oct 17 07:40 1916 Oct 17 12:57 1933 Oct 21 02:24 1933 Oct 21 02:52 1934 Oct 21 12:14 1934 Oct 21 12:28 1935 Oct 21 13:26 1935 Oct 22 05:16 1936 Oct 21 16:19 1936 Oct 22 06:28 1937 Oct 21 20:24 1937 Oct 21 23:02 1938 Oct 21 21:21 1938 Oct 22 02:24 1993 Oct 17 22:48 1993 Oct 18 00:14 1993 Oct 18 02:26 1993 Oct 18 02:40 2006 Oct 21 02:09 2006 Oct 23 03:38 2007 Oct 21 18:14 2007 Oct 22 00:28 2007 Oct 22 04:36 2007 Oct 22 09:21 2007 Oct 22 10:04 2007 Oct 23 01:12 2008 Oct 23 05:16 2008 Oct 20 14:40 2008 Oct 21 07:28 2009 Oct 21 15:07 2009 Oct 21 19:43 2010 Oct 21 21:36 2010 Oct 21 23:31 2010 Oct 22 02:24 2070 Oct 18 19:12 2070 Oct 18 19:26 (J2000.0) 207.551 208.212 208.344 208.522 208.698 208.702 210.324 208.258 204.771 204.990 208.170 208.190 208.321 208.330 208.120 208.771 208.982 209.568 208.890 208.998 208.675 208.883 204.662 204.724 204.774 204.782 207.452 209.461 207.858 208.118 208.252 208.486 208.513 209.140 210.050 207.412 208.114 208.214 208.367 208.225 208.302 208.424 204.770 204.779 ∆r (AU) +0.0012 +0.0027 -0.0013 +0.0010 +0.0021 -0.0001 -0.0008 +0.0037 +0.0002 +0.0014 +0.0011 +0.0086 +0.0044 +0.0002 +0.0034 +0.0087 -0.0073 +0.0058 +0.0067 +0.0022 -0.0091 +0.0001 +0.0039 +0.0041 -0.0017 +0.0088 +0.0005 -0.0069 +0.0043 -0.0004 +0.0032 +0.0047 +0.0002 -0.0071 +0.0073 +0.0022 -0.0019 +0.0077 -0.0008 -0.0011 +0.0056 +0.0013 -0.0029 +0.0003 14 2006-2010 (Rendtel 2007, Trigo-Rodriguez et al. 2007, Arlt et al. 2008, Kero et al. 2011, IMO database) due to 1:6 resonant meteoroids with our theoretical simulations. These correlations are very promising and give us great confidence in confirming theory with observations. Using similar techniques one could also predict similar events for the future. We foresee a meteor outburst in 2070 (due to the 2:13 resonance) similar to the 1993 outburst. Using the data (ZHR and mass flow rate) from previous observations it is also possible to roughly estimate the mass influx in outburst years (see Section 3.2). Although non-resonant particles can produce random outbursts, our calculations show that a substantial majority of Orionid outbursts are due to resonant structures in the meteor stream. However it must be pointed out that much older meteoroids (ejected before 1404 B.C.) may also contribute to all these outbursts (which makes further backward integrations and calculations very crucial). 6 Conclusions and Future Work Most of the theoretical aspects of these two resonances can have very significant and interesting effects on real observations. The compact dust trails getting preserved for many 10 kyr due to 1:6 MMR hint at an exciting possibility that strong meteor outbursts could occur in the future even after the comet becomes extinct, i.e. survival times of some resonant structures could be much higher than the physical lifetime of the parent body. It is well known and obvious that understanding the history of comets is crucial to predicting meteor showers. In this work we find that Orionid outbursts in 1436-1440,1916,1933-1938, 1993 and 2006-2010 were caused by resonant particles ejected from Halley before 240 B.C., the date beyond which there are no direct observational records of the comet. As a corollary to the above point about the importance of knowing comets' histories, one could argue that non-uniform meteor rates can act as a great tool to backtrack the history of a comet beyond the time frame in which there are direct sightings of the comet itself. All of these prove how useful the comparison between meteor observations and these simulations are. In short it is an indirect confirmed observation of the comet beyond 240 B.C. Even though the Eta Aquariid shower is considerably different (McIntosh & Hajduk 1983) from the Orionid shower in many ways, it would be worthwhile to verify whether all these resonant phenomena and enhanced activity are applicable in its case as well (CBET 944, 2007). The low number (compared to Orionids) of credible observations of Eta-Aquariids is a limitation in this regard though. Near future releases of radio results on Eta Aquariids (personal communications with Campbell-Brown and Jenniskens) would be very promising in this direction. Negative observations, comprising diminished meteor rates of Orionids in some particular years compared to adjacent years (e.g. ZHR reaching only 7 in 1900: Kronk 1988) could be as scientifically valuable as enhanced meteor phenomena which we have investigated in this work. A future careful study of such events can also be intriguing in many aspects. As the next step we plan to design an ejection model to simulate the Orionid stream beginning 12,000 years in the past and to correlate more past and present observations as accurately as possible. 15 7 Acknowledgements The authors wish to thank both the anonymous reviewers for the helpful comments and also intend to express their gratitude to the Department of Culture, Arts and Leisure of Northern Ireland for the generous funding to pursue astronomical research at Armagh Observatory. 8 References Arlt R., Rendtel J. and Bader P. 2008. The 2007 Orionids from visual observations. WGN (Journal of the International Meteor Organisation) 36: 55-60 Asher D. J. and Emel'yanenko V. V. 2002. The origin of the June Bootid outburst in 1998 and determination of cometary ejection velocities. Monthly Notices of the Royal Astronomical Society 331: 126-132 Asher D. J., Bailey M. E. and Emel'yanenko V. V. 1999. Resonant meteoroids from Comet Tempel-Tuttle in 1333: the cause of the unexpected Leonid outburst in 1998. Monthly Notices of the Royal Astronomical Society 304: L53-56 Bailey M. E. and Emel'yanenko V. V. 1996. Dynamical evolution of Halley-type comets. Monthly Notices of the Royal Astronomical Society 278: 1087-1110 Central Bureau Electronic Telegram No. 944 dated 2007 Apr 25 (http://www.cbat.eps.harvard.edu/iau/cbet/000900/CBET000944.txt) Chambers J. E. 1999. A hybrid symplectic integrator that permits close encounters between massive bodies. Monthly Notices of the Royal Astronomical Society 304: 793-799 Christou A. A., Vaubaillon J., and Withers P. 2008. The P/Halley Stream: Meteor Showers on Earth, Venus and Mars. Earth, Moon, and Planets, 102: 125-131 Denning W. F. 1899. Meteoric showers in autumn, winter, and spring from Ursa Major and the region near. The Observatory 22: 90-91 Dubietis A. 2003. Long-term activity of meteor showers from Comet 1P/Halley. WGN (Journal of the International Meteor Organisation) 31: 43-48 Emel'yanenko V. V. 1988. Meteor-stream motion near commensurabilities with Jupiter. Soviet Astronomy Letters 14: 278-281. Emel'yanenko V. V. 2001. Resonance structure of meteoroid streams. In Proceedings of the Meteoroids 2001 Conference (ESA SP -- 495), Edited by Warmbein B. Noordwijk: ESA. pp. 43-45. Everhart E. 1985. An efficient integrator that uses Gauss-Radau spacings. In Proceedings of IAU Colloq. 83, Edited by Andrea Carusi and Giovanni B. Valsecchi. Astrophysics and Space Science Library 115: 185-202 Giorgini J.D., Yeomans D.K., Chamberlin A.B., Chodas P.W., Jacobson R.A., Keesey M.S., Lieske J.H., Ostro S.J., Standish E.M., and Wimberly R.N. 1996. JPL's On-Line Solar System Data Service, Bulletin of the American Astronomical Society 28(3), 1158 16 Hajduk A. 1986. Meteoroids from Comet Halley and the comets mass production and age. In ESA Proceedings of the 20th ESLAB Symposium on the Exploration of Halley's Comet, 2: 239-243 Halley E. 1705. Synopsis Astronomiae Cometicae. London Herschel A. S. 1866. Radiant points of shooting stars. Monthly Notices of the Royal Astro- nomical Society 26: 51-53 Hughes D. W. 1985. The size, mass, mass loss and age of Halley's comet. Monthly Notices of the Royal Astronomical Society 213, 103-109 Hughes D.W. and McBride N. 1989. The mass of meteoroid streams. Monthly Notices of the Royal Astronomical Society 240: 73-79 Imoto S. and Hasegawa I. 1958. Historical Records of Meteor Showers in China, Korea, and Japan. Smithsonian Contributions to Astrophysics 2: 131-144 International Meteor Organisation Records (http://www.imo.net/zhr) Jenniskens P. 2006. Meteor showers and their parent comets. Cambridge, UK: Cambridge University Press. Jenniskens P., Lyytinen E., Nissinen M., Yrjola I., and Vaubaillon J. 2007. Strong Ursid shower predicted for 2007 December 22. WGN (Journal of the International Meteor Organisa- tion) 35: 125-133. Kepler J. 1609. Astronomia Nova. Heidelberg Kepler J. 1619. Harmonices Mundi Libri V. Linz. Kero J., Szasz C., Nakamura T., Meisel D.D., Ueda M., Fujiwara Y., Terasawa T., Miyamoto H. and Nishimura K. 2011. First results from the 2009-2010 MU radar head echo observation programme for sporadic and shower meteors: the Orionids 2009. Monthly Notices of the Royal Astronomical Society 416: 2550-2559 Kondrat'eva E.D. and Reznikov E.A. 1985. Comet Tempel-Tuttle and the Leonid meteor swarm. Solar System Research 19: 96-101. Kozai Y. 1979. Secular perturbations of asteroids and comets. In Dynamics of the solar sys- tem. Proceedings IAU Symposium 81. Edited by Duncombe R.L. Dordrecht: Reidel Publishing Co., pp. 231-236 Kresak L. 1987. The Evolution of the Small Bodies of the Solar System. Proceedings of the International School of Physics "Enrico Fermi". Edited by M. Fulchignoni, and L. Kresak. Amsterdam: Elsevier. pp. 202-216 Kronk G. W. 1988. Meteor Showers: a descriptive catalog. Hillside, NJ: Enslow Publishers. Lindblad B. A. and Porubcan V. 1999. Orionid Meteor Stream. Contributions of the Astro- nomical Observatory of Skalnat´e Pleso 29: 77-88 Lovell A. C. B. 1954, Meteor Astronomy. Oxford: Clarendon Press. 17 Marsden B. G. and Williams G. V. 2008. Catalogue of Cometary Orbits, 17th ed. Cambridge, MA: Minor Planet Center/Central Bureau for Astronomical Telegrams. Marsden B.G., Sekanina Z. and Yeomans D.K. 1973. Comets and non-gravitational forces. V. The Astronomical Journal 78: 211-225 McIntosh B. A. and Hajduk A. 1983. Comet Halley meteor stream - A new model. Monthly Notices of the Royal Astronomical Society 205: 931-943 McIntosh B. A. and Jones J. 1988. The Halley comet meteor stream- Numerical modelling of its dynamic evolution. Monthly Notices of the Royal Astronomical Society 235: 673-693 Millman P. M. 1936. Meteor News (Observation of the Orionids in 1936). Journal of the Royal Astronomical Society of Canada 30: 416-418 Miskotte K. 1993. High Orionid activity on October 18, 1993. WGN (Journal of the Inter- national Meteor Organisation) 21: 292 Murray C.D. and Dermott S.F. 1999. Solar system dynamics. Cambridge, UK: Cambridge University Press. Newton I. 1687. Philosophiae Naturalis Principia Mathematica. London. Olivier C.P. 1921. Parabolic orbits of meteor streams. Publications of the Leander Mc- Cormick Observatory 2: 201-268 Rendtel J. 2007. Three days of enhanced Orionid activity in 2006 - Meteoroids from a resonance region?. WGN (Journal of the International Meteor Organisation) 35: 41-45 Rendtel J. 2008. The Orionid Meteor Shower Observed Over 70 Years. Earth Moon, and Planets, 102: 103-110 Rendtel J. and Betlem H. 1993. Orionid meteor activity on October 18, 1993. WGN (Journal of the International Meteor Organisation) 21: 264-268 Ryabova G. 2003. The comet Halley meteoroid stream: just one more model. Monthly Notices of the Royal Astronomical Society 341: 739-746 Ryabova G. O. 2006. Meteoroid streams: mathematical modelling and observations. In Asteroids, Comets, Meteors, Proceedings IAU Symposium 229, edited by Lazzaro D., Ferraz- Mello S., Fern´andez J. A. Cambridge, UK: Cambridge University Press. pp. 229-247. Saha P. and Tremaine S. 1993. The orbits of the retrograde Jovian satellites. Icarus 106: 549-562 Sato M. and Watanabe J. 2007. Origin of the 2006 Orionid Outburst. Publications of the Astronomical Society of Japan 59: L21-L24 Soja R. H., Baggaley W. J., Brown P., and Hamilton D. P. 2011. Dynamical resonant structures in meteoroid stream orbits. Monthly Notices of the Royal Astronomical Society 414: 1059 -- 1076. Spurny P. and Shrbeny L. 2008. Exceptional Fireball Activity of Orionids in 2006. Earth, Moon, and Planets 102: 141-150 18 Steel D. I. 1987. The dynamical lifetime of comet P/Halley. Astronomy and Astrophysics 187: 909-912 Trigo-Rodriguez J.M., Madiedo J.M., Llorca J., Gural P.S., Pujols P., Tezel T. 2007. The 2006 Orionid outburst imaged by all-sky CCD cameras from Spain: meteoroid spatial fluxes and orbital elements. Monthly Notices of the Royal Astronomical Society 380: 126-132 Vaubaillon J., Lamy, P., and Jorda, L. 2006. On the mechanisms leading to orphan meteoroid streams. Monthly Notices of the Royal Astronomical Society 370: 1841-1848. Whipple F. L. 1951. A Comet Model. II. Physical Relations for Comets and Meteors. The Astrophysical Journal 113: 464-474 Whipple A. L. and Shelus P. J. 1993. A secular resonance between Jupiter and its eighth satellite?. Icarus 101: 265-271 Williams I. P. 2011. The origin and evolution of meteor showers and meteoroid streams. Astronomy and Geophysics 52: 2.20-2.26 Yeomans D. K. and Kiang T. 1981. The long-term motion of comet Halley. Monthly Notices of the Royal Astronomical Society 197: 633-646 Zhuang T. S. 1977. Ancient Chinese reports of meteor showers. Chinese Astronomy 1: 197-220 19
1906.02697
1
1906
2019-06-06T16:50:20
Simulated Phase-dependent Spectra of Terrestrial Aquaplanets in M Dwarf Systems
[ "astro-ph.EP" ]
Orbital phase-dependent variations in thermal emission and reflected stellar energy spectra can provide meaningful constraints on the climate states of terrestrial extrasolar planets orbiting M dwarf stars. Spatial distributions of water vapor, clouds, and surface ice are controlled by climate. In turn, water, in each of its thermodynamic phases, imposes significant modulations to thermal and reflected planetary spectra. Here we explore these characteristic spectral signals, based on 3D climate simulations of Earth-sized aquaplanets orbiting M dwarf stars near the habitable zone. By using 3D models, we can self-consistently predict surface temperatures and the location of water vapor, clouds, and surface ice in the climate system. Habitable zone planets in M dwarf systems are expected to be in synchronous rotation with their host star and thus present distinct differences in emitted and reflected energy fluxes depending on the observed hemisphere. Here we illustrate that icy, temperate, and incipient runaway greenhouse climate states exhibit phase-dependent spectral signals that enable their characterization.
astro-ph.EP
astro-ph
Affiliations: Title: Simulated phase dependent spectra of terrestrial aquaplanets in M-dwarf systems Authors: E.T. Wolf1,2,4, R.K. Kopparapu2,3,4,5, J. Haqq-Misra2,5 1University of Colorado, Boulder, Laboratory for Atmospheric and Space Physics, Department of Atmospheric and Oceanic Sciences, Boulder CO, USA 2Virtual Planetary Laboratory, Seattle WA, USA 3NASA Goddard Space Flight Center, Greenbelt MD, USA 4Sellers Exoplanet Environments Collaboration, NASA Goddard Space Flight Center, Greenbelt MD, USA 5Blue Marble Space Institute of Science, Seattle WA, USA Abstract: Orbital phase dependent variations in thermal emission and reflected stellar energy spectra can provide meaningful constraints on the climate states of terrestrial extrasolar planets orbiting M-dwarf stars. Spatial distributions of water vapor, clouds, and surface ice are controlled by climate. In turn water, in each of its thermodynamic phases, imposes significant modulations to thermal and reflected planetary spectra. Here, we explore these characteristic spectral signals, based on 3D climate simulations of Earth-sized aquaplanets orbiting M-dwarf stars near the habitable zone. By using 3D models, we can self-consistently predict surface temperatures and the location of water vapor, clouds, and surface ice in the climate system. Habitable zone planets in M-dwarf systems are expected to be in synchronous rotation with their host star and thus present distinct differences in emitted and reflected energy fluxes depending on the observed hemisphere. Here, we illustrate that icy, temperate, and incipient runaway greenhouse climate states exhibit phase dependent spectral signals that enable their characterization. 1. Introduction Planets in M-dwarf star systems provide our best chance for characterizing habitable terrestrial exoplanetary atmospheres in the next decade. Observing planets orbiting M- dwarf stars has several key positive biases favoring observation. M-dwarf stars are small and dim. Their low stellar luminosities and small radii mean that planetary emitted and reflected radiances will be relatively easier for humans to observe, compared to those from an Earth-like planet orbiting a Sun-like star. Habitable zone planets around M-dwarfs must also orbit quite close to their host stars, and thus their orbital periods will be much less than an Earth-year, allowing for observations to be feasibly stacked over many orbits within nominal mission lifetimes. Habitable extrasolar planets are assumed to have significant amounts of liquid water on their surfaces and thus in contact with the atmosphere. This provides a quantifiable constraint on the temperature regimes and climate states allowable for habitable worlds. Nominally, the surface temperature of a habitable planet must be maintained between the freezing and boiling points of water, however, locally surface temperatures may violate these limits. Broadly surface habitable worlds generally must maintain global mean surface temperatures somewhere between 280 K and 310 K, or else they risk having widespread areas of severe temperature extremes (Wolf et al. 2017). The presence of water vapor, perhaps uniquely among atmospheric constituents, acts as an inherently destabilizing force in the climate system. In its vapor phase, water-vapor is both a strong greenhouse gas, absorbing thermal emission by the planet, and also a significant absorber of stellar radiation at near infrared wavelengths. Both of these features of water vapor spectroscopy, combined with the temperature dependence of the Clausius- Clapeyron relationship, accelerate temperature perturbations (i.e. a positive climate feedback). The thermal absorption component is commonly referred to as the water-vapor greenhouse feedback. The near-infrared stellar absorption component also represents a positive feedback, and is greatly accentuated for planets around M-dwarf stars, given their low effective temperatures and thus peak emission in the near-infrared region of stellar spectra (Kopparapu et al. 2013). Furthermore, water also readily condenses at typical atmospheric temperatures and pressures for habitable worlds, forming clouds, sea-ice, and snow, all of which can add significantly to the planetary albedo. In particular, the temperature dependence of sea-ice and snow surface coverage also accelerates temperature perturbations, through the sea-ice albedo feedback. However, the spectral reflectivity of water ice and snow decreases significantly into the near-infrared which mutes the sea-ice albedo feedback for planets around M-dwarf stars (Joshi et al. 2011; Shields et al. 2013; Shields et al. 2014). Clouds have competing effects depending on their location. Liquid water clouds that form in the boundary layer (i.e. stratus clouds) contribute greatly to the planetary albedo, however ice water clouds that form in the upper troposphere (i.e. cirrus clouds) contribute to the greenhouse effect while remaining optically thin to incoming stellar radiation. Due to their close proximity to the host star, habitable zone planets around late-K and M-dwarf stars are expected to be tidally locked into synchronous rotation states (Kasting et al. 1993). For a synchronously rotating planet, its rotation rate equals its orbital period, and the same side of the planet always faces the host star. Numerous climate models have explored these novel atmospheric states for terrestrial extrasolar planets (e.g. Joshi et al. 1997; Merlis & Schneider 2010; Yang et al. 2013; 2014; Hu & Yang 2014; Shields et al. 2016, 2018; Kopparapu et al 2016, 2017; Fujii et al. 2017; Turbet et al. 2017, 2018; Haqq-Misra et al. 2018; Del Genio et al. 2019; Komacek & Abbot 2019; Yang et al. 2019; Adams et al. 2019). The Coriolis force plays a central role in modulating the atmospheric circulation, and the assumption of synchronous rotation allows precise constraints of planetary rotation rates. For slowly rotating planets (𝑃≲20 Earth days), the severe weakening of the Coriolis force allows for a thermally direct circulation, with strong rising motions on the permanent day- side and sinking air on the permanent night-side. Such a circulation tends to produce thick reflective clouds on the substellar hemisphere, greatly increasing the planetary albedo, and stabilizing the climate against a runaway greenhouse under relatively high total stellar irradiances (Yang et al. 2013). However, the longitudinal extent of the substellar clouds is sensitive to the specific rotation rate. As the planetary rotation rate increases for habitable zone planets around late M-dwarfs, strengthening Coriolis forces drive increased zonal motions that advect clouds eastward around the planet (Kopprapu et al. 2017; Haqq-Misra et al. 2018). Water vapor, clouds, and ice are variable constituents in the climate system of habitable planets, while other typical gas species (i.e. N2, CO2) tend to be well-mixed. Here, we postulate that the spatial variability of water in its various thermodynamic phases, coupled with its strong interaction with thermal and stellar radiation, provides an avenue for atmospheric characterization through phase dependent observations. Broadband phase curves and phase dependent spectra may provide an opportunity to characterize atmospheric circulation, heat transport, and cloud formation in planetary atmospheres, beyond what may be capable with transit and thermal emission spectroscopy alone (Koll & Abbot 2016). Phase dependent information captures multi-dimensional process in planetary atmospheres, which are particularly relevant for tidally locked planets around M- dwarf stars, where day-night asymmetries in climate and atmospheric structure are much more likely to occur (e.g. Joshi 2003; Yang et al. 2013; Koll & Abbot 2016; Haqq-Misra et al. 2018). Along with increasing the albedo, the presence of thick substellar clouds on slowly rotating planets can also significantly reduce the outgoing longwave radiation from the substellar hemisphere due to the greenhouse effect of high altitude clouds (Yang et al. 2013). For habitable planets, thermal emission phase curve morphology is controlled by emission from the planet surface and its interaction with water-ice clouds in the atmosphere, which are controlled by atmospheric circulations in the upper troposphere/lower stratosphere region (Haqq-Misra et al. 2018). This stands in contrast to the case of hot Jupiters, where thermal emission phase curve morphology is thought to be controlled by direct stellar heating of the atmosphere and subsequent advection of heat eastward by broad equatorial superrotating jets (e.g. Knutson et al. 2007; Showman & Polvani 2011). Here, we analyze the phase dependence of thermal emission and reflected stellar energy from Earth-sized ocean covered planets. We examine both broadband fluxes and spectral radiances. Our focus is in how phase dependent planetary energy signals are modulated by water vapor, liquid water clouds, ice water clouds, and surface ice. Planetary energy fluxes and spectra are determined based on a suite of 3D theoretical models of Earth- sized aquaplanets located in and near the habitable zones of late-K and M-dwarf stars. We use the habitable zone planet models of Kopparapu et al. (2017), along with complimentary simulations that extend to lower stellar fluxes and thus to colder planetary temperatures. This paper marks the third in a series of papers on 3D climate modeling of terrestrial planets around M-dwarf stars. Kopparapu et al. (2017) describes the climatology and habitability constraints of planets approaching the moist and runaway greenhouse limits. Haqq-Misra et al. (2018) describes the atmospheric circulation patterns as a function of rotation rate and orbital distance. This work focuses on thermal and reflected phase curves from these worlds, and how phase curve information may be used to characterize terrestrial exoplanetary atmospheres in M-dwarf systems. This manuscript is organized as follows. In section 2 we outline our methodology. In section 3 we briefly summarize climate modeling results. In section 4 we describe broadband thermal emission and broadband planetary albedo phase curves. In section 5 we describe phase dependent thermal emission and reflection spectra. In section 6 we describe the variation spectra. Sections 7 and 8 contain discussions and conclusions respectively. 2. Methods We follow the methods used in the general circulation modeling study of the inner edge of the habitable zone for synchronous rotators by Kopparapu et al. (2017) except where noted. Here we give a brief summary of our modeling methods, but refer the reader to section 2 in Kopparapu et al. (2017) and Neale et al. (2010) for progressively in-depth model descriptions. We use a modified version of the Community Atmosphere Model (CAM) version 4 from the National Center for Atmospheric Research in Boulder, CO (Neale et al. 2010). Specific code changes, including radiative transfer, model configurations, and initial conditions files needed to facilitate exoplanet studies with CAM, are available via the ExoCAM and ExoRT code packages, which we have made publicly available at the lead authors' Github page1. In this work, planets are assumed to be Earth-sized, completely ocean 1 https://github.com/storyofthewolf covered, and with atmospheres composed of 1 bar N2 and variable amounts of water vapor, liquid water clouds, and ice water clouds. There is no CO2, CH4 or other non-condensable (at Earth-like temperatures) greenhouse species in these simulations. Water vapor and clouds are prognostically determined using the standard parameterizations of CAM version 4 (Neale et al. 2010; see also Kopparapu et al. 2017, section 2.3). For simplicity and computational efficiency, the ocean is treated as a 50-meter deep "slab" ocean with no ocean heat transport (i.e. q-fluxes are set to zero everywhere). Sea ice forms where sea-surface temperatures fall below the freezing point of sea-water at 271.36 K. We assume the default snow and ice albedo parameterization, which divides ocean, sea ice, and snow albedos into two bands: visible and near-IR divided at ~0.7 μm. Following Shields et al. (2013), we set the visible (near-IR) sea-ice albedo to 0.67 (0.3), the snow albedo to 0.8 (0.68), and the ocean albedo to 0.07 (0.06), respectively. This two-band parameterization of surface albedos is hard-wired within sea-ice, snow, and land modules respectively, and is not easily changed to accommodate multiple bands. A band resolved surface albedo parameterization would surely improve its treatment (e.g. Joshi et al. 2003), however this is left for future work. Simulations were run with 4° ´ 5° horizontal resolution and 40 vertical layers extending from the surface up to ~1 mb pressures, using a finite volume dynamical core (Lin & Rood, 1996). The radiative transfer module considers H2O line and continuum absorption and N2- N2 collision-induced absorption (CIA). Correlated-K absorption coefficients are derived for H2O using the HITRAN 2012 database (Rothman et al. 2013) and HELIOS-K, an ultrafast open-source spectral sorting program that runs on graphics processing units (Grimm & Heng 2015). Water vapor continuum is treated using the formalism of Paynter & Ramaswamy (2011). N2-N2 CIA is included from Borysow & Frommhold (1986). Cloud overlap is treated using the Monte Carlo Independent Column Approximation assuming maximum random overlap (Pincus et al 2003). In this study, the radiative transfer module is configured with 68 spectral intervals across the whole spectra, each with 8 gauss points per spectral interval. Note, here we have increased the spectral resolution compared to Kopparapu et al. (2017), which used only 42 spectral intervals. Differences in the radiative flux calculations between the 42 and 68 bin versions are less than ~1% at worst for moist greenhouse atmospheres, and introduce no discernable differences into the mean states of the climate simulations. Here, we have chosen to use the higher resolution radiative transfer in order to extract higher quality synthetic spectra from our simulated worlds. Longwave and shortwave radiative streams share the same spectral interval grid, and were computed each over the entire available spectrum. We have modified our 3D model to directly output spectrally resolved fluxes. We simulated planets around stars with stellar effective temperatures, Teff, of 2600 K, 3000 K, 3300 K, 3700 K, 4000 K, and 4500 K, assuming BT-SETTL spectra (Allard et al. 2007), with stellar metallicities of [Fe/H] = 0.0, and log g = 4.5 (Figure 1). The planets are assumed to be in synchronous rotation. The incident stellar fluxes and orbital periods (and thus planetary rotation rates) are calculated with consistency to Kepler's laws. For stars with Teff ≥ 3300 K, we use the empirically derived stellar mass-luminosity relationships from Boyajian et al. (2013). For cooler stars we use stellar mass-luminosity relationships derived from stellar evolutionary models (Baraffe et al. 2002), assuming an age of ~3 Gyr. We have duplicated the simulation grid of Kopparapu et al. (2017), and have also included two new simulations per star, with 680 Wm-2 (0.5 S0) and 1080 Wm-2 (0.8 S0) incident stellar fluxes respectively. We included these new simulations to the grid in order to expand our sampling of theoretical climate states to cold worlds. In total, around each stellar type, we have conducted simulations of terrestrial planets with global mean surface temperatures (Ts) varying from ~205 K up through the point where a runaway greenhouse is triggered. Based on our suite of 3D model simulations, we then compute shortwave and longwave phase curves and phase dependent spectra directly from 3D GCM spectrally resolved outputs, using the methods described by Koll and Abbot (2015, Appendix C) for calculating the disk-averaged and observer-projected fluxes from a planet, assuming an edge-on view relative to the observer. Because our 3D model directly outputs spectrally resolved information, no off-line radiative calculations are needed. For climatologically stable simulations, phase curves and phase dependent spectra are calculated based on averages of the last 5 Earth-years of simulation. Our testing has found that a 5 year duration for a temporal average is sufficient for removing bias from variability. For incipient runaway greenhouse cases, as described in Kopparapu et al. (2017, section 3.3), temperatures rapidly increase and the model becomes numerically unstable when model temperatures approach ~400 K within ~10 Earth-years time in most cases. For incipient runaway greenhouse cases, we calculate phase curves based on the last 90 Earth-days of simulation before the numerical instability occurs. This short duration for a temporal average is chosen as to sample the hottest part of the simulation, but as a consequence retains some variability due to weather. Here, we limit the scope of our focus to planets that are also transiting, thus variations in the observed phase curves caused by observer obliquity are not considered here. Planets that are transiting will provide the best targets, allowing us to combine information between 3. Summary of Climate Models thermal and reflected phase curves, transit spectroscopy, secondary eclipse mapping, and radial velocity methods. Most of the climate models used here were first presented in Kopparapu et al. (2017), with continued analysis in Haqq-Misra et al. (2018). Here we have expanded our simulation grid to include lower stellar fluxes (0.5 S0 and 0.8 S0), in order to include cold and icy states in our ensemble of climates. We refer the reader to Kopparapu et al. (2017) for an in-depth discussion of planetary climate, clouds, and moist and runaway greenhouse processes occurring near the inner edge of the habitable zone. We refer the reader to Haqq-Misra et al. (2018) for an in-depth discussion of atmospheric dynamical regimes that exist on these tidally locked worlds as a function of changing planetary rotation rate. In this section we summarize the climate modeling results that are most relevant for interpreting phase dependent thermal emission and reflected stellar energy spectra from these worlds. We first divide climate into four different categories, sorted by their global mean surface temperature, Ts. These categorizations will be referenced throughout the paper. Cold climates have 200 K < Ts < 250 K. While these worlds are very cold and their surfaces are dominated by sea ice, all still maintain a small fraction (4% - 25%) of open-ocean at the substellar point and thus are habitable based on the standard criteria requiring surface liquid water to exist anywhere on the planet surface. Cool climates have 250 K < Ts < 273 K. These climate states have open-ocean fractions up to ~50%, thus their substellar hemispheres have little to no ice, while their antistellar hemispheres are permanently ice- covered. Temperate climates have 273 K < Ts < 310 K and represent the most optimistic cases for habitable planets, dominated by open-oceans and largely with modern Earth-like surface temperatures. Finally, incipient runway states are those where a runaway greenhouse instability has been triggered. As described in Kopparapu et al. (2017, section 3.3), a strong energy imbalance is present and model temperatures rapidly increase, however our climate model becomes numerically unstable as temperatures near ~400 K. Simulating a fully realized runaway greenhouse, with atmospheric temperatures beyond the critical point of water, is beyond the operational bounds of our code and remains a significant challenge for 3D climate modelers (e.g. Ishiwatari et al. 2002; Kane et al 2018). A fully realized runaway greenhouse would evaporate the entirety of the oceans into its atmospheres, yielding ~270 bars of water vapor for every Earth ocean originally present (Kasting 1988), and would have a surface temperature of ≥1600 K in equilibrium (Goldblatt et al. 2013). Thus, our "runaway" cases represent a snapshot of atmospheres undergoing the runaway greenhouse process and are in a transient state, not a steady-state. Our presented runaway states fall well short of the extremes of the classical runaway greenhouse (Goldblatt & Watson 2012). With respect to observable phenomenon, our incipient runaway states are perhaps best considered as a sampling of hot, moist, and optically thick atmospheres (e.g. Goldblatt et al. 2015). To further sort our results, we divide atmospheric circulation regimes into three categories. For the remainder of the paper we will adopt the circulation regime nomenclature suggested by Haqq-Misra et al. (2018). "Slow" rotators have rotation periods (P) of >20 Earth-days, which occurs for habitable zone planets around stars with 3300 K £ Teff £ 4500 K. Slow rotators are characterized by a weak Coriolis force, strong rising motions at the subtellar point, cooling and descending air on the antistellar hemisphere, and radially symmetric day to night side heat transport. "Rapid" rotators have P < 5 Earth-days, which can occur only for habitable zone planets around the coolest of stars, Teff < 2600 K. Rapid rotators are characterized by a strong Coriolis force, weak upwelling, strong superrotation, and west to east zonal heat transport. Lastly, an intermediate dynamical regime, coined "Rhines" rotators by Haqq-Misra et al. (2018), features rotation rates of 5 £ P £ 20 Earth- days, which occurs for habitable zone planets around stars with 3000 K £ Teff £ 3300 K. Defined in terms of the Rhines length scale (Rhines, 1975; Showman et al. 2010), Rhines rotators are characterized by a relatively strong Coriolis force, strong upwelling, and moderate superrotation. While here we define discrete categorizations based on specific rotation periods, in actuality atmospheric circulation evolves continuously as a function of planet rotation. Note, also it has been suggested that bistability in atmospheric circulation regimes could occur for terrestrial sized planets with rotation rates of less than ~10 days depending upon assumed boundary conditions (Carone et al. 2018), the possibilities for which we have not yet explored. Here, we show maps of surface temperature (K, Figure 2), vertically integrated water vapor column (kg m-2, Figure 3), vertically integrated liquid cloud water path (g m-2, Figure 4), and vertically integrated ice cloud water path (g m-2, Figure 5). Figures 2, 3, 4, and 5 have the same format. Each column shows illustrative cases for a cold planet (~200 K), a cool planet (~260 K), a temperate planet (~300 K), and a hot, moist planet where a transition to a runaway greenhouse has been triggered (~380 K). Each row corresponds to a different host star, with stellar effective temperatures ranging from 2600 K to 4500 K. The stellar effective temperature, range of orbital periods, and range of incident stellar fluxes are listed in the left-hand margin. Naturally, cold planets are furthest from their host star and have the longest orbital periods and lowest received stellar fluxes. Planets where a runaway greenhouse has been triggered are closest to their host star, and have the shortest orbital periods, and highest received stellar fluxes. Recall, that all of the worlds simulated are assumed to be in synchronous rotation, thus their orbital period equals the planetary rotational period. To first order, surface temperature maps vary as expected, with climate warming as the stellar flux is increased (Figure 2). Note that day-to-night heat transport becomes more efficient as the climate warms, and thus day-night temperature gradients are reduced for increasingly warm states. For incipient runaway greenhouse states, the atmosphere is optically thick and the surface temperature varies little between substellar and antistellar hemispheres for all rotation rates. For cold climates a sharp day-night temperature gradient is maintained as the substellar point retains surface liquid water, while the night-side remains frigid. Slow rotating planets (Teff ≥ 3300 K, P ≥ 20 days) generally maintain surface temperature distributions that are radially symmetric about the substellar point. However for planets with faster rotation rates, found for habitable zone planets around low mass stars, increased zonal transport results in the emergence of longitudinally banded patterns. The water vapor column also varies as expected with warming climate (Figure 3), with the amount of water vapor that the atmosphere can contain increasing exponentially as a function of temperature, as determined by the Clausius-Clapeyron relation. Recall, here we have assumed completely ocean-covered planets, which thus provide an unlimited source of water into the atmosphere. Cold planets have only a maximum of < 10 kg m-2 of water vapor in their atmospheres centered over a small substellar patch of open-ocean, and virtually no water vapor on their ice-covered night sides. On the other hand, our incipient runaway greenhouse cases have 5000 to 10,000 kg m-2 in their atmospheres at the point of model termination. Cool climates may have up to ~75 kg m-2 and temperate climates have up to several hundred kg m-2 of water vapor in their atmospheres. Note water vapor distributions are generally symmetric about the substellar point for slow rotators, but begin to take on zonal patterns due to eastward advection that occurs for faster rotators. In Figures 4 and 5 we show liquid water cloud and ice water cloud paths respectively. Note that clouds are shown in white. While surface temperatures and the water vapor column behave predictably and monotonically with increasing incident stellar flux, the behavior of clouds is more complex. Liquid water clouds are sparse for cold climates, because these atmospheres generally remain cold and dry. However, cold climates have the highest concentrations of ice water clouds. However, unlike for Earth, where ice clouds (i.e. cirrus) are only formed high in the atmosphere and thus contribute to the greenhouse effect, here for cold climate ice clouds exist near the surface and thus do not imply a significant greenhouse effect. Liquid water clouds are thickest for cool and temperate climates. For slow rotating planets (Teff ≥ 3300 K) thick clouds decks form around the substellar hemisphere (e.g. Yang et al. 2013). However, note that the for planets around low mass stars, the increased planetary rotation rates for habitable zone planets results in a stronger Coriolis force, which drives zonal flows (Haqq-Misra et al. 2018) and correspondingly advects clouds eastward from the substellar hemisphere (Kopparapu et al. 2017). Incipient runaway greenhouse planets tend to have fewer clouds on the substellar hemisphere. Strong stellar irradiance and strong absorption in the near-infrared by water vapor heats the atmosphere significantly, lowering relative humidities, and resulting in fewer clouds in the substellar hemisphere despite overwhelming water vapor burdens (Kopparapu et al. 2017). Liquid water clouds are generally confined to the marginally cooler night-side, and ice- clouds are few. Maps of liquid and ice cloud water paths are critical for the interpretation of thermal emitted and reflected stellar phase curves, as will be made evident in the following sections. 4. Broadband Phase Curves 4.1. Thermal Emission First, we compute broadband thermal emission phase curves based on all simulations conducted (Figure 6). Each panel shows top-of-atmosphere thermal emission phase curves (Wm-2) integrated across the entire spectrum, for all simulations conducted around each star respectively. Stellar effective temperatures, Teff, are labeled in the upper left-hand corners of each panel. Note the illustration of phases at the top of the figure. Thermal emission phase curves are color coded according to climate state, with relevant global mean temperature ranges also given. The solid lines indicate stable climate states which have reached equilibrium. Dark blue lines indicate cold climates (200 K < Ts < 250 K). Light blue lines indicate cool climates (250 K < Ts < 273 K). Solid gold lines represent temperate climates (273 K < Ts < 310 K), our ideal habitable scenarios. The dashed red lines are for incipient runaways, where a transition to a runaway greenhouse has been triggered, and provide a representation of a hot and moist atmosphere (see Section 3). Planets in slow, Rhines, and rapid rotating circulation regimes exhibit fundamentally different climate state dependent morphologies of their broadband thermal emission phase curves. Slow and Rhines rotators (i.e. Teff ³ 3000 K) exhibit an inversion of their thermal phase curve amplitudes as the climate warms from cold to temperate states. As noted first by Yang et al. (2013), temperate synchronously rotating planets tend to have efficient day- night heat redistribution (Figure 2), coupled with thick clouds on the substellar hemisphere that extend high into the troposphere (Figures 4, 5). The strong greenhouse effect provided by these high substellar ice clouds creates a minimum in the observed thermal emission from the planet observed at secondary eclipse (i.e. viewing the day-hemisphere). The night-side of the planet remains cloud free and thus can emit thermal radiation to space originating from much lower and thus warmer layers of the atmosphere. Conversely, for cold planets, day-night temperature gradients are large with the night side remaining icy and very cold, but with a small patch of open-ocean confined to the substellar point (Figure 2). On cold planets, cloud coverage is sparse and confined to the boundary layer, where their emitting temperature remains near to that of the surface (Figure 4, 5). The result is that for cold planets the thermal emission phase curve tends to have a maximum at secondary eclipse corresponding with when the small patch of open-ocean at the substellar point comes into view. Cool planets, represent an in between case, and an additional mode in their broadband thermal phase curves is evident. On these worlds, clouds are only concentrated within ~30° of the substellar point, resulting in a local minimum at the substellar point, and emission maxima observed near western and eastern quadrature respectively. The inversion of the thermal phase curve with warming climate is predicted for planets around stars with Teff ³ 3000 K, including both slow and Rhines rotators. However, this trend is not predicted for rapidly rotating planets that are found in the habitable zone of 2600 K stars. Around such cool stars, all climate states yield similar thermal phase curve morphologies, with maximum emission observed near secondary eclipse. The thermal emission minimum is generally found at eastern quadrature. For rapidly rotating planets, the Coriolis effect becomes important and increases the strength of the zonal circulation. Here, superrotating winds advect clouds eastward of the substellar point, limiting their reflectivity at the substellar point and confining ice clouds to reside near the eastern terminator (Figure 5). An additional mode of variation in thermal emission phase curves involves the phase offset of maximum and minimum emission. Slow rotators tend to maintain strict day-night symmetry of thermal emission phase curves for cold, cool, and temperate climate states, holding true even as the phase curve inverts for warm climates as described above. Thermal emission phase curves for cold climates remain symmetric for all circulation regimes, because clouds are few and emission is dominantly from the surface, whose energy budget is defined simply by the solar insolation and surface albedo. However, Rhines and fast rotators exhibit symmetry breaking of their thermal emission phase curves for warmer climates. As climates warm, clouds take a leading order importance in the determination of thermal emission. For rapid rotators, clouds, and in particular the location of high ice clouds (Haqq-Misra et al. 2018), are shifted significantly eastward of the substellar point, causing the thermal emission minimum to be observed nearer to eastern quadrature. Interestingly, for Rhines rotators, the high ice clouds are shifted westward, causing the thermal emission minimum to be observed nearer to western quadrature (Figure 5). Lastly, incipient runaway simulations represent an outlier case, which do not conform with trends deduced from stable climate states. We find, similar to other recent studies (e.g. Leconte et al. 2013; Kane et al. 2019), that as a runaway greenhouse is triggered the atmospheric relative humidity and cloudiness decline, despite large specific humidities. The low relative humidity ensures that the magnitude of thermal emission to space generally exceeds the theoretical radiation limit from saturated atmospheres of 282 Wm-2, determined by Goldblatt et al. (2013) from line-by-line calculations. The large water vapor burden results in strong absorption of stellar near-infrared radiation (Figure 3). The day side becomes preferentially heated in the middle atmosphere, and then can effectively reradiate thermal energy back to space. Clouds in our incipient runaway cases tend to be few and are greater on the night-side of the planet (Figure 4, 5). Note that Earth presently has a mean outgoing longwave radiation of 239.7±3.3 Wm- 2 (Stephens et al. 2012). On Earth, the global mean outgoing longwave radiation varies by only ~4 Wm-2 (~1%) over monthly timescales due to cloud variability (i.e. weather) while interannual variability is even smaller (Harris and Belotti, 2010). There would be little inferred phase dependence of thermal emission from an Earth-twin planet orbiting a Solar- twin star. Cowan et al. (2012), using 3D model simulations of Earth from the ECHAM climate model, predict little more than a 2% variation in deduced thermal phase variations for an Earth-twin planet observed at a viewing obliquity of 0°. By contrast, the phase dependent variations in thermal emission is much larger for synchronously rotating habitable planets around M-dwarfs. Here, while global mean outgoing longwave radiation remains on the same order as Earth (~200 to 300 Wm-2), for synchronously rotating worlds the amplitude of variation can be up to ~70 Wm-2, or roughly 20% - 35% of the global mean. In Figure 7, we show broadband albedo phase curves based on all simulations conducted. Figure 7 is arranged and color coded identically as Figure 6. Similar to 4.2. Albedo broadband thermal phase curves, planets in different circulation regimes, and in different climate states, exhibit different morphologies of phase dependent albedo variation. Note, unlike thermal emission which emanates from all parts of the planet, albedo is only calculated from the illuminated fraction of the planet. At secondary eclipse the entire substellar hemisphere is illuminated relative to the observer. At crescent phases, only a small sliver of the planet near the terminator is illuminated relative to the observer. By definition the albedo at transit is undefined. Perhaps surprisingly, for slow rotating planets (Teff ³ 3300 K), we find that temperate and cool climates exhibit larger planetary albedos than do cold icy climates states, at all orbital phases except at crescent phases. Cold climate states generally have few clouds and are covered by sea ice, except for small regions confined to the substellar point. For cool and temperate climate states, the illuminated hemisphere has little or no surface ice, and has significant cloudiness enshrouding much of the substellar hemisphere (Figure 2, 4). Irradiated by the red spectra of M-dwarf stars (Figure 1), liquid water clouds have a greater broadband reflectivity than does sea-ice and snow (Shields et al. 2013). Thus, cloudy climates of habitable planets around M-dwarf stars will display higher albedos than will a water-ice covered "Snowball" world. For slow rotators, cold climates exhibit a strong increase in albedo at crescent phases. For cold climates, small areas of open-ocean along with cloud free conditions around the substellar point effectively suppress the disk-averaged albedos observed at secondary eclipse. However, at crescent phases, only ice-covered regions of the planet are visible to the observer, and thus exhibit a higher apparent albedo at these phase angles, compared to at secondary eclipse. For cool and temperate worlds around slow rotators, phase dependent albedos remain fairly uniform, except at western crescent phases. This may be attributed to a subtle decreases in cloudiness along western terminator, at trailing cusp of the substellar cloud deck relative to the direction of planetary rotation (Figure 4). As we move to smaller and cooler host stars, where Rhines and rapid rotation circulation regimes are present, we note changes to albedo phase curve morphologies. As planet rotation rate increases, eastward zonal flow strengthens and water clouds are advected to the east of the substellar point (Kopparapu et al. 2017; Haqq-Misra et al. 2018). This causes a tilting of the broadband albedo phase curve, with slightly higher albedos being observed at eastern quadrature and crescent phases for habitable planets around 3000 K, and 3300 K and cooler stars. Rapidly rotating planets found around 2600 K stars again are an outlier compared to the other cases. Opposite of the behavior seen for slower rotating planets, for rapidly rotating planets the albedo monotonically increases for progressively colder planets. This is because superrotation is strong and effectively obliterates the subtellar cloud deck. Clouds are quickly sheared downstream, away from the subtellar point (Figure 4). Thus for fast rotators, the planetary albedo at secondary eclipse is controlled primarily by the surface albedo, rather than by clouds as is the case for slower rotating circulation regimes. For temperate planets around 2600 K stars, the strong advection of clouds eastward results in highest albedos found at eastern quadrature and crescent phases. Increased albedos are also found at western crescent phase, due to strong weather systems which circle the planet, carrying moisture and clouds originating from the substellar hemisphere (Figure 3, 4). Around all stars, planets undergoing a runaway greenhouse process exhibit exceedingly low albedos. This is due to an obvious lack of sea-ice, a relative lack of clouds, and large amounts of water vapor in the atmosphere which readily absorbs stellar energy in the near-infrared. For stars with Teff ³ 3000 K, ensuing runaway greenhouse climate states may be easily discerned via their low albedos, particularly when viewed at secondary eclipse. Clouds are generally isolated to the terminators and night-side of the planet, where the free atmosphere is marginally cooler than the highly irradiated substellar hemisphere, allowing water vapor to condense into cloud (Figure 4, 5). Increases in the phase dependent albedo at crescent phases can be contributed to these clouds located near the terminators. For fast rotators around 2600 K stars, the differentiation between runaway and temperate climates is less obvious, as albedos in both cases remain low and of similar magnitude. Still phase dependent changes to the albedo may yield a tell-tale sign. For temperate planets, tropospheric water clouds are present but are advected eastward by strong supperotating winds, resulting in the highest reflection at eastern quadrature. However, for incipient runaway cases, the advection of heat eastward maintains high temperatures, low relative humidities, and few clouds in the eastern hemisphere. Clouds preferentially form in the subtly cooler western hemisphere, resulting in the highest reflection being observed at western crescent phases (Figure 4). Similar patterns of cloud formation in hot, supperrotating atmospheres, with more cloud mass found in the cooler western hemisphere, have also been noted in models of hot Jupiters (Parmentier et al. 2016). Finally, note that for all cases considered, there is a general trend that for a given climate state the planetary albedos are smaller around cooler stars, because the incident stellar light becomes increasingly shifted into the near-infrared (Figure 1), where both water vapor and sea ice are significant absorbers. 5. Phase Dependent Spectra 5.1. Thermal Emission Spectra In this section we shift our focus to spectrally resolved phase curves. In Figure 8 we show thermal emission spectra measured at the top-of-atmosphere, from cold, temperate, and ensuing runaway greenhouse climates states (columns), around each of our 6 stars (rows). Spectra are shown for the 4 primary phases; transit (dark blue), western quadrature (gold), secondary eclipse (red), and eastern quadrature (light blue). In all cases, we see negligible phase dependent variation in thermal emission within water vapor absorption bands. Phase dependent differences are small at long wavelengths (<15 µm) where water vapor rotational bands imply a relatively smooth absorption spectra, and they are also small coincident with the~6 µm water vapor absorption band. At even shorter wavelengths (<5 µm), the magnitude of phase dependent thermal emission differences is damped due to there being very little planetary emission. The most significant phase dependent spectral variations in thermal emission are found in the water vapor window region, which extends approximately between 8 - 13 µm. The water vapor window region is also coincident with maximum in thermal emission from terrestrial planets at roughly Earth-like temperatures. In the water vapor window region, the spectral signals of clouds and their variability are manifested most strongly, because there are no other absorbers in this spectral region. The presence and location of water clouds is driven to first order by the planet temperature and atmospheric circulation regime, and thus can yield clues as to the planet's climate state. Recall in our model, the only absorbers are water vapor, clouds, and N2-N2 CIA. N2-N2 CIA only weakly affects thermal radiation longward of ~20 µm and is likely swamped by H2O absorption anyway (Wordsworth & Pierrehumbert, 2013). Note, N2-N2 CIA also has an absorption feature at 4.3 µm which may be detectable at <10 part-per-million levels in transmission spectra (Schwieterman et al. 2015), however it lies at the shoulder of both stellar and planetary energy distributions and thus has minimal effects on reflection and emission spectra, and does not affect climate. Cold planets yield phase dependent thermal emission spectra that are quite similar, regardless of the host star and atmospheric circulation regime. For cold planets, there is very little water vapor and few clouds in the atmosphere, and thus the effects of constituent advection are less important. Thermal emission in the water vapor window region is dominated by emission from near the surface directly to space. Clouds that are present in cold cases are confined to close to the surface, and thus do not impose a significant greenhouse effect. The spatial distribution of surface temperature (and thus thermal emission) is controlled by the patterns of incident stellar energy and surface albedo, both of which are symmetric about the substellar point. In all cold cases, even with global mean temperatures dipping into the low 200s K, a small circular patch of open-ocean is present at the substellar point, with temperatures marginally above freezing and few clouds (Figure 2, 4, 5). All cold cases have the emission maximum observed at secondary eclipse when the open-ocean patch on the day-side is in view of the observer, and the emission minimum at transit when the icy cold night-side is in view of the observer. For cold worlds, phase dependent thermal emission spectra are generally symmetric, with spectra at eastern and western quadrature being identical in Figure 8. Temperate planets exhibit more complex changes to their phase dependent emission spectra. Still, the water vapor window region provides the optimal bandpass to observe variations in planetary emission spectra as a function of phase. For slow rotating planets (Teff ³ 3300 K) the emission maximum is observed at transit, because high ice clouds over the substellar hemisphere effectively trap thermal radiation, while thermal radiation readily escape to space through the clear-sky conditions on the night hemisphere (e.g. Yang et al. 2013; Haqq-Misra et al. 2018). Temperate slow rotators exhibit remarkable symmetry, with thermal emission at eastern and western quadrature being identical and overlapping nearly perfectly in Figure 8. For Rhines rotators (Teff = 3000 K and 3300 K), day-night differences in thermal emission spectra become muted. Differences, even in the water vapor window region, are small due to advection of both heat and clouds around the planet (Kopparapu et al. 2016; Kopparapu et al. 2017). Finally, for rapid rotators, maximum thermal emission through the water vapor window region is observed at secondary eclipse, followed closely by the emission observed at western quadrature. For fast rotators, clouds are strongly advected eastward, completely off of the substellar point into the eastern hemisphere. Emission from the warm planet surface, located immediately under and west of the substellar point, is able to escape directly to space. Ice cloud layers form only over the eastern terminator, reducing outgoing longwave radiation in those regions (Haqq-Misra et al. 2018). Incipient runaway cases tend to yield an emission maximum in the water vapor window region observed at secondary eclipse, due to intense stellar absorption by near- infrared water vapor absorption bands, strong middle atmosphere heating, and subsequent thermal emission to space. However, some variability in this trend is observed due to patchy and inconsistent clouds, probably due to the short duration and short temporal averaging period for these simulations. More strikingly incipient runaway cases may be discerned by the obvious thermal emission maximum also found in the ~4 µm water vapor window 5.2. Albedo Spectra region. This result is qualitatively similar to the results of Leconte et al. (2013) for runaway greenhouse states on highly irradiated water-limited land planets, and the 1D line-by-line calculations of Goldblatt et al. (2013). However, Goldblatt et al. (2013) predicts that thermal emission near ~4 µm is only possible when a planet reaches the terminal state of the runaway, with surface temperatures of ~1800 K coincident with the emitting layer of the atmosphere becomes hot enough to emit at ~4 µm. Note the ~4 µm emission peak shown in Goldblatt et al. (2013) is significantly larger than those found in our study. Goldblatt et al. (2013) assumed a saturated atmosphere. Both here and in the Leconte et al. (2013), the runaway cases simulated remain subsaturated despite atmosphere temperatures beyond ~400 K and large specific humidities. Presumably, if not for numerical limitations of our model, our incipient runaway greenhouse cases would continue to warm until excessively hot steam atmospheres are obtained, as depicted with 1D radiative transfer modeling of Goldblatt et al. (2013). As such, the thermal emission feature shown here at ~4 µm should continue to strengthen as the runaway process continues. However, the subsaturation of the atmosphere means that the ~4 µm feature may appear very earlier in the temporal evolution of a runaway greenhouse state, and also possibly for other warm water-rich worlds. Perhaps more pertinently, the emergence of a modest ~4 µm emission feature may allow us to discern between planets with surface temperatures of 300 K and ~400 K, providing an important observable constraint on planetary habitability. In Figure 9, we show phase dependent spectrally resolved albedos for cold, temperate, and incipient runaway cases (columns), around each of our 6 stars (rows). Figure 9 is arranged and color coded similarly to Figure 8, except that the albedo at transit is by definition undefined and thus omitted. Significant differences are found in the reflectance spectrum between climate states. Spectra are shown for 3 phases; western quadrature (gold), secondary eclipse (red), and eastern quadrature (light blue). In general, water vapor absorption features manifest as low-albedo regions of the albedo spectrum. Near-infrared water vapor window regions are controlled by a combination of cloud and surface reflectivities, and manifest as maxima in the albedo spectrum. Water vapor absorption is weak at visible wavelengths, thus visible albedos are also controlled by cloud and surface albedos, as well as Rayleigh scattering. For cold climates, reflection off of the surface dominates the albedo spectrum, because these atmospheres have relatively few clouds and sparse water vapor. In our model, there is a step-wise division between the visible and near-infrared snow and ice albedos at ~0.7 µm. This division is clearly evident in phase dependent albedo spectra, indicating that the surface dominates the albedo for these cold and relatively dry and cloud free worlds. For slow rotating (Teff ³ 3300 K) cold climates, there is a high degree of symmetry in reflected light phase curves. At visible wavelengths, reflection is always greatest at eastern and western quadrature and tend to overlap on the plot, while the albedo at secondary eclipse is slightly lower. This is due to the presence of a small open-ocean spot on the substellar hemisphere, which increases the disk averages albedo, compared to the completely ice- covered terminator regions. For cold climates that are faster rotating (Teff £ 3000 K) we observe different patterns. At visible wavelengths, the albedo at eastern quadrature is greatest, while the albedo at western quadrature and at secondary eclipse are less respectively. For these planets, increased zonal flow advects the sparse clouds eastward, and also results in enhanced snow fall over sea ice in the eastern hemisphere, leading to slightly higher surface albedos compared to the western hemisphere. For cold climates, near-infrared water vapor absorption features are discernable but weak. Some phase dependent variability is observed in the near-infrared window regions, generally mimicking the trends seen at visible wavelengths. For temperate climates, visible albedos are controlled by clouds rather than the surface albedo, and near-infrared water vapor absorption features show prominently in the albedo spectrum. Recall that temperate climates have thick cloud cover over the substellar hemisphere. For slow rotating temperate worlds, eastern quadrature and secondary eclipse phases have about the same visible wavelength albedo, while the albedo is slightly lower at western quadrature. Even for slow rotators, the subtle effect of a small Coriolis parameter exists, preferentially favoring cloudiness in the eastern hemisphere over the western hemisphere (Figure 4). This trend holds true for planets around all stars with Teff ³ 3000 K. For rapidly rotating temperate worlds, the strong advection of clouds eastward means that the eastern quadrature has a substantially higher visible albedo than is seen at secondary eclipse or western quadrature. For all temperate climates, near-infrared water vapor absorption bands show prominently. This is due to the combined effect of thick substellar clouds which raise albedos in water vapor window regions, and enhanced upper atmosphere water vapor amounts (e.g. Kopparapu et al. 2017; Fujii et al. 2017) which increase absorption within water bands. The resolution of the near-infrared water-vapor bands in reflectance spectra may be an important signpost for habitable worlds. Incipient runaway cases manifest similarly around each stellar type. In these hot moist climates, there are few clouds on the irradiated hemisphere, large water-vapor specific 5.3. Secondary Eclipse Spectra humidities, and no sea ice. Ocean albedos are low, thus reflected light is dominated by Rayleigh scattering. The spectral albedo longward of 0.7 µm is generally less than ~0.2. Shortward of 0.7 µm reveals a very steep Rayleigh slope, as water is only very weakly absorbing, but enhanced specific humidities contribute to increased scattering. To get a clearer comparison between the spectra of different climate states, here we collect thermal emission and albedo secondary eclipse spectra from each climate state, around each of our 6 stars (Figure 10). Generally, in the thermal emission (Figure 10, left column), the hotter the climate state, the more emission from the planet. This is true at all wavelengths, but is particularly evident in the water-vapor window region of 8 - 13 µm. Subsaturation of the atmospheres prevents the water vapor window region from closing, preventing the emergence of a hard radiation limit for moist atmospheres (Goldblatt et al. 2013). As noted earlier, for incipient runaway cases strong emission becomes evident in the ~4 µm water vapor window region. The ~4 µm feature may be a signpost for a planet undergoing a runaway greenhouse (see also Goldblatt et al. 2013; Leconte et al. 2013). However, based on secondary eclipse spectra alone, it may be difficult to discern between cold, cool, and temperate climate states. To discern these climate states, one also needs to additionally consider variations in phase dependent thermal emission spectra (Figure 8), and also planetary albedos. For phase dependent albedos (Figure 9, right column), as noted earlier, cool and temperate climates have the highest albedos in both the visible and near-infrared wavelengths due to the omnipresent thick substellar cloud deck for slow and Rhines circulation regimes. For slow rotators (Teff ³ 3300 K) spectral albedos for cool and temperate climates are nearly indistinguishable, with high albedos in the visible and prominent water vapor absorption features evident in the near-infrared. Cold climates, have a lower visible albedo than temperate worlds, and also have muted water vapor features in the near- infrared. Incipient runway cases have the lowest spectral albedos except at ~0.2 µm, the shortwave limit of our model, where atmospheric scattering is significant. For rapid rotating cases (Teff = 2600 K), the albedo is inversely related to the planet temperature, evident at all wavelengths. Rhines rotation cases Teff = 3000 K) represent a transition between slow and rapid rotating regimes. In an effort to constrain observability, in Figure 11 we show the combined light secondary eclipse spectra for each climate state, around each of our 6 stars. Spectra are given in planet/star contrast in units of parts-per-million (ppm). The combined light spectra shows the reflected stellar light plus the thermal emitted light from the planet in a single spectrum, here plotted relative to the emitted light from the star. Generally, the combined light spectral signals measured at secondary eclipse from any climate state only exceeds ~1 ppm at wavelengths longer than ~12 µm for planets around 4500 K stars, and ~6 µm for planets around 2600 K stars. Lower effective temperature stars are smaller, and thus dimmer, meaning that planet spectral signals are relatively larger. Part-per-billion (ppb) precision would allow us to spectrally resolve all thermal (³4 µm) water vapor and cloud features for planets around M-dwarf stars. Reflected light at wavelengths less than ~2 µm, are also generally amenable to detection at ppb precision in all cases, except for runaway greenhouse climates around 4500 K stars. The most difficult spectral region to detect lies between 2 - 3 µm, where part-per-trillion (ppt) precision would be required to resolve 6. Variation Spectra planet spectral features, except around the smallest M-dwarf stars where ppb may suffice. While ppb and ppt precisions are out of bounds currently for observations of combined light secondary eclipse spectra, our calculations here demonstrate that climate states may be discriminated based on the spectral signals resulting from expected water vapor and cloud distributions. Future missions and instruments should aim for ppb precision in order to characterize terrestrial planets around M-dwarf stars in thermal emission and reflected stellar light. Selsis et al. (2011) coined the term "variation spectrum" as the amplitude of variation in the radiance emitted or reflected by an exoplanet over the course of its orbit. In practice, the variation spectrum is calculated as the maximum observed radiance minus the minimum observed radiance, measured over the planet's orbit. In Figure 12, we plot the combined light variation spectra for incipient runaway (red), temperate (gold), cool (light blue), and cold (dark blue) climates, around each of our 6 stars. The left column shows the variation spectra in units of radiance (W m-2 µm-1) measured at the top of the atmosphere, while the right column shows the variation spectra in planet/star contrast in units of ppm. Similar to Figure 11, combined light is considered as the reflected stellar plus the outgoing thermal emitted radiation from the planet. Spectral differences in the variation spectra can yield information about climate states. In radiance units, the variation spectra at visible wavelengths is large. For shortwave variation spectra, the maximum reflection is observed at secondary eclipse, while the minimum in reflection occurs by transit where by definition no light is reflected towards to observer. Thus, the shortwave the variation spectrum indicates the reflectivity of the substellar hemisphere. The absolute maximum of the variation spectra in radiance units corresponds with the wavelength of peak emission from the star (Figure 1). As noted previously, temperate and cool climate states generally reflect more light than do cold and icy climates because water clouds have higher broadband albedos than does sea ice, when irradiated by M-dwarf stars (Shields et al. 2013). Around each star and for all climate states, a local minimum in the variation spectra is found corresponding with the ~2.7 µm water vapor absorption band. The magnitude of the variation spectra at 2.7 µm is inversely related to the planet temperature, with cold planets having the most variation, and incipient runaway states having the least. In the ~4 µm water vapor window, runaway greenhouse climates exhibit a local maximum that is not found for other climate states. Other climate states tend to exhibit either a modest minimum in the variation spectra or remain flat at ~4 µm, due to it residing at the shoulder of both thermal and stellar radiative fields. Generally, a local minimum is also observed near the ~6 µm water vapor feature, which is particularly sharp for incipient runaway cases. All climates around all stars exhibit maximum in the thermal variation spectra in the water vapor window region. Figure 12 right column shows the combined light variation spectra of planets relative to the stellar brightness. The brightness of the stars relative to planets means that the planet/star contrast at short wavelengths is significantly lower than that observed at thermal wavelengths. Still, many of the same trends hold true that were noted for variation spectra in radiance units. Local minima are seen in water vapor absorption bands, most notably at ~2.7 µm and ~6 µm. At ~4 µm, runaway greenhouse states exhibit a maximum, while all other climate states display either a modest minimum or flat spectra. All climates exhibit their largest variation at thermal wavelengths, where the stellar emission is weak. Cold climates exhibit the largest variation spectra beyond ~10 µm, while that from runaway states is muted. The precision required to observe variation spectra is greater than that required to measure secondary eclipse spectra (Figure 11), because phase dependent variations in radiance are, naturally, less than their magnitude. To capture variation spectra at wavelengths ³6 µm, ppb precision is generally sufficient for planets around all stars studied. However, some of the most interesting and tell-tale spectral signals reside in the 2 - 6 µm range, which becomes increasingly challenging to observe from planets around large stars. In this section we discuss numerous caveats to the work presented here. In this study, we have only considered Earth-sized aquaplanets, with 1 bar of N2 along with H2O and clouds. The presence of different absorbing species would affect climate, phase curves, and phase dependent spectra. Naturally, the addition of other greenhouse gases would result in warmer climate states for a given stellar flux. Also, gases such as CO2 and CH4 are generally well-mixed species, meaning that they are evenly distributed about the planet, and thus may uniformly reduce the magnitude of broadband thermal emission phase curves. While well- mixed greenhouse gases would surely be evident in spectra, it is unlikely that they would contribute significantly to the variation spectra (i.e. phase dependent variability). For instance, Boutle et al. (2017) showed, based on tidally-locked simulations of Proxima b, that phase dependent emission spectra are insensitive to phase variations in the 15 µm CO2 7. Discussions absorption band. A similar conclusion was also reached by Turbet et al. (2016) based on Proxima b simulations with a different model. Clouds dominate phase dependent variability, not gaseous absorbing species, and thus phase dependent variability is most evident in window regions of the spectra. Still, strong absorption by CO2 at 15 µm or by O3 at 9.6 µm, could encroach upon the water-vapor window region, and partially obscure the variation spectra that have been presented here. Additionally, planets with drastically different atmospheric compositions and total surface pressures would be expected to have significant differences in temperature, atmospheric circulation, heat transport, lapse rates, convection, and cloud formation. Each of these aspects would affect thermal and reflected phase curves morphologies and amplitudes. Analysis of Kepler planets indicates that terrestrial planets could feasibly be as large as 1.5 Earth radii, with larger planets being Neptune-like instead of terrestrial (Rogers, 2015; Wolfgang et al. 2016). Changes in planetary radius could have modest effects on atmospheric dynamics and circulation regimes, quantified by changes to the Rossby radius of deformation vis-à-vis the planet radius (Carone et al. 2018). Changes in planet size and internal composition would also be expected to change the surface gravity. Changes in surface gravity would affect many aspects of the climate system, including the lapse rate, scale heights, and particle sedimentation rates, all of which could strongly affect the nature of clouds and climate on extrasolar planets. Naturally, changes to clouds and climate would feedback upon the phase dependent thermal and reflected light signals. Larger sized planets would be more favorable to observation, yielding larger planet/star contrasts than are shown here, and vice-versa. We have also assumed that these planets are all tidally locked into synchronous rotation with their host star. Resonant rotation states and eccentric orbits would also influence phase dependent observations (e.g. Wang et al. 2014; Bolmont et al. 2016), which we have not considered here. Here, for simplicity, computational convenience, and continuity with past works (Kopparapu et al. 2017; Haqq-Misra et al. 2018), we have assumed a slab ocean with zero ocean heat transport. Ocean heat transport may have a significant effect on climate and also the location of clouds (Way et al. 2018). Dynamic ocean heat transport on completely ocean covered worlds (i.e. no continents), leads to warmer global mean temperatures and a reduction in day-night temperature differences on synchronously rotating planets (Hu & Yang 2014; Del Genio et al. 2019). Yang et al. (2019) argue that ocean heat transport is critical for the treatment of cold, partially ice-covered planets around M-dwarf stars, however it does not have a meaningful effect on the climate and thermal emission phase curves for warm terrestrial planets (Ts ~ 300 K) residing close to the inner edge of the habitable zone. The presence of continents also presents a significant uncertainty in the net effect of ocean heat transport on climate, as continents can reroute or eliminate day-to-night ocean heat transport entirely, resulting in climates states that are similar to those found in simulations with zero ocean heat transport (Del Genio et al. 2019; Yang et al. 2019). In this work, we have argued that clouds are likely more important in modulating phase curves for habitable planets, compared to surface energy transports. Still, ocean heat transport in continental configurations and topographic patterns combined with variations prominences could have significant complex feedbacks on the surface temperature, sea-ice and snow cover, and cloud distributions respectively, all of which affect thermal and reflected light phase curves. In future works, our group plans to couple our GCM with a dynamic ocean model to study the interactions between oceans, continents, topography, and observable features. However, the long equilibration times for a coupled dynamic atmosphere-ocean models still presents a significant computational hinderance for conducting 3D simulations across wider parameter spaces, as we have done here. The suite of simulations studied here, represents only a small fraction of the possible worlds that could reside in the habitable zones of M-dwarf stars. While here we have demonstrated how water vapor and clouds may affect thermal and reflected light phase curves for Earth-like aquaplanets, more work is needed to systematically map how other parameters would affect phase dependent observations. Future planned work will expand our modeling to include ocean heat transport, different atmospheric compositions, planet sizes, and surface configurations. Terrestrial planets orbiting low-mass stars will provide our best opportunity for detecting and verifying a habitable world in the coming decades. Here, we suggest that broadband and spectrally resolved thermal emitted and reflected light phase curves can be used to characterize the climate states of tidally locked terrestrial extrasolar planets orbiting M-dwarf stars. The spatial distributions of water vapor, clouds, and surface ice significantly impact phase dependent spectral signals from terrestrial planets. In turn, the mean climate state of a planet controls the partitioning of water into vapor, cloud, and ice phases. Thus, by carefully studying the phase dependent signals of water in each of its thermodynamic phases, we may connect the dots between observations and climate states for terrestrial extrasolar planets around M-dwarf stars. Here, we have calculated theoretical broadband and spectrally resolved thermal emission and reflected light phase curves, as well as 8. Conclusions Acknowledgements variation spectra, based on 3D climate modeling simulations of extrasolar planets in a variety of climate states, ranging from cold, ice dominated worlds with Ts ~ 205 K, to hot, moist, optically thick worlds with Ts ~ 400 K and with a runaway greenhouse initiated. Cold, temperate and incipient runaway greenhouse climate states exhibit unique phase dependent spectra in both thermal and reflected light. For potentially habitable worlds, we recommend probing the spectral window regions, where variability in clouds and surface properties will show most strongly. In particular, we recommend probing water vapor window regions in the near-infrared, 4 µm, and in the 8 - 13 µm region. Furthermore, we recommend that future missions and instruments aim for part-per-billion precisions, in order to sufficiently characterize water vapor and cloud features in the atmospheres of terrestrial planets. R.K., E.T.W., and J.H.-M. acknowledge funding from NASA Habitable Worlds grant NNX16AB61G. R.K. and E.T.W. acknowledge funding from the Sellers Exoplanet Environments Collaboration. R.K., E.T.W., and J.H.-M. acknowledge funding from the Virtual Planetary Laboratory under Cooperative Agreement number NNA13AA93A. E.T.W. and R.K. acknowledge NASA Astrobiology through participation in the Nexus for Exoplanet System Science. The 3D simulation outputs used in this work have been publicly archived, and are available at the following address: https://archive.org/details/SimulatedPhaseDependentSpectraOfTerrestrialAquaplanetsIn MdwarfSystems, or by contacting the lead author, [email protected]. Institute CAN7 award NNH13DA017C References Adams, A.D., Boos, W.R. & Wolf, E.T. 2019, accepted to ApJ, arXiv:1903.06216 Allard, F., Allard, N.F., Homeier, D., et al. (2007), A&A, 474, L21 Baraffe, I., Chabrier, G., Allard, F, & Hauschildt, P.H. (2002) A&A, 382, 563 Bolmont, E., Libert, A.-S., Leconte, J. & Selsis, F. (2016) A&A 591, A106 Borysow, A. & Frommhold, L. 1986, ApJ, 331, 1043 Boyajian, T.S., Von Braun, K., van Belle, G., et al. (2013) ApJ, 771, 40 Carone, L., Keppens, R., Decin, L., & Henning, Th. 2018, MNRAS, 473, 4672 Cowan, N.B., Voigt, A. & Abbot, D.S. (2012) ApJ 757:80 (13 pp) Del Genio, A.D., Way, M.J., Admundsen, D.S. et al. 2018, AsBio, 19, 2 Del Genio, A. Kiang, N., Way, M. et al. 2019, submitted to ApJ, arXiv:1812.06606 Fujii, Y., Del Genio, A., & Amundsen, D.S. 201, ApJ, 848, 100 Goldblatt, C. & Watson, A.J. 2012, RSPTA, 370, 4197 Goldblatt, C., Robinson, T.D., Zahnle, K.J., & Crisp, D. 2013, NatGeo, 6, 661 Goldblatt, C. 2015, 15, 362 Grimm S., & Heng K. 2015, ApJ, 808, 182 Haqq-Misra, J., Wolf, E.T., Joshi, M. et al. 2018, ApJ, 852:67 (16pp) Harris, J.E. & Belotti C. 2009, JClim, 23, 1277 Hu, Y. & Yang, J., 2014, PNAS, 111, 629 Ishiwatari, M., Takehiro, S.-I., Nakajima, K., & Hayashi, Y.-Y., 2002, JAS, 59(22), 3223 Joshi, M.M., Haberle, R.M., & Reynolds, R.T. 1997, Icarus, 129, 40 Joshi, M.M. 2003, Asbio, 3, 2, 415 Joshi, M.M. & Haberle, R.M. 2011, Asbio, 12, 1, 3 Kane, S.R., Ceja, A.Y., Way, M.J., & Quintana, E.V. 2018, ApJ, 869,1 Kasting, J.F. 1998, Icarus, 74, 472 Kasting, J.F., Whitmire, D.P. & Reynolds, R.T. 1993, Icarus, 101, 108 Koll D.D.B. & Abbot D.S. 2015, ApJ, 802, 21 (15pp) Koll D.D.B. & Abbot, D.S. 2016, ApJ, 825, 99 (22 pp) Knutson, H.A., Charbonneau, D., Allen, L.E., et al.(2007, Nature, 447, 183 Komacek, T.D. and Abbot, D.S. 2019, arXiv:1901.00567 Kopparapu, R.K., Ramirez, R., Kasting, J.F., et al. 2013, ApJ, 765, 131 (16pp) Kopparapu, R.K., Wolf, E.T., Haqq-Misra, J. et al. 2016, ApJ, 819, 1 Kopparapu, R. k., Wolf, E.T., Arney, G., et al. 2017, ApJ, 845,5 (16pp) Leconte, J., Forget, F., Charnay, B. et al. 2013, A&A, 554, A69 Lin, S-J., & Rood, R.B. 1995, MWRv, 124, 2046 Meadows, V.S., Arney, G.N., Schwieterman, E.W. et al. 2018, AsBio, 18(2), 133 Merlis, T.M. & Schneider, T. 2010, JAMES, 2, 13 Neale, R.B., Richter, J.H., Conlet, A.J. et al. 2010, TN-486/TN-486+STR Pincus, R., Barker, H.W., & Morcrette, J. 2003, JGRD, 108, 4376 Parmentier, V., Fortney, J.J., Showman, A.P. et al. 2016, ApJ, 828,22 Rhines, P. B. 1975, JFM, 69, 417 Robinson, T.D., Meadows, V.S., & Crisp, D. 2010, ApJ, 721, 1 Rogers, L. A. 2015, ApJ, 801, 41 Rothman, L.S., Gordon, L.E., Babikov, Y. et al. 2013, JQSRT, 130, 4 Schwieterman, E.W., Robinson, T.D., Meadows, V.S. et al. 2015, ApJ, 810,57 Selsis, F., Wordsworth, R.D., & Forget, F. 2011, A&A, 532A, 1S Shields, A.L., Meadows, V.S., Bitz, C.M., et al. 2013, Asbio, 13(8), 715 Shields, A.L., Barnes, R., Agol, E., et al. 2016, AsBio, 16, 6, 443 Shields, A.L. & Carns, R.C. 2018, ApJ, 867, 11 Showman, A.P., Cho, J.Y.-K., & Menou, K. 2010, in Exoplanets, ed. S. Seager (Tucson, AZ: Univ. Arizona Press), 471 Showman, A.P. & Polvani, L.W. 2011, ApJ, 738:71 (24pp) Stephens, G.L., Li, J., Wild, M., et al. 2012, NatGeo, 5, 691 Turbet, M., Leconte, J., Selsis, F., et al. 2016, A&A, 596, A112 Turbet, M., Bolmont, E., Leconte, J. et al. 2018, A&A, 612, A86 Wang, Y., Tian, F., & Hu, Y. 2014, ApJ, 791, L12 (6 pp) Way, M.J., Del Genio, A.D., Aleinov, I., et al. 2018, ApJS, 239, 24 Wolf E.T., Shields, A.L., Kopparapu, R.K., et al. 2017, ApJ, 827:107 (11 pp). Wolfgang, A., Rogers, L.A. & Ford, E.B. 2016, ApJ, 825(1) Wordsworth, R. & Pierrehumbert, R. 2013, Science, 339, 64 Yang, J., Cowan, N., & Abbot, D.S. 2013, ApJL, 771, L45 Yang, J., Boue, G., Fabrycky, D.C. & Abbot, D.S. 2014, ApJL, 787, L2 Yang, J., Abbot, D.S., Koll, D.D.B., et al. 2019, ApJ, 871, 29 Figure Captions Figure 1: Stellar spectra used in this study at the binned resolution of our model (solid lines.) The solar spectra is shown for comparative purposes (dotted line). Figure 2: Maps of surface temperature from cold, cool, temperate, and incipient runaway greenhouse worlds (columns), around each of our 6 stars (rows). The panels are centered on the substellar point. The white line indicates the extent of sea-ice in each simulation. Note the range in orbital periods and incident stellar fluxes are indicated in the left-hand column, with the longest (lowest) orbital period (incident stellar flux) corresponding with the cold simulations, and the shortest (highest) orbital period (incident stellar flux) corresponding with the incipient runaway greenhouse. Figure 3. Maps of the water vapor column (kg m-2) for various climate states around each star. Same format as Figure 2. Figure 4. Maps of the liquid cloud water column (g m-2) for various climate states around each star. Same format as Figure 2. Note in the colorbar that clouds are shown in white. Figure 5. Maps of the ice cloud water column (g m-2) for various climate states around each star. Same format as Figure 2. Note in the colorbar that clouds are shown in white. Figure 6: Broadband thermal emission phase curves calculated for all simulations, around each of our 6 stars. The fluxes given are the top-of-atmosphere values, disk-averaged as a function of phase. Phase curves are color coded according to climate state. Solid gold lines indicate temperate climates, solid light blue lines indicate cool climates, and solid dark blue lines indicate cold climates, with surface temperature ranges given in the legend. Dashed lines indicate incipient runaway greenhouse cases, and are representative of hot, moist, optically thick atmospheres. Figure 7: Broadband albedo phase curves calculated for all simulations, around each of our 6 stars. Same format and color coding as Figure 6. Figure 8: Phase dependent thermal emission spectra for cold, temperate, and incipient runaway cases, around each of our 6 stars. Plotted are the disk-averaged spectra at transit, western quadrature, secondary eclipse, and eastern quadrature. Note, for all cold climates and for temperate climates around star with Teff ≥ 3700 K, phase curves appear symmetrical and thus emission spectra at eastern and western quadrature are identical and tend to overlap and thus are difficult to distinguish without close inspection. Figure 9: Phase dependent albedo spectra for cold, temperate, and incipient runaway cases, around each of our 6 stars. Identical format and color coding as Figure 8. Plotted are the disk-averaged spectra at western quadrature, secondary eclipse, and eastern quadrature. Note, some overlaps between phase dependent reflectance spectra occur, and thus can be difficult to distinguish without close inspection. Figure 10: Thermal emission (left column) and albedo (right column) secondary eclipse spectra for incipient runaway, temperate, cool, and cold climates cases, around each of our 6 stars. Figure 11: Combined light spectra at secondary eclipse for cold, cool, temperate, and incipient runaway, climate states around each of our 6 stars. Combined light is the reflected stellar light plus the thermal emitted light. Spectra are given in units planet/star contrast in parts-per-million. Figure 12: Combined light variation spectra for cold, cool, temperate, and incipient runaway, climate states, around each of our 6 stars. The right column is in units of top-of- atmosphere radiances, and the left column shows planet/star contrasts in part-per-million. 2600 K 3000 K 3300 K 3700 K 4000 K 4500 K Sun 1 2 Wavelength (µm) 3 4 5 Figure 1: 2000 1500 1000 500 ) m µ 2 - m W ( e c n a d a R i 0 0 273 250 2 3 0 2 5 0 2 3 0 Cold 350 175 0 0 2 2 5 0 175 2 7 3 2 2 0 5 0 0 2 5 0 2 0 0 175 2 5 0 2 0 0 2 7 3 2 5 1 0 7 200 5 0 5 2 2 0 0 1 7 5 400 175 1 7 5 175 2 7 3 175 Figure 2 150 3 6 5 3 2600 K 4.0 - 6.9 days 1375 - 680 W m-2 Runaway 7 369 6 Surface temperature (K) 200 250 Temperate 273 2 8 0 0 0 3 315 300 280 273 300 Cool 0 3 2 0 5 2 0 9 2 378 367 3 0 0 2 7 2 8 0 3 2 9 0 285 3 0 5 3 0 0 2 7 3 0 3 2 2 5 0 0 9 2 273 290 285 0 3 2 2 7 3 250 2 7 3 250 2 7 2 3 5 0 290 2 7 3 250 1373 3 6 7 3 7 3 6 9 356 4 6 3 3 6 8 3 6 0 2 6 3 8 6 3 3 6 5 3 8 5 5 1 3 3 1 0 2 6 3 362 0 1 3 3 0 3 5 0 0 3 1 0 3 8 9 3 8 7 5 8 3 3 3 0 0 5 0 3000 K 7.1 - 14.4 days 1600 - 680 W m-2 8 7 3 0 8 3 3300 K 18.3 - 38.1 days 1750 - 680 W m-2 3700 K 35.7 - 80.2 days 1950 - 680 W m-2 4000 K 53.7 - 125.0 days 2050 - 680 W m-2 362 4500 K 78.3 - 207.9 days 2350 - 680 W m-2 Water Vapor Column (kg m-2) Temperate 100 2 0 0 150 100 50 25 10 10 1 0 2 5 5 0 100 150 6 0 120 0 0 3 60 120 3 0 0 0 0 1 8 0 0 8 1 1 1 4 6 0 0 3 7 5 3 3 5 2 0 5 1 7 5 2 5 0 2 2 2 5 0 0 175 0 0 1 1 5 0 1 2 5 5 1 0 2 5 5 10 20 Cool 75 5 0 5 0 4 0 30 1000 25 10 5 2 0 1 0 5 0 4 3 0 20 0 4 3 0 20 0 4 3 0 20 1 0 4 0 10 3 0 2 0 Cold 3 5 9 7 5 3 7 5 3 7 5 3 7 5 3 5 3 10000 Runaway 5000 4 8 0 0 4600 6000 5000 8 2 0 0 5 4 0 0 4 4 0 0 0 0 2 5 Figure 3 1 6000 8 0 0 0 8 2 0 0 2600 K 6.9 - 4.0 days 1375 - 680 W m-2 3000 K 14.4 - 7.1 days 1600 - 680 W m-2 3300 K 38.1 - 18.3 days 1750 - 680 W m-2 5200 0 0 4 5 3700 K 80.2 - 35.7 days 1950 - 680 W m-2 0 0 4 4 4000 K 125.0 - 53.7 days 2050 - 680 W m-2 4500 K 207.9 - 78.3 days 2350 - 680 W m-2 4200 0 0 4 5 0 0 0 5 0 5 7 4 4500 Liquid cloud water path (g m-2) Cool 100 50 Cold 1000 10 Temperate 100 400 0 1 0 0 8 50 800 1 0 0 0 0 2 400 200 400 100 200 50 200 1 0 0 5 0 1 0 0 5 0 0 5 1 0 0 0 0 4 0 0 4 2 0 0 2 5 0 0 0 100 5 0 0 0 2 200 4 0 2 0 0 0 100 0 0 4 5 0 0 400 1 2 0 0 0 5 0 5 0 1 0 400 5 2 0 1 0 0 0 0 5 50 5 2 5 0 0 01 2 5 5 50 2 5 5 5 0 2 5 5 2 5 5 2 5 5 Figure 4 2600 K 6.9 - 4.0 days 1375 - 680 W m-2 1 Runaway 100 200 3 0 0 500 3 0 0 500 100 0 0 4 400 200 0 0 2 2 0 0 50 100 100 0 0 3 200 100 200 200 0 0 4 0 0 0 4 0 8 0 0 2 0 5 100 1 0 0 200 1 0 0 0 5 0 0 1 0 0 03 0 2 200 8 0 0 4 0 0 2 50 1 0 0 200 300 1 0 0 0 5 100 0 0 0 0 0 1 5 0 0 0 0 8 100 2 4 0 0 8 0 0 4 0 0 0 0 2 0 0 2 3000 K 14.4 - 7.1 days 1600 - 680 W m-2 300 2 0 0 3300 K 38.1 - 18.3 days 1750 - 680 W m-2 3700 K 80.2 - 35.7 days 1950 - 680 W m-2 4000 K 125.0 - 53.7 days 2050 - 680 W m-2 4500 K 207.9 - 78.3 days 2350 - 680 W m-2 Cold 50 2 5 1 0 1 5 0 1 02 5 100 1 1 5 2 5 0 1 0 1 2 0 55 10 1 5 2 5 0 1 0 1 5 2 5 0 10 1 Ice cloud water path (g m-2) 1 Temperate Cool 10 50 25 10 10 25 5 0 1 10 2 5 10 5 1 5 1 25 10 1 0 1 1 1 0 1 1 5 5 1 1 0 1 0 0 1 1 0 5 1 0 1 0 10 0 1 5 5 2 5 10 5 1 0 1 10 2 5 5 2 10 5 1 5 1 1 5 0 1 1 0 2 5 5 5 Figure 5 0.1 Runaway 2600 K 6.9 - 4.0 days 1375 - 680 W m-2 5 3000 K 14.4 - 7.1 days 1600 - 680 W m-2 3 5 3 5 5 13 5 1 3 3 3 3 5 10 25 5 0 25 1 2 5 0 1 1 5 1 5 1 0 3 1 5 1 0 5 5 5 3 3 1 3 13 3 1 5 5 3 5 5 3 3 3300 K 38.1 - 18.3 days 1750 - 680 W m-2 3700 K 80.2 - 35.7 days 1950 - 680 W m-2 4000 K 125.0 - 53.7 days 2050 - 680 W m-2 4500 K 207.9 - 78.3 days 2350 - 680 W m-2 Figure 6: transit west secondary east transit transit west secondary east transit runaway: TS > 350 K temperate: 273 < TS < 310 K cool: 250 < TS < 273 K cold: 200 < TS < 250 K 90 Orbital Phase (degrees) 45 0 90 Orbital Phase (degrees) 45 0 135 180 3300 K 2600 K 500 400 300 200 100 0 500 400 300 200 100 0 500 400 300 200 100 0 -180 -135 -90 -45 4000 K ) 2 - m W ( x u F l ) 2 - m W ( x u F l ) 2 - m W ( x u F l 135 180 3000 K 3700 K 500 400 300 200 100 0 500 400 300 200 100 0 500 400 300 200 100 0 -180 -135 -90 -45 4500 K ) 2 - m W ( x u F l ) 2 - m W ( x u F l ) 2 - m W ( x u F l transit transit west secondary east transit 3000 K 3700 K 4500 K 0.75 0.60 0.45 0.30 0.15 0.00 0.75 0.60 0.45 0.30 0.15 0.00 0.75 0.60 0.45 0.30 0.15 0.00 -180 -135 -90 -45 90 Orbital Phase (degrees) 45 0 135 180 -180 -135 -90 -45 90 Orbital Phase (degrees) 45 0 135 180 Figure 7: transit west 2600 K secondary east runaway: TS > 350 K temperate: 273 < TS < 310 K cool: 250 < TS < 273 K cold: 200 < TS < 250 K 3300 K 4000 K o d e b A l o d e b A l o d e b A l 0.75 0.60 0.45 0.30 0.15 0.00 0.75 0.60 0.45 0.30 0.15 0.00 0.75 0.60 0.45 0.30 0.15 0.00 o d e b A l o d e b A l o d e b A l 0 5 10 15 20 25 30 35 40 Wavelength (µm) 5 10 15 20 25 Wavelength (µm) 30 35 40 Phase Dependent Thermal Emission Spectra Temperate Runaway Cold at transit at western quadrature at secondary eclipse at eastern quadrature 5 10 15 20 25 30 35 40 Wavelength (µm) 0 Figure 8: 40 30 20 10 0 40 30 20 10 0 40 30 20 10 0 40 30 20 10 0 40 30 20 10 0 40 30 20 10 0 0 Teff ) 1 - e c n a d a R i m µ 2 - m W ( ) 1 - e c n a d a R i m µ 2 - m W ( ) 1 - e c n a d a R i m µ 2 - m W ( ) 1 - e c n a d a R i m µ 2 - m W ( ) 1 - e c n a d a R i m µ 2 - m W ( ) 1 - e c n a d a R i m µ 2 - m W ( K 0 0 6 2 K 0 0 0 3 K 0 0 3 3 K 0 0 7 3 K 0 0 0 4 K 0 0 5 4 Cold 0 0 0.5 1.5 1.0 2.0 Wavelength (µm) 2.5 3.0 0.0 0.5 1.5 1.0 2.0 Wavelength (µm) 2.5 3.0 0.5 1.5 1.0 2.0 Wavelength (µm) 2.5 3.0 Phase Dependent Albedo Spectra Temperate Runaway at western quadrature at secondary eclipse at eastern quadrature Figure 9: 1.0 0.8 0.6 0.4 0.2 0.0 1.0 0.8 0.6 0.4 0.2 0.0 1.0 0.8 0.6 0.4 0.2 0.0 1.0 0.8 0.6 0.4 0.2 0.0 1.0 0.8 0.6 0.4 0.2 0.0 1.0 0.8 0.6 0.4 0.2 0.0 Teff K 0 0 6 2 o d e b A l K 0 0 0 3 o d e b A l K 0 0 3 3 o d e b A l K 0 0 7 3 o d e b A l K 0 0 0 4 o d e b A l K 0 0 5 4 o d e b A l 5 10 15 25 Wavelength (µm) 20 30 35 40 0.0 0.5 1.0 2.0 Wavelength (µm) 1.5 2.5 3.0 Secondary Eclipse Spectra runaway temperate cool cold o d e b A l o d e b A l o d e b A l o d e b A l o d e b A l o d e b A l 1.0 0.8 0.6 0.4 0.2 0.0 1.0 0.8 0.6 0.4 0.2 0.0 1.0 0.8 0.6 0.4 0.2 0.0 1.0 0.8 0.6 0.4 0.2 0.0 1.0 0.8 0.6 0.4 0.2 0.0 1.0 0.8 0.6 0.4 0.2 0.0 Figure 10: 40 30 20 10 0 40 30 20 10 0 40 30 20 10 0 40 30 20 10 0 40 30 20 10 0 40 30 20 10 0 0 Teff ) 1 - e c n a d a R i m µ 2 - m W ( ) 1 - e c n a d a R i m µ 2 - m W ( ) 1 - e c n a d a R i m µ 2 - m W ( ) 1 - e c n a d a R i m µ 2 - m W ( ) 1 - e c n a d a R i m µ 2 - m W ( ) 1 - e c n a d a R i m µ 2 - m W ( K 0 0 6 2 K 0 0 0 3 K 0 0 3 3 K 0 0 7 3 K 0 0 0 4 K 0 0 5 4 Figure 11: Teff Combined Light Secondary Eclipse Spectra runaway temperate cool cold 1.0 Wavelength (µm) 10.0 100.0 1 ppm 1 ppb 1 ppt 1 ppm 1 ppb 1 ppt 1 ppm 1 ppb 1 ppt 1 ppm 1 ppb 1 ppt 1 ppm 1 ppb 1 ppt 1 ppm 1 ppb 1 ppt 102 100 10-2 10-4 10-6 102 100 10-2 10-4 10-6 102 100 10-2 10-4 10-6 102 100 10-2 10-4 10-6 102 100 10-2 10-4 10-6 102 100 10-2 10-4 10-6 0.1 K 0 0 6 2 ) m p p ( s F / p F K 0 0 0 3 ) m p p ( s F / p F ) K 0 0 3 3 m p p ( s F / p F ) K 0 0 7 3 m p p ( s F / p F ) K 0 0 0 4 m p p ( s F / p F ) K 0 0 5 4 m p p ( s F / p F Figure 12: Teff 1 - K 0 0 6 2 m µ 2 - m W 1 - K 0 0 0 3 m µ 2 - m W 1 - K 0 0 3 3 m µ 2 - m W 1 - K 0 0 7 3 m µ 2 - m W 1 - K 0 0 0 4 m µ 2 - m W 1 - K 0 0 5 4 m µ 2 - m W 104 102 100 10-2 10-4 104 102 100 10-2 10-4 104 102 100 10-2 10-4 104 102 100 10-2 10-4 104 102 100 10-2 10-4 104 102 100 10-2 10-4 ) m p p ( S F / P F ) m p p ( S F / P F ) m p p ( S F / P F ) m p p ( S F / P F ) m p p ( S F / P F ) m p p ( S F / P F 0.1 1.0 Wavelength (µm) 10.0 100.0 102 100 10-2 10-4 10-6 102 100 10-2 10-4 10-6 102 100 10-2 10-4 10-6 102 100 10-2 10-4 10-6 102 100 10-2 10-4 10-6 102 100 10-2 10-4 10-6 1 ppm 1 ppb 1 ppt 1 ppm 1 ppb 1 ppt 1 ppm 1 ppb 1 ppt 1 ppm 1 ppb 1 ppt 1 ppm 1 ppb 1 ppt 1 ppm 1 ppb 1 ppt 0.1 Top-of-Atmosphere Radiance Planet/Star Contrast Combined Light Variation Spectra runaway temperate cool cold 1.0 Wavelength (µm) 10.0 100.0
1301.0770
1
1301
2013-01-04T16:52:53
Crantor, a short-lived horseshoe companion to Uranus
[ "astro-ph.EP" ]
Stable co-orbital motion with Uranus is vulnerable to planetary migration but temporary co-orbitals may exist today. So far only two candidates have been suggested, both moving on horseshoe orbits: 83982 Crantor (2002 GO9) and 2000 SN331. (83982) Crantor is currently classified in the group of the Centaurs by the MPC although the value of its orbital period is close to that of Uranus. Here we revisit the topic of the possible 1:1 commensurability of (83982) Crantor with Uranus and also explore its dynamical past and look into its medium-term stability and future orbital evolution. (83982) Crantor currently moves inside Uranus' co-orbital region on a complex horseshoe orbit. The motion of this object is primarily driven by the influence of the Sun and Uranus, although Saturn plays a significant role in destabilizing its orbit. The precession of the nodes of (83982) Crantor, which is accelerated by Saturn, controls its evolution and short-term stability. Although this object follows a temporary horseshoe orbit, more stable trajectories are possible and we present 2010 EU65 as a long-term horseshoe librator candidate in urgent need of follow-up observations. Available data indicate that the candidate 2000 SN331 is not a Uranus' co-orbital.
astro-ph.EP
astro-ph
Astronomy & Astrophysics manuscript no. aa20646-12 November 2, 2018 c(cid:13) ESO 2018 3 1 0 2 n a J 4 . ] P E h p - o r t s a [ 1 v 0 7 7 0 . 1 0 3 1 : v i X r a Crantor, a short-lived horseshoe companion to Uranus (cid:63) C. de la Fuente Marcos and R. de la Fuente Marcos Universidad Complutense de Madrid, Ciudad Universitaria, E-28040 Madrid, Spain Received 29 October 2012 / Accepted 2 January 2013 ABSTRACT Context. Stable co-orbital motion with Uranus is vulnerable to planetary migration but temporary co-orbitals may exist today. So far only two candidates have been suggested, both moving on horseshoe orbits: 83982 Crantor (2002 GO9) and 2000 SN331. Aims. (83982) Crantor is currently classified in the group of the Centaurs by the MPC although the value of its orbital period is close to that of Uranus. Here we revisit the topic of the possible 1:1 commensurability of (83982) Crantor with Uranus and also explore its dynamical past and look into its medium-term stability and future orbital evolution. Methods. Our analysis is based on the results of N-body calculations that use the most updated ephemerides and include perturbations by the eight major planets, the Moon, the barycentre of the Pluto-Charon system, and the three largest asteroids. Results. (83982) Crantor currently moves inside Uranus' co-orbital region on a complex horseshoe orbit. The motion of this object is primarily driven by the influence of the Sun and Uranus, although Saturn plays a significant role in destabilizing its orbit. The precession of the nodes of (83982) Crantor, which is accelerated by Saturn, controls its evolution and short-term stability. Although this object follows a temporary horseshoe orbit, more stable trajectories are possible and we present 2010 EU65 as a long-term horseshoe librator candidate in urgent need of follow-up observations. Available data indicate that the candidate 2000 SN331 is not a Uranus' co-orbital. Conclusions. Our calculations confirm that (83982) Crantor is currently trapped in the 1:1 commensurability with Uranus but it is unlikely to be a primordial 1:1 librator. Although this object follows a chaotic, short-lived horseshoe orbit, longer term horseshoe stability appears to be possible. We also confirm that high order resonances with Saturn play a major role in destabilizing the orbits of Uranus co-orbitals. Key words. minor planets, asteroids: general -- minor planets, asteroids: individual: 83982 Crantor (2002 GO9) -- minor planets, asteroids: individual: 2010 EU65 -- minor planets, asteroids: individual: 2000 SN331 -- celestial mechanics -- methods: numerical 1. Introduction An object librating in a trajectory which encompasses the Lagrangian points L4, L5 and L3 of a host planet evolves on a so-called regular horseshoe orbit. Viewed in a frame of refer- ence that co-rotates with the host planet, the shape of such an orbit projected onto the ecliptic plane resembles that of an ac- tual horseshoe although its three-dimensional layout looks more like a corkscrew around the orbit of the host planet while both revolve around the Sun. The size of the horseshoe orbit depends on the mass of the host planet, being wider and having a better chance of survival when this mass decreases (Dermott & Murray 1981a). In general, horseshoe orbits are not considered to be long-term stable (Dermott & Murray 1981a; Murray & Dermott 1999). Horseshoe orbits were originally predicted by Brown (1911) and Darwin (1912) and further studied later by, for exam- ple, Thuring (1959), Rabe (1961), Giacaglia (1970), Weissman & Wetherill (1974) and Garfinkel (1977) but were largely con- sidered theoretical curiosities until the Saturnian moons Janus and Epimetheus were identified as horseshoe librators (Smith et al. 1980; Synnott et al. 1981; Dermott & Murray 1981b). The existence of minor planets moving on horseshoe orbits around the major planets was first postulated by Milani et al. (1989) and Michel et al. (1996) on theoretical grounds. Send offprint requests [email protected] to: C. de la Fuente Marcos, e-mail: (cid:63) Figures 2 and 6 (animations) are available in electronic form at http://www.aanda.org In the solar system, there are several real examples of mi- nor bodies moving on such orbits. The first minor body to be confirmed to follow a horseshoe orbit was 3753 Cruithne (1986 TO) (Wiegert et al. 1997, 1998), in this case with the Earth. Karlsson (2004) found multiple objects moving in temporary horseshoe orbits with Jupiter. Connors et al. (2004) identified an object, 2003 YN107, following a compound horseshoe-quasi- satellite orbit with the Earth. Additional objects moving in com- parable trajectories are 2002 AA29 and 2001 GO2 (Brasser et al. 2004). 2001 CK32 follows an orbit similar to that of (3753) Cruithne but hosted by Venus, not the Earth (Brasser et al. 2004). (36017) 1999 ND43 is a horseshoe librator with Mars (Connors et al. 2005). Yet another horseshoe companion of the Earth was found in 2010 SO16 (Christou & Asher 2011). Additional horse- shoe librators with Jupiter were recently identified by Wajer & Kr´olikowska (2012). Finally, (310071) 2010 KR59 is following a temporary and rather complex horseshoe orbit with Neptune (de la Fuente Marcos & de la Fuente Marcos 2012a). Numerical simulations predict that Uranus may have re- tained a certain amount of its primordial co-orbital minor planet population (Holman & Wisdom 1993; Wiegert et al. 2000; Nesvorn´y & Dones 2002; Marzari et al. 2003). However, Uranus appears not to be able to efficiently capture objects into the 1:1 commensurability today even for short periods of time (Horner & Evans 2006). The stability of hypothetical Uranus co-orbitals, specifically those moving in tadpole orbits, has been studied by Dvorak et al. (2010) and they have found that the orbital inclina- tion is the key parameter regarding stability, only the inclination intervals (0, 7)◦, (9, 13)◦, (31, 36)◦ and (38, 50)◦ appear to be sta- 1 C. de la Fuente Marcos and R. de la Fuente Marcos: Crantor, a horseshoe companion to Uranus ble. This scarcity of Uranus co-orbitals seems to be confirmed by current observational results. Although hundreds of objects have been discovered in the outer solar system during the var- ious wide-field surveys carried out during the past decade, the two objects pointed out by Gallardo (2006) remain as the only Uranus' co-orbital candidates identified to date. Calculations by Gallardo (2006) revealed that two objects were moving in a 1:1 mean motion resonance with Uranus. One of them, 83982 Crantor (2002 GO9), is the main object of study of this paper. Since 2006, the orbit of this object has been im- proved and here we make use of the most updated ephemerides to reassess the current dynamical status of this minor body. In this paper, we use N-body simulations to confirm the co-orbital nature with Uranus of the asteroid (83982) Crantor, currently classified as Centaur by both the Minor Planet Center (MPC) and the Jet Propulsion Laboratory (JPL). The numerical model is described in the next section and available data on (83982) Crantor are presented in Section 3. The results of our N-body calculations are shown in Section 4. These results are discussed in Section 5. A new Uranus' horseshoe librator candidate is presented in Section 6 and our conclusions are summarized in Section 7. 2. Numerical integration The orbital evolution of 83982 Crantor (2002 GO9) was com- puted for 0.5 Myr forward and backward in time using the Hermite integration scheme described by Makino (1991) and implemented by Aarseth (2003). This N-body code has been extensively tested by the authors and used in a variety of re- cent solar system numerical studies (de la Fuente Marcos & de la Fuente Marcos 2012a,b,c,d). The standard version of this se- quential code is publicly available from the IoA web site1. Our integrations include the perturbations by the eight major planets, the Moon, the barycentre of the Pluto-Charon system, and the three largest asteroids. For accurate initial positions and veloc- ities we used the heliocentric ecliptic Keplerian elements pro- vided by the Jet Propulsion Laboratory on-line solar system data service2 (Giorgini et al. 1996) and initial positions and veloci- ties based on the DE405 planetary orbital ephemerides (Standish 1998) referred to the barycentre of the solar system. In addition to the orbital calculations completed using the nominal elements in Table 1, we have performed 50 control simulations with sets of orbital elements obtained from the nominal ones and the quoted uncertainties (3-σ). The derived sample of control orbits follows a Gaussian distribution in the 6-dimensional space of orbital el- ements and they are compatible with the observations within the 3-σ uncertainties. The analysis of the control orbits provides some insight on the predictability of the trajectory of the object. The numerical rather than analytical approach to the study of this object is more appropriate because planets other than Uranus, as we will discuss later, play a role on the dynamics of this aster- oid, rendering the traditional perturbational approach within the framework of the three-body problem particularly limited in this case. There exists practically no analytical means to study horse- shoe orbits. 1 http://www.ast.cam.ac.uk/sverre/web/pages/nbody. htm 2 http://ssd.jpl.nasa.gov/?planet_pos 2 Table 1. Heliocentric Keplerian orbital elements of 83982 Crantor (2002 GO9) used in this research. Values include the 1- σ uncertainty. The orbit is based on 104 observations spanning a data-arc of 2,654 days or 7.27 yr, from 2001-03-20 to 2008- 06-25. (Epoch = JD2456200.5, 2012-Sep-30.0; J2000.0 ecliptic and equinox. Source: JPL Small-Body Database.) semi-major axis, a eccentricity, e inclination, i longitude of the ascending node, Ω argument of perihelion, ω mean anomaly, M perihelion, q aphelion, Q absolute magnitude, H = = = = = = = = = 19.3553±0.0014 AU 0.27496±0.00004 12.78489±0.00003 ◦ 117.4097±0.0003 ◦ 92.599±0.003 ◦ 43.964±0.006 ◦ 14.0333±0.0002 AU 24.677±0.002 AU 8.5±0.8 mag 3. 83982 Crantor (2002 GO9) in perspective 83982 Crantor (2002 GO9) was discovered on April 12, 2002 by E. F. Helin, S. Pravdo, K. Lawrence, M. Hicks and R. Thicksten working for the Near-Earth Asteroid Tracking (NEAT) project at Palomar Observatory (Gilmore et al. 2002). It was originally reported as a scattered disk object with a = 54.24 AU and e = 0.81 but soon after, a number of precovery images of the object were uncovered: it first appears in images obtained on March 20, 2001 from the Air Force Maui Optical and Supercomputing (AMOS) observatory located at the summit of Haleakala, then on images acquired on April 16, 2001 from the Apache Point Observatory as part of the Sloan Digital Sky Survey (SDSS), and again on new images obtained on March 26, 2001 and January 15, 2002 from Haleakala-AMOS (Ticha et al. 2002). All this ob- servational material enabled the computation of a reliable orbit characterized by a value of the semi-major axis (19.36 AU) close to that of Uranus, significant eccentricity (∼0.3), and moderate inclination (∼13◦). Therefore, its orbit is now relatively well de- termined with 104 observations spanning a data-arc of 2,654 days and it is clearly not compatible with that of a scattered disk object. Consistently, (83982) Crantor is currently listed by both the Minor Planet Center (MPC) Database3 and the JPL Small- Body Database4 as a Centaur. Early photometric work (Tegler et al. 2003) pointed out the red surface color of (83982) Crantor and suggested that this object and many others like it were formed farther away from the Sun than distant minor bodies characterized by grey sur- face color. In fact, (83982) Crantor is one of the reddest ob- jects of the solar system, close to fellow ultra-red Centaur (5145) Pholus; its surface should be partially covered by tholins in or- der to explain its redness and low albedo (Cruikshank et al. 2007). The rotation period of (83982) Crantor is 6.97 h or 9.67 h with a light-curve amplitude of 0.14 mag (Ortiz et al. 2003). (83982) Crantor was observed with the near-infrared integral field spectrograph SINFONI at the Very Large Telescope (VLT) that found evidence of an absorption feature in its spectrum at 2.0 µm, probably associated with water ice and another feature at 2.3 µm which could be associated with methanol (Alvarez- Candal et al. 2007). These results confirmed previous hints ob- tained by Doressoundiram et al. (2005). Visible spectra further support the very red nature of the object (Alvarez-Candal et al. 2008). Additional near-infrared spectra taken with the Keck I Telescope confirmed previous results (Barkume et al. 2008). 3 http://www.minorplanetcenter.net/db_search 4 http://ssd.jpl.nasa.gov/sbdb.cgi C. de la Fuente Marcos and R. de la Fuente Marcos: Crantor, a horseshoe companion to Uranus Near-infrared photometry was obtained by Doressoundiram et al. (2007). Incomplete photometry was obtained with FORS1 at VLT (DeMeo et al. 2009). Recent Hubble Wide Field Camera 3 results (Fraser & Brown 2012) validate previous findings, in- dicating the presence of irradiated organics and tholins on its surface. The asteroid (83982) Crantor is relatively large. The object has a diameter of < 66.7+18.7−19.6 km with visible geometric albedo of 8.60+8.62−3.36% (Stansberry et al. 2008). Its period of revolution around the Sun, approximately 85.15 yr at present, is very close to that of Uranus, 84.32 yr. As a result, (83982) Crantor and Uranus appear to follow each other in their paths around the Sun, although (83982) Crantor's orbital plane is currently tilted to that of the Earth by 12.8◦ (Uranus' is 0.8◦). Its dynamical half-lives have been estimated to be 2.93 Myr (for forward inte- gration) and 3.67 Myr (for backward integration) by Horner et al. (2004). (83982) Crantor was originally proposed as a possi- ble co-orbital of Uranus together with 2000 SN331 by Gallardo (2006). Both objects would be following horseshoe trajectories. In the same research work, it is pointed out that Uranus' Trojans are affected by high order resonances with Saturn. The asteroid 2000 SN331 has not been reobserved since its discovery and its orbit remains very poorly known (see below). In the following section we focus on the dynamical evolution of (83982) Crantor. 4. Dynamical evolution In order to study the librational properties of 83982 Crantor (2002 GO9) and following the work of Mikkola et al. (2006), we define the relative deviation of the semi-major axis from that of Uranus by α = (a − aU)/aU, where a and aU are the semi-major axes of the object and Uranus, respectively, and also the rela- tive mean longitude λr = λ − λU, where λ and λU are the mean longitudes of the object and Uranus, respectively. If λr oscillates around 0◦, the object is considered a quasi-satellite; Trojan bod- ies are characterized by λr oscillating around +60◦ (L4 Trojan) or -60◦ (or 300◦, L5 Trojan); finally, an object librating with am- plitude > 180◦ follows a horseshoe orbit (see, e.g., Murray & Dermott 1999). Our N-body calculations confirm that (83982) Crantor cur- rently is a co-orbital companion to Uranus and follows a horse- shoe orbit, all in agreement with what was originally pointed out by Gallardo (2006), see Fig. 1. The apparent overlap with Uranus' position in Fig. 1 is the result of the moderate orbital in- clination of the object. The orbital behaviour of (83982) Crantor is illustrated by the animation displayed in Fig. 2 (available on the electronic edition as a high resolution animation or embed- ded at lower resolution in the pdf file associated to this paper). The orbit is presented in three frames of reference: heliocen- tric (left), co-rotating with Uranus (top-right) and Uranocentric (bottom-right). (83982) Crantor moves in a non-regular horse- shoe orbit with a period of about 8500 years. Non-regular means that compound horseshoe-quasi-satellite loops are possible. The dynamical evolution of an object moving in a horseshoe orbit as- sociated to Uranus can be decomposed into a slow guiding cen- tre motion and a superimposed short period three-dimensional epicyclic motion viewed in a frame of reference co-rotating with Uranus. The object spirals along Uranus' orbit at a rate of nearly 0.◦08 per year, but each time it gets close to Uranus is effectively repelled by the planet. The reversals of the net motion of the object with respect to Uranus are obvious in the accompanying animation and given the fact that they take place when the dis- tance to Uranus is the smallest, the gravitational interaction with Fig. 1. The motion of 83982 Crantor (2002 GO9) over the time range (-3, 5) kyr is displayed projected onto the ecliptic plane in a coordinate system rotating with Uranus. The orbit and the position of Uranus are also indicated. In this frame of refer- ence and as a result of its non-negligible eccentricity, Uranus describes a small ellipse. The trajectory of the object spiralizes along Uranus' orbit at a rate of nearly 0.◦08 per year. Uranus is, at that moment, the strongest. Although the inclina- tion of the asteroid is high enough to avoid close encounters with Uranus when the relative mean longitude approaches zero, these close encounters that can only occur in the vicinity of the nodes play a major role on the activation and deactivation of the horse- shoe behaviour of this object (see below). The relative deviation of the semi-major axis, α, as a func- tion of the relative mean longitude, λr, is displayed in Fig. 3, for selected time intervals. The relative mean longitude of (83982) Crantor librates around the unstable Lagrangian point L3 at 180◦ with large amplitude, allowing the object to come quite close to Uranus and to suffer destabilizing close encounters. Typical horseshoe behaviour is observed but brief (half a loop) quasi- satellite episodes also take place. During these events, the ob- ject still moves in a 1:1 commensurability with Uranus but λr librates about 0◦. This exchange between horseshoe and quasi- satellite paths (or compound horseshoe-quasi-satellite orbits) is observed in other horseshoe librators, for example 3753 Cruithne (1986 TO) (Wiegert et al. 1997, 1998). A plot of the orbital elements of (83982) Crantor over a 100 kyr interval centred on the present is shown in Fig. 4. The dis- tance of (83982) Crantor from Uranus displayed in Fig. 4, panel A shows that the object undergoes close encounters with Uranus. A very close encounter, almost a collision at 0.17 AU or 0.37 Hill radii (which is 0.45 AU for Uranus), took place 19248 yr ago. The evolution of λr in panel B indicates that the current horseshoe episode will end in about 20000 yr from now; (83982) Crantor will decouple from Uranus with λr circulating not librat- ing. The semi-major axis exhibits an oscillatory behaviour that is characteristic of the effects of a 1:1 mean motion resonance. The eccentricity decreases by 10% during the very close encounter 3 -30-20-10 0 10 20 30X (AU)-30-20-10 0 10 20 30Y (AU) C. de la Fuente Marcos and R. de la Fuente Marcos: Crantor, a horseshoe companion to Uranus Play Pause Fig. 2. Three-dimensional evolution of the orbit of 83982 Crantor (2002 GO9) in three different frames of reference: heliocentric (left), frame co-rotating with Uranus but centered on the Sun (top-right), and Uranocentric (bottom-right). The red point is (83982) Crantor, the green one is Uranus, and the yellow one is the Sun. The osculating orbits are outlined and the viewing angle changes slowly to make easier for the reader to visualize the orbital evolution. pointed out above and then increases again after nearly 19000 yr, see Fig. 4, panel D. During the horseshoe episode, the orbital inclination remains in the interval (12, 13)◦, see Fig. 4, panel E. The argument of perihelion, Fig. 4, panel F, circulates. In Fig. 4, panel B, we also show the evolution of the relative mean longi- tude for a particular control orbit that has been chosen close to the 3-σ limit; in this way, its orbital elements are most different from the nominal ones in Table 1. For this particular orbit and prior to entering the horseshoe dynamical state, the object was an L4 Trojan. All the control calculations indicate that (83982) Crantor has been co-orbital with Uranus for at least 100 kyr. 5. Discussion The characteristic e-folding time of 83982 Crantor (2002 GO9) during the present horseshoe dynamical state has been found to be of order of 1 kyr. This value gives the reader the idea of how small the timescale required for two initially infinitesimally close trajectories to separate significantly is in this case. An ad- ditional test, that confirms how sensitive to small changes the dynamical evolution of (83982) Crantor is, can be obtained by repeating the calculations, this time excluding one of the three largest asteroids; the orbit significantly diverges from the stan- dard one after just a few 10 kyr. Therefore, simulations over long timescales (e.g. 1 Myr) are not appropriate in this case and be- cause of that, we restrict our figures to a few 10 kyr. Since the orbit of the asteroid is chaotic, its true phase-space trajectory will diverge exponentially from that obtained in our calculations. However, the evolution of the control orbits exhibits very simi- lar secular behaviour of the orbital elements in the time inter- val (-10, 10) kyr. We also note that at the ends of the interval, a close encounter between the asteroid and Uranus happened. Therefore, the dynamical evolution of (83982) Crantor as de- scribed by our integrations can be considered reliable within that relatively short time interval but outside, we should regard our results as an indication of the probable dynamical behaviour of the object. Our calculations show that the orbit of (83982) Crantor is predictable only within a relatively short time interval. We con- sider that an orbit is predictable when all the control orbits give comparable results. It is clear that the current horseshoe be- haviour is not stable on a timescale longer than a few 10 kyr. The analysis of the control orbits indicates that this object may have been co-orbital for less than 200 kyr and it may leave Uranus' co-orbital region in less than 100 kyr. However, some control or- bits remain in the co-orbital region for about 1 Myr. About half the studied control orbits started co-orbital motion about 10-20 kyr ago but one third left the circulation regime to become co- orbitals about 80 kyr ago. The rest appear to have been switching between the various co-orbital states for several 100 kyr. The 4 C. de la Fuente Marcos and R. de la Fuente Marcos: Crantor, a horseshoe companion to Uranus Fig. 3. Resonant evolution of the asteroid 83982 Crantor (2002 GO9). The relative deviation of its semi-major axis from that of Uranus, α, as a function of the relative mean longitude, λr, during the time intervals (-22, -10) kyr (black squares) and (2, 12) kyr (dots) is displayed. The first interval covers the close encounter observed in Fig. 4, panel A (see the text). The second interval lasts an entire cycle of the horseshoe orbit and includes a quasi-satellite half libration, i.e., λr goes beyond -60◦, passes 0◦ but does not reach 60◦. changes in dynamical state are always associated to close en- counters with Uranus. Regarding the future orbital evolution of (83982) Crantor, the majority (70%) of control orbits continue in the horseshoe state for about 15 kyr. The remaining orbits di- vide evenly among those lasting less and those lasting more than 15 kyr. Twenty per cent of control orbits leave the co-orbital re- gion permanently after ending their horseshoe stage but nearly 70% return after 5 to 15 kyr. The rest switch to the quasi-satellite phase. Out of the returning co-orbitals, the vast majority make a comeback as horseshoe librators with only 15% becoming quasi- satellites or Trojans. Nearly 25% continue switching between the various co-orbital states for several 100 kyr and the rest leave the co-orbital region after nearly 80 kyr. Among the co-orbital states, the vast majority of the episodes produce irregular horseshoe or- bits and they also last longer; quasi-satellite events are a distant second and tadpole orbits are observed in just a few cases. Most discrete episodes last for less than 10 kyr. As the reader can see, it is rather difficult to provide a clean picture of the medium-term past and future of this object. The amount of time spent in the horseshoe state prior to the origin of time considered in this re- search is similar in all the control models and in most of them the relative mean longitude circulates before entering and after leaving the state but the details beyond 20 kyr into the past and after 30 kyr into the future are quite heterogeneous both in terms of the actual type of orbital behaviour and its duration. The asteroid 83982 Crantor (2002 GO9) follows an eccen- tric orbit (e ≈ 0.3) but it currently crosses only the orbit of Uranus. In general, minor bodies that cross the paths of one or more planets can be rapidly destabilized by scattering resulting from close planetary approaches if their orbital inclinations are small. (83982) Crantor moves in a relatively highly inclined or- bit (i ≈ 13◦). In the solar system and for a minor body moving in an inclined orbit, close encounters with major planets are only possible in the vicinity of the nodes. The distance between the Fig. 4. Time evolution of various parameters. The distance of 83982 Crantor (2002 GO9) from Uranus (panel A); the value of the Hill sphere radius of Uranus, 0.447 AU, is displayed. The resonant angle, λr (panel B) for the nominal orbit in Table 1 (thick line) and one of the control orbits (thin line). This partic- ular control orbit has been chosen close to the 3-σ limit so its orbital elements are most different from the nominal ones. The orbital elements a (panel C) with the current value of Uranus' semi-major axis, e (panel D), i (panel E), and ω (panel F). Sun and the nodes is given by r = a(1 − e2)/(1 ± e cos ω), where the "+" sign is for the ascending node and the "-" sign is for the descending node. Figure 5 shows the evolution of the distance to the nodes of (83982) Crantor in the time range (-50, 50) kyr. The evolution of the orbital elements in Fig. 4 shows that changes in ω dominate those in a and e for (83982) Crantor and largely con- trol the positions of the nodes. The current precession rate of the nodes is nearly +0.◦2 per century. This value decreases when λr circulates. We found that, in accordance with theory (Namouni 1999), if (83982) Crantor moves in a horseshoe orbit, ω > 0 but during the brief quasi-satellite loops, the value of the argument 5 -0.02-0.01 0 0.01 0.02-150-100-50 0 50 100 150αλ-λU (o) 0 50 100 150 200 250 300 350-40-20 0 20 40ω (o)time (kyr, 0 = 2012-Sep-30.0)F 6 7 8 9 10 11 12 13 14 15i (o)E 0.25 0.26 0.27 0.28 0.29 0.3 0.31 0.32 0.33eD 19 19.5 20a (AU)C-150-100-50 0 50 100 150λr (o)B 0.1 1 10distance Uranus-Crantor (AU)A C. de la Fuente Marcos and R. de la Fuente Marcos: Crantor, a horseshoe companion to Uranus precession of the nodes is the mechanism by which minor plan- ets are placed or removed from horseshoe orbits (Wiegert et al. 1998). On the other hand, the repetitive episodes described above (see also Fig. 4, panel B) in which (83982) Crantor's rela- tive mean longitude librates for several cycles, then circulates for a few more cycles before restarting libration once again are char- acteristic of a type of dynamical behaviour known as resonance angle nodding (Ketchum et al. 2013). These authors conclude that nodding often occurs when a small body is in an external (near) mean motion resonance with a larger planet. This type of complicated dynamics has been observed in other horseshoe librators. Gallardo (2006) suggested that two objects were trapped in a 1:1 mean motion resonance with Uranus, both moving on horse- shoe orbits: (83982) Crantor and 2000 SN331. We have just con- firmed that (83982) Crantor actually moves in a relatively short- lived horseshoe orbit associated to Uranus. However, we cannot do the same with the other object, 2000 SN331. Our calculations indicate that this object is not co-orbital but may move in a tem- porary 10:9 inner resonance with Uranus, i.e., the object com- pletes 10 revolutions around the Sun while Uranus goes around the Sun just 9 times. This result is robust, in principle, but the reliability of the orbit of this object is extremely poor as it is based on just 6 observations with a data-arc span of only 1 day. Therefore, we may say that it is a candidate to move in resonance with Uranus but not a co-orbital. Given the significant destabilizing role played by Saturn on the orbital evolution of (83982) Crantor, the question concerning the existence of long-term Uranus co-orbitals moving on stable orbits takes a new twist, are they at all possible in the light of our present results? In the following section, we present early results on an object that, because of moving in a low eccentricity orbit, may be able to survive as Uranus co-orbital for a longer period of time. In any case, it must be pointed out that longer calculations (several Myr long) suggest that (83982) Crantor may become (600 kyr from now) a long-term Uranus' L5 Trojan. However and given the chaotic nature of the orbit of this object, the reader should take this result with caution. 6. 2010 EU65: a promising Uranus horseshoe candidate Asteroid 2010 EU65 was discovered on March 13, 2010 in im- ages obtained by the European Southern Observatory (ESO) at La Silla (Rabinowitz et al. 2012). The object was reobserved in the following days from the same location and also from Cerro Tololo Observatory at La Serena (Rabinowitz et al. 2010). In to- tal, 26 observations with a data-arc span of 85 days. At the time of discovery its apparent magnitude in R was estimated to be 21.2. The Heliocentric Keplerian orbital elements of 2010 EU65 appear in Table 2. As a relatively recent discovery, its orbit is poorly constrained and it is included here mainly to encourage follow-up observations. Little is known of the physical proper- ties of this object with the exception of its absolute magnitude of 9.1. This probably suggests a medium-sized object with an estimated diameter in the range 28-90 km, for an albedo range of 0.5-0.05. Its period, 84.16 yr, matches well that of Uranus, 84.32 yr, so it appears to follow a 1:1 resonant orbit with Uranus yet it is classified as a Centaur by the MPC. In general and with the exception of the inclination, its orbit is remarkably similar to that of Uranus. Our calculations for the nominal orbit in Table 2 indicate that 2010 EU65 also follows a horseshoe orbit, this time very regular, associated to Uranus. In this case and due to its small eccentricity, the orbit is far more stable than that of Fig. 5. Heliocentric distance to the descending (thick line) and ascending nodes (dotted line) of 83982 Crantor (2002 GO9). Saturn's aphelion, Uranus' semi-major axis, and Neptune's per- ihelion distances are also shown. Both nodal distances are at present relatively close to the value of Uranus' semi-major axis which explains why close encounters with Uranus are possible and relatively frequent, but most of the time the asteroid remains at a safe distance from Uranus. of perihelion decreases (see Fig. 4, panel F). The values of the nodal distances are currently very close to the value of the semi- major axis of Uranus. Close encounters are therefore possible at both nodes doubling the probability of having a strong grav- itational interaction with Uranus that may significantly change the orbit. When the object is east of Uranus, encounters occur at the ascending node. In contrast, close encounters take place at the descending node when (83982) Crantor approaches Uranus from the west. In this way, the gravitational perturbations from Uranus are most effective and both Saturn and Neptune are sec- ondary perturbers for this object. Currently and as a result of its horseshoe trajectory, the object approaches Uranus every 4354 yr although most of the time the asteroid remains at a safe dis- tance from Uranus. Although (83982) Crantor's eccentricity is not large enough to cross Saturn's orbit, Saturn appears to play a non-negligible role in destabilizing the object's motion. Even if close encoun- ters with Saturn and Neptune are not possible, this object (to- gether with Uranus) currently moves in near resonance with the other three giant planets: 1:7 with Jupiter, 7:20 with Saturn, and 1:2 with Neptune. In order to study the role of these three giant planets on the resonant evolution of (83982) Crantor, we have re- peated the calculations considering negligible masses for Jupiter, Saturn and Neptune. This alteration removes the associated near resonance. The "toy model" with no Neptune gives very similar results for the evolution of this minor body. Therefore, the role of the near resonance with Neptune can be regarded as negligible. In sharp contrast, the cases of the toy models with no Jupiter or no Saturn are dramatically different. The lack of Jupiter has im- mediate effects on the results because it amplifies the dynamical effect of Saturn but the absence of Saturn has major effects on the overall orbital evolution of (83982) Crantor. The precession rate decreases by 30% and the object remains as a complex Uranus horseshoe librator for several 100 kyr. Though (83982) Crantor moves primarily under the influence of the Sun and Uranus, Saturn (mainly), Jupiter and Neptune play an important role by influencing, through torque-induced precession, the position of the asteroid's nodes. Variations in the nodal distance strongly affect the interaction of (83982) Crantor with Uranus and may change or terminate the horseshoe orbit currently observed. This 6 10 12 14 16 18 20 22 24 26 28 30-40-20 0 20 40distance (AU)time (kyr, 0 = 2012-Sep-30.0) C. de la Fuente Marcos and R. de la Fuente Marcos: Crantor, a horseshoe companion to Uranus Table 2. Heliocentric Keplerian orbital elements of 2010 EU65 used in this research. The orbit is based on 26 observations spanning a data-arc of 85 days, from 2010-03-13 to 2010-06- 06. (Epoch = JD2455300.5, 2010-Apr-14.0; J2000.0 ecliptic and equinox. Source: JPL Small-Body Database.) semi-major axis, a eccentricity, e inclination, i longitude of the ascending node, Ω argument of perihelion, ω mean anomaly, M perihelion, q aphelion, Q absolute magnitude, H = = = = = = = = = 19.2041 AU 0.05402 14.8382 ◦ 4.61965 ◦ 180.824 ◦ 0.37496 ◦ 18.1668 AU 20.2414 AU 9.1 mag 83982 Crantor (2002 GO9); now the effect of Saturn is rather negligible and the object always remains at a safe distance from Uranus. The immediate future orbital evolution of 2010 EU65 is illustrated by the animation displayed in Fig. 6 (available on the electronic edition as a high resolution animation or embed- ded at lower resolution in the pdf file associated to this paper). As in the case of (83982) Crantor, the orbit is presented in three frames of reference: heliocentric (left), co-rotating with Uranus (top-right) and Uranocentric (bottom-right). Our numerical in- tegrations suggest that, in sharp contrast with (83982) Crantor, this object remains in co-orbital motion with Uranus for Myr timescales. Due to its poorly known orbit, we must insist that the object is a mere horseshoe librator candidate (and in dire need of follow-up observations) although all the studied control orbits (with errors below 1%) give consistent results. 7. Conclusions In this paper we have analyzed the orbital behaviour of Uranus' current horseshoe librator 83982 Crantor (2002 GO9) numeri- cally in order to better understand its current dynamical status, past dynamics and future evolution as well as to gain some in- sight about its stability. We have also shown that the evolution of this object is mainly controlled by the Sun and Uranus. In fact, close encounters with Uranus generate instability in its orbit and throw the object in and out of the horseshoe dynamical state. (83982) Crantor is remarkable in several respects: it is the first known minor body to be trapped in a 1:1 mean motion reso- nance with Uranus; it currently moves in a complex, horseshoe- like orbit when viewed in a frame of reference co-rotating with Uranus; and it could be the "Rosetta Stone" for understanding why the overall number of Uranus co-orbitals appears to be sig- nificantly below that of Jupiter or Neptune. The object is placed and removed from its horseshoe orbit by the mechanism of the precession of the nodes. This precession is accelerated by the perturbative effects of Saturn. The chaotic nature of the orbit of this object constraints the degree of predictability of its dy- namical evolution on timescales longer than a few 10 kyr. This strongly suggests that its dynamical age is much shorter than that of the solar system; therefore, (83982) Crantor is unlikely to be a member of a hypothetical primordial population of objects moving in a 1:1 mean motion resonance with Uranus. Horner & Evans (2006) claimed that Uranus cannot currently efficiently trap objects in the 1:1 commensurability even for short periods of time. Our results suggest that, contrary to this view and in spite of the destabilizing role of Saturn, Uranus still can actively capture temporary co-orbitals. Regarding the issue of stability, (83982) Crantor's orbital inclination is close to the edge of one of the stability islands in i identified by Dvorak et al. (2010) but 2010 EU65 moves outside the stability islands proposed in that study yet it seems to be more stable than (83982) Crantor. This comparatively better stability strongly suggests that not only in- clination but also eccentricity play an important role on the long- term dynamics of these objects. In summary, key questions still remain open and further work is necessary to better understand the complex issue of the stability of Uranian co-orbitals. Acknowledgements. The authors thank the referee for his/her helpful sugges- tions regarding the presentation of this paper. The authors would like to thank S. J. Aarseth for providing the code used in this research. This work was par- tially supported by the Spanish 'Comunidad de Madrid' under grant CAM S2009/ESP-1496 (Din´amica Estelar y Sistemas Planetarios). We thank M. J. Fern´andez-Figueroa, M. Rego Fern´andez and the Department of Astrophysics of Universidad Complutense de Madrid (UCM) for providing excellent computing facilities. Most of the calculations and part of the data analysis were completed on the 'Servidor Central de C´alculo' of the UCM and we thank S. Cano Als´ua for his help during that stage. In preparation of this paper, we made use of the NASA Astrophysics Data System and the ASTRO-PH e-print server. References Aarseth, S. J. 2003, Gravitational N-Body Simulations, Cambridge University Press, Cambridge, p. 27 Alvarez-Candal, A., Barucci, M. A., Merlin, F., Guilbert, A., & de Bergh, C. 2007, A&A, 475, 369 Alvarez-Candal, A., Fornasier, S., Barucci, M. A., de Bergh, C., & Merlin, F. 2008, A&A, 487, 741 Bailey, B. L., & Malhotra, R. 2009, Icarus, 203, 155 Barkume, K. M., Brown, M. E., & Schaller, E. L. 2008, AJ, 135, 55 Brasser, R., Innanen, K. A., Connors, M., et al. 2004, Icarus, 171, 102 Brown, E. W. 1911, MNRAS, 71, 438 Christou, A. A., & Asher, D. J. 2011, MNRAS, 414, 2965 Connors, M., Veillet, C., Brasser, R., et al. 2004, Meteoritics Planet. Sci., 39, 1251 Connors, M., Stacey, G., Brasser, R., & Wiegert P. 2005, Plan. Space Sci., 53, 617 Cruikshank, D. P., Barucci, M. A., Emery, J. P., Fern´andez, Y. R., Grundy, W. M., Noll, K. S., Stansberry, J. A. 2007, in Protostars & Planets V, ed. B. Reipurth, D. Jewitt, K. Keil, University of Arizona Press, Tucson, 879 Darwin, G. 1912, MNRAS, 72, 642 DeMeo, F. E., Fornasier, S., Barucci, M. A., et al. 2009, A&A, 493, 283 Dermott, S. F., & Murray, C. D. 1981a, Icarus, 48, 1 Dermott, S. F., & Murray, C. D. 1981b, Icarus, 48, 12 Doressoundiram, A., Barucci, M. A., Tozzi, G. P., et al. 2005, Plan. Space Sci., 53, 1501 Doressoundiram, A., Peixinho, N., Moullet, A., Fornasier, S., Barucci, M. A., Beuzit, J.-L., & Veillet, C. 2007, ApJ, 134, 2186 Dvorak, R., Schwarz, R., Suli, ´A., & Kotoulas, T. 2007, MNRAS, 382, 1324 Dvorak, R., Bazs´o, ´A, & Zhou, L.-Y. 2010, Celest. Mech. Dyn. Astron., 107, 51 de la Fuente Marcos, C., & de la Fuente Marcos, R. 2012a, A&A, 547, L2 de la Fuente Marcos, C., & de la Fuente Marcos, R. 2012b, A&A, 545, L9 de la Fuente Marcos, C., & de la Fuente Marcos, R. 2012c, MNRAS, 427, 728 de la Fuente Marcos, C., & de la Fuente Marcos, R. 2012d, MNRAS, 427, L85 Fraser, W. C., & Brown, M. E. 2012, ApJ, 749, 33 Gallardo, T. 2006, Icarus, 184, 29 Garfinkel, B. 1977, AJ, 82, 368 Giacaglia, G. E. O. 1970, in Periodic Orbits, Stability and Resonances, ed. G. E. O. Giacaglia, Dordrecht, Reidel, 515 Gilmore, A. C., Pravec, P., Helin, E. F., et al. 2001, MPEC 2002-H03 Giorgini, J. D., Yeomans, D. K., Chamberlin, A. B., et al. 1996, BAAS, 28, 1158 Holman, M. J., & Wisdom, J. 1993, AJ, 2015, 1987 Horner, J., & Evans, N. 2006, MNRAS 367, L20 Horner, J., Evans, N. W., & Bailey, M. E. 2004, MNRAS 354, 798 Karlsson, O. 2004, A&A, 413, 1153 Ketchum, J. A., Adams, F. C., & Bloch, A. M. 2013, ApJ, 762, 71 Makino, J. 1991, ApJ, 369, 200 Marzari, F., Tricarico, P., & Scholl, H. 2003, A&A, 410, 725 Michel, P., Froeschl´e, C., & Farinella, P. 1996, A&A, 313, 993 Mikkola, S., Innanen, K., Wiegert, P., Connors, M., & Brasser, R. 2006, MNRAS, 369, 15 7 C. de la Fuente Marcos and R. de la Fuente Marcos: Crantor, a horseshoe companion to Uranus Fig. 6. Three-dimensional evolution of the orbit of 2010 EU65 in three different frames of reference: heliocentric (left), frame co- rotating with Uranus but centered on the Sun (top-right), and Uranocentric (bottom-right). The orange point represents 2010 EU65, the green one is Uranus, and the yellow one is the Sun. The osculating orbits are outlined and the viewing angle changes slowly to make easier for the reader to visualize the orbital evolution. Play/Pause Milani, A., Carpino, M., Hahn, G., & Nobili, A. M. 1989, Icarus, 78, 212 Murray, C. D., & Dermott, S. F. 1999, Solar System Dynamics, Cambridge University Press, Cambridge, p. 97 Namouni, F. 1999, Icarus, 137, 293 Nesvorn´y, D., & Dones, L. 2002, Icarus, 160, 271 Ortiz, J. L., Guti´errez, P. J., Casanova, V., & Sota, A. 2003, A&A, 407, 1149 Rabe, E. 1961, AJ, 66, 500 Rabinowitz, D., Tourtellotte, S., & Marsden, B. G. 2010, MPEC 2010-H80 Rabinowitz, D., Schwamb, M. E., Hadjiyska, E., Tourtellotte, S. 2012, AJ, 144, 140 Smith, B. A., Reitsema, H. J., Fountain, J. W., & Larson, S. M. 1980, BAAS, 12, 727 Standish, E. M. 1998, JPL Planetary and Lunar Ephemerides, DE405/LE405, Interoffice Memo. 312.F-98-048, Jet Propulsion Laboratory, Pasadena, California Stansberry, J., Grundy, W., Brown, M., Cruikshank, D., Spencer, J., Trilling, D., & Margot, J.-L. 2008, in The Solar System Beyond Neptune, ed. M. A. Barucci, H. Boehnhardt, D. P. Cruikshank, A. Morbidelli, University of Arizona Press, Tucson, 161 Synnott, S. P., Peters, C. F., Smith, B. A., & Morabito, L. A. 1981, Science, 212, 191 Tegler, S. C., Romanishin, W., & Consolmagno, G. J. 2003, ApJ, 599, L49 Thuring, B. 1959, AN, 285, 71 Ticha, J., Tichy, M., Haver, R., et al. 2002, MPEC 2002-L55 Wajer, P., & Kr´olikowska, M. 2012, Acta Astronomica, 62, 113 Weissman, P. R., & Wetherill, G. W. 1974, AJ, 79, 404 Wiegert, P., Innanen, K. A., & Mikkola, S. 1997, Nature, 387, 685 Wiegert, P. A., Innanen, K. A., & Mikkola, S. 1998, AJ, 115, 2604 Wiegert, P., Innanen, K., & Mikkola, S. 2000, AJ, 119, 1978 8
1706.08444
3
1706
2018-01-14T08:07:53
The ARIEL Mission Reference Sample
[ "astro-ph.EP" ]
The ARIEL (Atmospheric Remote-sensing Exoplanet Large-survey) mission concept is one of the three M4 mission candidates selected by the European Space Agency (ESA) for a Phase A study, competing for a launch in 2026. ARIEL has been designed to study the physical and chemical properties of a large and diverse sample of exoplanets, and through those understand how planets form and evolve in our galaxy. Here we describe the assumptions made to estimate an optimal sample of exoplanets - including both the already known exoplanets and the "expected" ones yet to be discovered - observable by ARIEL and define a realistic mission scenario. To achieve the mission objectives, the sample should include gaseous and rocky planets with a range of temperatures around stars of different spectral type and metallicity. The current ARIEL design enables the observation of ~1000 planets, covering a broad range of planetary and stellar parameters, during its four year mission lifetime. This nominal list of planets is expected to evolve over the years depending on the new exoplanet discoveries.
astro-ph.EP
astro-ph
Noname manuscript No. (will be inserted by the editor) The ARIEL Mission Reference Sample Tiziano Zingales · Giovanna Tinetti · Ignazio Pillitteri · Jérémy Leconte · Giuseppina Micela · Subhajit Sarkar Received: date / Accepted: date Abstract The ARIEL (Atmospheric Remote-sensing Exoplanet Large-survey) mission concept is one of the three M4 mission candidates selected by the Eu- ropean Space Agency (ESA) for a Phase A study, competing for a launch in 2026. ARIEL has been designed to study the physical and chemical properties of a large and diverse sample of exoplanets, and through those understand how planets form and evolve in our galaxy. Here we describe the assumptions made to estimate an optimal sample of exoplanets -- including both the already known exoplanets and the "expected" ones yet to be discovered -- observable by ARIEL and define a realistic mission scenario. To achieve the mission objectives, the sample should include gaseous and rocky planets with a range of temperatures around stars of different spec- tral type and metallicity. The current ARIEL design enables the observation of ∼1000 planets, covering a broad range of planetary and stellar parameters, Tiziano Zingales University College London INAF- Osservatorio Astronomico di Palermo Giovanna Tinetti University College London Ignazio Pillitteri INAF - Osservatorio Astronomico di Palermo Jérémy Leconte CNRS, Université de Bordeaux Laboratoire d'Astrophysique de Bordeaux Giuseppina Micela INAF - Osservatorio Astronomico di Palermo Subhajit Sarkar Cardiff University, Cardiff, UK 8 1 0 2 n a J 4 1 . ] P E h p - o r t s a [ 3 v 4 4 4 8 0 . 6 0 7 1 : v i X r a 2 Tiziano Zingales et al. during its four year mission lifetime. This nominal list of planets is expected to evolve over the years depending on the new exoplanet discoveries. Keywords Exoplanets · ARIEL space mission · Planetary population 1 Introduction 1.1 Mission overview Today we know over 3500 exoplanets and more than one third are transiting (http://exoplanets.eu/). These include Earths, super-Earths, Neptunes and Giant planets around a variety of stellar types. The Kepler space mission has discovered alone more than 1000 new transiting exoplanets between 2009 and 2015 and more than 3000 still unconfirmed planetary candidates. The number of known exoplanets is expected to increase in the next decade thanks to current and future space missions (K2, GAIA, TESS, CHEOPS, PLATO) and a long list of ground-based surveys (e.g. HAT-NET, HARPS, WASP, MEarth, NGTS, TRAPPIST, Espresso, Carmenes). They will detect thousands of new transiting exoplanets. ARIEL (Atmospheric Remote-sensing Exoplanet Large-survey) is one of the three candidate missions selected by the European Space Agency (ESA) for its next medium-class science mission due for launch in 2026. The goal of the ARIEL mission is to investigate the atmospheres of several hundreds planets orbiting distant stars in order to address the fundamental questions on how planetary systems form and evolve. Key objective of the mission is to find out whether the chemical composition of exoplanetary atmospheres correlate with basic parameters such as the planetary size, density, temperature, and stellar type and metallicity. During its four-year mission, ARIEL aims at ob- serving a statistically significant sample of exoplanets, ranging from Jupiter- and Neptune-size down to super-Earth and Earth-size in the visible and the infrared with its meter-class telescope. The analysis of ARIEL spectra and photometric data will allow to extract the chemical fingerprints of gases and condensates in the planets' atmospheres, including the elemental composition for the most favorable targets. It will also enable the study of thermal and scattering properties of the atmosphere as the planet orbit around the star. The main purpose of this paper is to estimate an optimal list of targets observable by ARIEL or a similar mission in ten years time and quantify a realistic mission scenario to be completed in 4 year nominal mission lifetime, including the commissioning phase. To achieve the mission objectives, the sample should include gaseous and rocky planets with a range of temperatures around stars of different spectral type and metallicity. With this aim, it is necessary to consider both the already known exoplanets and the "expected" ones yet to be discovered. The data col- lected by Kepler allow to estimate the occurrence rate of exoplanets according to their size and orbital periods. Using this planetary occurrence rate and the number density of stars in the Solar neighbourhood, we can estimate the The ARIEL Mission Reference Sample 3 number of exoplanets expected to exist with a particular size, orbital period range and orbiting a star of a particular spectral type and metallicity. Here we describe the assumptions made to estimate an optimal sample of exoplanets observable by ARIEL and define the Mission Reference Sample (MRS). It is clear that this nominal list of planets will change over the years depending on the new exoplanetary discoveries. In Section 2 we explain the method used to estimate the number and the parameters of the planetary systems yet to be discovered. All the potential ARIEL targets will be presented in Section 3, where we show all the planets that can be observed individually during the mission lifetime, and out of which we want to select the optimal sample. Section 4 is dedicated to the selection and description of an ARIEL MRS fulfilling the mission requirements, we compare the proposed ARIEL MRS to the sample expected to be discovered by TESS, confirming that TESS could provide a large fraction of the ARIEL targets. A sample including only planets known today is identified. In Section 5 we show a possible MRS which maximises the coverage of the planetary and stellar physical parameters. 1.2 Description of the models We use the ESA Radiometric Model (Puig et al., 2015) to estimate the per- formances of the ARIEL mission given the planetary, stellar and orbital char- acteristics: namely the stellar type and brightness, the planetary size, mass, equilibrium temperature and atmospheric composition, the orbital period and eccentricity. This tool takes into account the mission instrumental parameters and planetary system characteristics to calculate: -- The SNR (Signal to Noise Ratio) that can be achieved in a single transit; -- The SNR that can be achieved in a single occultation; -- The number of transit/occultation revisits necessary to achieve a specified SNR; -- The total number and types of targets that can be included in the mission lifetime. In this work, the input planet list for the radiometric model is a combina- tion of known and simulated exoplanets, as detailed in the following sections. We used the instrument parameters of the ARIEL payload as designed dur- ing the phase A study. To increase the efficiency of our simulations we used a Python tool as a wrap of the ESA Radiometric Model, so we could test different mission configurations that fulfil the mission science objectives. The results were validated with ExoSim, a time domain simulator used for the ARIEL space mission, but thanks to its modularity it can be used to study any transit spectroscopy instrument from space or ground. ExoSim has been developed by Sarkar et al. (2016); Sarkar and Pascale (2015); Pascale et al. (2015) (see App A). 4 Tiziano Zingales et al. 2 Simulations of planetary systems expected to be discovered in the next decade 2.1 Star count estimate We used the stellar mass function as obtained from the 10-pc RECONS (RE- search Consortium On Nearby Stars) to estimate the number of stars as a func- tion of the K magnitude. We assume mass-luminosity-K magnitude conversions from Baraffe et al. (1998). The same procedure was adopted by Ribas and Lo- vis (2013). The number of main sequence stars with limit K-mag mK = 7 used to infer the number density of stars in the Solar neighbourhood is shown in Tab 1. Spectral type N∗ (K < 7) Mass (M(cid:12)) 1.25 - 1.09 1.09 - 0.87 0.87 - 0.65 0.65 - 0.41 0.41 - 0.22 0.22 - 0.10 F6 - F9 G0 - G8 K0 - K5 K7 - M1 M2 - M3 M4 - late M 5646 3356 1167 386 81 28 Table 1 Star counts considering different spectral types with limiting magnitude mK = 7. The number density and the number of stars are related through Eq 1: ρ∗ = N∗(K < 7) 4 3 πd3 (1) where the distance d has been calculated in the ARIEL Radiometric Model (Puig et al., 2015) using the relation between K magnitude mK and the dis- tance d: mK = −2.5 log R2∗Ss(∆λ) d2SK0 (∆λ) (2) In Eq 2, R∗ is the stellar radius, SK0 (∆λ) is the zero point flux for the standard K-band filter profile, ∆λ is the filter band pass given in Cohen et al. (2003) and Ss(∆λ) the stellar flux density evaluated over the same bandwidth. We neglect the interstellar absorption since our stars are at a relatively short distance. 2.2 Planetary population and occurrence rate In this section we briefly review the current knowledge about the occurrence rate of planets, i.e. the average expected number of planets per star. Fressin et al. (2013) used the Kepler statistics to publish the planetary occurrence rates around F, G, K main sequence stars ordered by orbital periods and planetary types. An accurate planetary occurrence rate is pivotal to the reliability of the The ARIEL Mission Reference Sample 5 Density Star / pc3 ρ(F6-F9) ρ(G0-G8) ρ(K0-K5) ρ(K7-M1) ρ(M2-M3) ρ(M4 - late M) 0.0039 0.0044 0.0049 0.0074 0.0059 0.0118 Table 2 Main sequence star densities considering different spectral types with limiting magnitude mK = 7 estimate of the existing planets in the Solar neighbourhood. We used the plan- etary occurrence rate values for F,G,K and M stars from Fressin et al. (2013), being the most conservative and currently the most complete, i.e. covering all planetary types and stars. Therefore, our estimates for the ARIEL sample are very conservative. Mulders et al. (2016) updated the planetary occurrence rate for planets between 0.5R⊕ and 4R⊕ and orbital period < 50 days, using a more recent list of planets discovered by the Kepler satellite. Fig 2 shows the comparison between Mulders et al. (2016) and Fressin et al. (2013). The differences between the two occurrence rates can be up to an order of mag- nitude. Mulders et al. (2015) show that M stars have 3.5 times more small planets (1.0−2.8R⊕) than FGK stars, but two times fewer Neptune-sized and larger (> 2.8R⊕) planets. The fraction of M-stars considered in our work is only ∼ 7% of the total stellar sample, so we are significantly underestimat- ing the number of small planets around M-dwarfs, which are optimal targets for transit spectroscopy. More recent and complete results from Mulders and collaborators are expected to be published in the next months and they are not yet available for our simulations. Given the discrepancy between Mulders and Fressin's statistics we expect a substantial improvement in our estimates when the most recent Kepler statistics will become available. Fressin et al. (2013) provided the following statistics for different planetary classes: -- Jupiters: 6R⊕ < Rp ≤ 22R⊕ -- Neptunes: 4R⊕ < Rp ≤ 6R⊕ -- Small Neptunes: 2R⊕ < Rp ≤ 4R⊕ -- Super Earths: 1.25R⊕ < Rp ≤ 2R⊕ -- Earths: 0.8R⊕ < Rp ≤ 1.25R⊕ We adopted a size resolution of 1R⊕ in each of these classes. The number of planets can be estimated as: Np = 4 3 πd3ρ∗Pt,pPgeom (3) where d is the radius of a sphere with the Sun at the centre, ρ∗ is the number density of the stars, Pt,p is the probability of having a t-type planet orbit- ing with an orbital period p (See Fig 1). Pgeom = R∗/a is the geometrical probability of a transit. 6 Tiziano Zingales et al. We simulated all the transiting planets in the solar system neighbourhood up to mK = 14, all these planets described by Np constitute the "Mission Reference Population". To include in the population sample the exoplanets known today, every time we predict a system with the same physical properties of a known sys- tem we replace it with the known one. In Sec 3 we show that in the solar system neighbourhood there are ∼ 9500 planets for which the ARIEL science requirements can be reached in less that 6 transits or eclipses. The equilibrium temperature (Eq 4) of the planet can be evaluated as- suming the incoming and outgoing radiation at the planetary surface are in equilibrium: (cid:18) R∗ 2a Tp = T∗ 2(cid:18)1 − A (cid:19) 1 (cid:19) 1 4 ε (4) Here T∗ and R∗ are the stellar temperature and radius, a the semi-major axis of the orbit, A is the planetary albedo and ε is the atmospheric emissivity. The ARIEL space mission will focus on planets with an orbital period shorter than 50 days. As expected, shorter periods mean shorter semi-major axis and, therefore, from Eq 4, typically higher effective temperature. Fig. 1 Average number of planets per star and per size bin with an orbital period shorter than 85 days orbiting around F, G, K stars. The statistics was extracted from the Q1 - Q6 Kepler data (Fressin et al., 2013). The ARIEL Mission Reference Sample 7 Fig. 2 Comparison of three different distributions estimating the planetary occurrence rate as a function of orbital period for planets between 0.5R⊕ and 4R⊕. Blue and green lines: results from Mulders et al. (2016) for two metallicity classes. Red line: results from Fressin et al. (2013). The Fressin et al. (2013) statistics strongly underestimates the occurence of sub Neptune size planets compared to Mulders et al. (2016) and other more recent estimates. The reason is the large number of small planets discovered after 2013. 2.3 Planetary masses and densities To simulate a realistic planetary population we need to consider a distribution of masses given a planetary radius. The planetary mass affects directly the surface gravity and therefore the scale height (H) of the atmosphere: H = k T µ g (5) The mass estimate is not a trivial task, given the range of planetary den- sities observed today. We used a Python tool written by Chen and Kipping (2016) to estimate the mass of all the planets in our simulated sample. In the ARIEL planetary sample there are both known and simulated planets. Chen and Kipping (2016) use the currently known planets to derive the statistical distribution of the mass of a given planet when its radius is known. Thus, for each planet in our initial sample, the mass is randomly drawn following this distribution except for known systems. In Fig 3 we show the mass distribution for all the planets in our simulations. Moreover, as a very few planets have a radius larger than 20R⊕, we use that radius as an upper limit. There is already a well known degeneracy in the 7 − 20R⊕ range: objects with a radius within that range can be planets as well as very cool stars. However, this should not be too concerning, as observations have shown that very short-period, low-mass stellar companions are much less frequent than hot giant planets (Piskorz et al., 2015). 1020304050Period (days)10-410-310-210-1100Planetary occurrence rate (%)0.5R⊕−4.0R⊕Mulders (metal poor star)Mulders (metal rich star)Fressin 8 Tiziano Zingales et al. Fig. 3 Mass-Radius distribution for all the simulated planets. The mass-radius relationship has been calculated with the Chen and Kipping (2016) tool. The ARIEL Mission Reference Sample 9 3 ARIEL science goals and Mission Reference Population 3.1 The 3 tier approach The ARIEL primary science objectives call for atmospheric spectra or photo- metric lightcurves of a large and diverse sample of known exoplanets covering a wide range of masses, densities, equilibrium temperatures, orbital properties and host-stars. Other science objectives require, by contrast, the very deep knowledge of a select sub-sample of objects. To maximise the science return of ARIEL and take full advantage of its unique characteristics, a three-tiered approach has been considered, where three different samples are observed at optimised spectral resolutions, wavelength intervals and signal-to-noise ratios. (a summary of the three-tiers and observational methods is given below in table 3). In this section we present the pool of potential targets that could reach the specifications for each tier in a reasonable number of observations. The number of targets for the various Tiers are shown as a function of planetary radius in Fig 4, 6 and 8 and as a function of effective temperature in 5, 7 and 9. Note that the planets shown in these figures do not represent the final sample, as it would take too long to observe all of them. They are the pool from which the MRS can be selected to best address the scientific questions summarized below. The fact that the number of potential targets is much larger than the number that can be observed illustrates that ARIEL can choose the final sample among a great variety of observable planets, providing a lot a flexibility. Survey (∼37%) Deep (∼60%) Benchmark (∼3%) ARIEL 3-tiers Low Spectral Resolution observations of a large sample of planets in the Vis-IR, with SNR ≥7 Higher Spectral Resolution observations of a sub-sample in the VIS-IR Very best planets, re-observed multiple time with all techniques Table 3 Summary of the survey tiers and the observational methods they will be accom- plished. Each tier is expressed in terms of nominal mission lifetime ARIEL could spend on them. The key questions and objectives of each tier can be summarised as follows (see Tinetti et al., in prep. for further details): Survey: -- What fraction of planets are covered by clouds? -- Tier 1 mode is particu- larly useful for discriminating between planets that are likely to have clear atmospheres, versus those that are so cloudy that no molecular absorp- tion features are visible in transmission. Extremely cloudy planets may be identified simply from low-resolution observations over a broad wavelength 10 Tiziano Zingales et al. range. This preliminary information will therefore allow us to take an in- formed decision about whether to continue the spectral characterization of the planet at higher spectral resolution, and therefore include or not the planet in the Tier 2 sample. -- What fraction of small planets have still hydrogen and helium retained from the protoplanetary disk? -- Primordial (primary atmosphere) atmospheres are expected to be mainly made of hydrogen and helium, i.e. the gaseous composition of the protoplanetary nebula. If an atmosphere is made of heavier elements, then the atmosphere has probably evolved (secondary at- mosphere). An easy way to distinguish between primordial (hydrogen-rich) and evolved atmospheres (metal-rich), is to examine the transit spectra of the planet: the main atmospheric component will influence the atmospheric scale height, thus changing noticeably the amplitude of the spectral fea- tures. This question is essential to understand how super-Earths formed and evolved. -- Can we classify planets through colour-colour diagrams or colour-magnitude diagrams? -- Colour-colour or colour-magnitude diagrams are a traditional way of comparing and categorising luminous objects in astronomy. Simi- larly to the Herzsprung-Russell diagram, which led to a breakthrough in understanding stellar formation and evolution, the compilation of similar diagrams for exoplanets might lead to similar developments (Triaud et al., 2014). -- What is the bulk composition of the terrestrial exoplanets? -- The planetary density may constrain the composition of the planet interior. However this measurement alone may lead to non-unique interpretations (Valencia et al., 2007). A robust determination of the composition of the upper atmosphere of transiting planets will reveal the extent of compositional segregation between the atmosphere and the interior, removing the degeneracy origi- nating from the uncertainty in the presence and mass of their (inflated?) atmospheres. -- What is the energy balance of the planet? -- Eclipse photometric measure- ments in the optical and infrared can provide the bulk temperature and albedo of the planet, thereby allowing the estimation of the planetary en- ergy balance and whether the planet has an internal heat source or not. Deep: A key objective of ARIEL is to understand whether there is a correlation between the chemistry of the planet and basic parameters such as planetary size, density, temperature and stellar type and metallicity. Spectroscopic mea- surements at higher resolution will allow in particular to measure: -- The main atmospheric component for small planets; -- The chemical abundances of trace gases, which is pivotal to understand the type of chemistry (equilibrum/non equilibrium). -- The atmospheric thermal structure, both vertical and horizontal; -- The cloud properties, i.e. cloud particles size and distribution, The ARIEL Mission Reference Sample 11 -- The elemental composition in gaseous planets. This information can be used to constrain formation scenarios (Öberg et al., 2011). Benchmark: A fraction of planets around very bright stars will be observed repeatedly through time to obtain: -- A very detailed knowledge of the planetary chemistry and dynamics; -- An understanding of the weather, and the spatial and temporal variability of the atmosphere. Benchmark planets are the best candidates for phase-curve spectroscopic mea- surements. 3.2 Key science questions In this section we show a full list of potential targets for ARIEL: these are expected to evolve until launch, and will be updated regularly to include new planet discoveries. ARIEL Tier 1 (Survey) will analyse a large sample of exoplanets to address science questions where a statistically significant population of objects needs to be observed. ARIEL Tier 1 will also allow a rapid, broad characterisation of planets permitting a more informed selection of Tier 2 and Tier 3 planetary candidates. For most Tier 1 planetary candidates, Tier 1 performances can be reached between 1 and 2 transits/eclipses. In Fig 4 and 5 we show that in the solar system neighbourhood there are ∼ 9500 observable by ARIEL for which the science requirements can be reached in less than 6 transits or eclipses. ARIEL Tier 2 (Deep, the core of the mission) will analyse a sub-sample of Tier 1 planets with a higher spectral resolution, allowing an optimal charac- terisation of the atmospheres, including information on the thermal structure, abundance of trace gases, clouds and elemental composition. In Fig 6 and 7 we show the properties of all the planetary candidates that could be studied by ARIEL in the Deep mode with a small/moderate number of transit or eclipse events. The third ARIEL Tier (Benchmark, the reference planets) will study the best planets (section 4.3), i.e. the ones orbiting very bright stars which can be studied in full spectral resolution with a relatively small number of tran- sits/eclipses. For the planets observed in benchmark mode in 1 or 2 events, it is possible to study the spatial and temporal variability (i.e. study the weather and evaluate its impact when observations are averaged over time). In Fig 8 and 9 we show the properties of the Tier 3 planetary candidates. 12 Tiziano Zingales et al. Fig. 4 Complete set of Tier 1 planets from the ARIEL missione reference population. The final list of Tier 1 planets will include an optimal sub-sample. Different colours indicate the number of transits/eclipses needed to reach Tier 1 performances. The planets shown here can achieve the Tier 1 requirements combining the signal of ≤ 5 transits/eclipses. Fig. 5 Temperature distribution for the planets illustrated in fig. 4. Fig. 6 Planets from the ARIEL mission reference population in the Deep mode (Tier 2) with a small/moderate number of transits/eclipses, divided in size bins. The final list of Tier 2 planets will include an optimal sub-sample. Different colours indicate the number of transits/eclipses needed to reach Tier 2 performances. Fig. 7 Temperature distribution for the planets illustrated in fig. 6. 01234567891011121314151617181920Radius (R)110100100010000Number of planets9545 planets1 transit2 transits3 transits4 transits5 transits3005007009001100130015001700190021002300Temperature (K)0100200300400500600700800Number of planets1 transit2 transits3 transits4 transits5 transits01234567891011121314151617181920Radius (R)110100100010000Number of planets2869 planets1 transit<= 5 transits<= 10 transits<= 15 transits<= 20 transits<= 25 transits3005007009001100130015001700190021002300Temperature (K)050100150200250Number of planets1 transit<= 5 transits<= 10 transits<= 15 transits<= 20 transits<= 25 transits The ARIEL Mission Reference Sample 13 Fig. 8 Number of planets from the mission reference population observable by ARIEL in the Benchmark mode with a < 25 number of transits/eclipses, divided in size bins. Different colours indicate the number of transits/eclipses needed to reach Tier 3 performances. Fig. 9 Temperature distribution for the planets illustrated in fig. 8. 01234567891011121314151617181920Radius (R⊕)110100100010000Number of planets1594 planets1 transit<= 5 transits<= 10 transits<= 15 transits<=20 transits<= 25 transits3005007009001100130015001700190021002300Temperature (K)020406080100120140Number of planets1 transit<= 5 transits<= 10 transits<= 15 transits<= 20 transits<= 25 transits 14 Tiziano Zingales et al. 4 A possible scenario for the ARIEL space mission In Section 3 we presented a comprehensive list of planet candidates which could be observed with the ARIEL space mission. Here we discuss possible optimisations of the Mission Reference Sample, which ideally should include a large and diverse sample of planets, have the right balance among the three Tiers and, most importantly, must be completed during the nominal mission lifetime (4 years including the commissioning phase). Fig. 10 Overview of the ARIEL MRS, comparing the number of planets observable in the three tiers during the mission lifetime. In Fig 10 we show a possible MRS with all the three tiers nested together. The first MRS shows the maximum number of targets, taking into account the nominal mission lifetime. It has been built starting from all the targets feasible within one transit/eclipse, and adding all the targets that can be done within 2, 3, 4 and so on transits/eclipses in ascending order. This is just one of the possible configurations for the MRS, and one would expect the ARIEL MRS to evolve in response of new exoplanetary discoveries in the next decade. 4.1 MRS Tier 1: Survey Our simulations indicate that the current ARIEL design as presented at the end of the Phase A study, allows to observe 1002 planets in Tier 1. All the planets can be observed in 1538 satellite visits i.e. 37% of the mission lifetime. Most giant planets and Neptunes fulfil the Tier 1 science objectives in 1 tran- sit/eclipse, the smaller planets require up to 6 events (fig. 11 and 12 ). Fig. 13 and 14 illustrate how the 1002 planets are distributed in terms of planetary size, temperature, density and stellar type. The ARIEL Mission Reference Sample 15 Fig. 11 ARIEL MRS Tier 1 planets organised in size-bins. Different colours indicate the number of transits/eclipses needed to reach Tier 1 performances. Fig. 12 ARIEL MRS Tier 1 planets organised in temperature-bins. Different colours indi- cate the number of transits/eclipses needed to reach Tier 1 performances. Fig. 13 ARIEL MRS Tier 1 planets organised in size-bins. Different colours indicate dif- ferences in the simulated planetary densities. Fig. 14 ARIEL MRS Tier 1 planets organised in temperature-bins. Different colours indi- cate differences in the simulated stellar temperatures. 01234567891011121314151617181920Radius (R)1101001000Number of planets1002 planets36.9 %1 transit2 transits3 transits4 transits5 transits6 transits30050070090011001300150017001900210023002500Temperature (K)020406080100Number of planets1 transit2 transits3 transits4 transits5 transits6 transits01234567891011121314151617181920Radius (R)1101001000Number of planetsg/cm30.01 - 0.020.02 - 0.030.03 - 0.040.04 - 0.070.07 - 0.120.12 - 0.190.19 - 0.320.32 - 0.520.52 - 0.850.85 - 1.391.39 - 2.282.28 - 3.733.73 - 6.116.11 - 10.0030050070090011001300150017001900210023002500Planetary Temperature (K)020406080100Number of planetsT*(K)3000 - 33003300 - 36003600 - 39003900 - 42004200 - 45004500 - 48004800 - 51005100 - 54005400 - 57005700 - 60006000 - 63006300 - 6600 16 4.2 MRS Tier 2: Deep Tiziano Zingales et al. The Deep is the core of the mission. Our simulations indicate that the current ARIEL design as presented at the end of the Phase A study, allows to observe ∼ 500 planets in Tier 2 assuming 60% of the mission lifetime. Fig. 17 and 18 illustrate how the 500 planets are distributed in terms of planetary size, temperature, density and stellar type. Most gaseous planets fulfil the Tier 2 science objectives in less than five transits/eclipses, the small planets require up to twenty events (fig. 15 and 16 ). We included a variety of planets from cold (300 K) to very hot (2500 K) as shown in Fig 16. We scheduled also ∼ 50 planets that will be studied with both transit and eclipse methods, indicated by stripes in Fig 15). These are the best candidates for phase-curves observations, which can be included at the expenses of the number of Tier 2 planets observed. Fig. 15 ARIEL MRS Tier 2 planets organised in size-bins. Different colours indicate the number of transits/eclipses needed to reach Tier 2 performances. Stripes indicate planets that will be studied with both transit and eclipse methods Fig. 16 ARIEL MRS Tier 2 planets organised in temperature-bins. Different colours indi- cate the number of transits/eclipses needed to reach Tier 2 performances. 01234567891011121314151617181920Radius (R⊕)1101001000Number of planets502 planets 60.2 %49 planets can be done with both methods1 transit<= 5 transits<= 10 transits<= 15 transits<= 20 transitsboth methods30050070090011001300150017001900210023002500Temperature (K)01020304050Number of planets1 transit<= 5 transits<= 10 transits<= 15 transits<= 20 transits The ARIEL Mission Reference Sample 17 Fig. 17 ARIEL MRS Tier 2 planets organised in size-bins. Different colours indicate dif- ferences in the simulated planetary densities. Fig. 18 ARIEL MRS Tier 2 planets organised in temperature-bins. Different colours indi- cate differences in the simulated stellar temperatures. 01234567891011121314151617181920Radius (R)1101001000Number of planetsg/cm30.01 - 0.020.02 - 0.030.03 - 0.040.04 - 0.070.07 - 0.120.12 - 0.190.19 - 0.320.32 - 0.520.52 - 0.850.85 - 1.391.39 - 2.282.28 - 3.733.73 - 6.116.11 - 10.0030050070090011001300150017001900210023002500Planetary Temperature (K)01020304050Number of planetsT*(K)3000 - 33003300 - 36003600 - 39003900 - 42004200 - 45004500 - 48004800 - 51005100 - 54005400 - 57005700 - 60006000 - 63006300 - 6600 18 4.3 MRS Tier 3: Benchmark Tiziano Zingales et al. In the current MRS, we have selected as Tier 3, 67 gaseous planets for weather studies. Fig. (19) shows the temperature distribution covered by the Tier 3 sample. Only 3% of the mission lifetime is required to achieve the Tier 3 science objectives for this sample. Fig. 19 Temperature distribution of the planets observable by ARIEL in the Benchmark. 30050070090011001300150017001900210023002500Temperature (K)0123456789Number of planets1 transit2 transits The ARIEL Mission Reference Sample 19 4.4 Compliance with TESS expected yields The Transiting Exoplanet Survey Satellite (TESS) is expected to provide a large fraction of the targets observable by ARIEL. The numbers of targets envisioned in the sample presented here are perfectly in line with the expected yield from The Transiting Exoplanet Survey Satellite (TESS), as shown in Fig 20 where we compare the expected TESS discoveries and the ARIEL MRS. We see that the ARIEL MRS is well within the TESS sample (Sullivan et al., 2015). The success of the TESS mission will allow the characterisation of hundreds of planets by ARIEL. Fig. 20 Comparison between the TESS targets (Sullivan et al., 2015) and the ARIEL MRS (green bars). 4.5 ARIEL MRS with currently known targets In February 2017 ∼210 transiting planets fulfill the ARIEL previous criteria. It means that, even if ARIEL were launched tomorrow, it would observe at least 210 relevant targets. Using the planets known today, we could organise the MRS into the following three tiers: -- Survey: 210 planets using 30% of the mission lifetime (Fig 21); -- Deep: 158 planets using 60% of the mission lifetime (Fig 26); -- Benchmark: 67 planets using 10% of the mission lifetime (Fig 27). In Fig 21, 22 and 23 we show the key physical parameters of the known planets defining the current observable MRS. In Fig 24 and 25 we show the properties of the stellar hosts. As mentioned previously, the number of known planets is expected to increase dramatically in the future. Pictorial representation (M. Ollivier, private comm.) of the known planets sky coordinates and their sky visibility all over the year is given in Fig 28. It shows that objects far away from the ecliptic plane will be visible longer than the planet close to this plane. EarthsSuper EarthsSub NeptunesNeptunes\GiantsType100101102103104105106Number of Planets1170671k4862363k111168817k67ARIEL targets2105 TESS Targets StarsFull-Frame Images 20 Tiziano Zingales et al. Fig. 21 ARIEL MRS with currently available planets radius distribution. Fig. 22 ARIEL MRS with currently available planets temperature distribution. Fig. 23 ARIEL MRS with currently available planets density distribution. Fig. 24 Temperature distribution of the stellar hosts for the planets shown in fig. 21 01234567891011121314151617181920212223Radius (R)051015202530Number of Planets210 planetsTRAPPIST-1 planets3004005006007008009001000110012001300140015001600170018001900200021002200230024002500Temperature (K)0510152025Number of PlanetsTRAPPIST-1 planets0.100.130.170.220.280.360.460.600.771.001.291.672.152.783.594.645.997.7410.00Density (g/cm3)051015202530Number of Planets30003600420048005400600066007200Temperature (K)010203040Number of Stars The ARIEL Mission Reference Sample 21 Fig. 25 Metallicity distribution of the stellar hosts for the planets shown in fig. 21 Fig. 26 Planets known today and observable by ARIEL in Deep mode, distributed in size-bins (top) and temperature bins (bottom) -- 158 planets. -0.50-0.45-0.40-0.35-0.30-0.25-0.20-0.15-0.10-0.05-0.000.050.100.150.200.250.300.350.400.450.50Metallicity0369121518Number of Stars01234567891011121314151617181920212223Radius (R)05101520Number of Planets158 planets30050070090011001300150017001900210023002500Temperature (K)0510152025Number of Planets 22 Tiziano Zingales et al. Fig. 27 Planets known today and observable by ARIEL in Benchmark mode, distributed size-bins (top) and temperature bins (bottom) -- 67 planets. 01234567891011121314151617181920212223Radius (R)02468101214Number of Planets67 planets30050070090011001300150017001900210023002500Temperature (K)01234567Number of Planets The ARIEL Mission Reference Sample 23 Fig. 28 A plot illustrating the fraction of the year for which a given location in the sky (in equatorial coordinates) is visible to ARIEL, as seen from a representative operational orbit of ARIEL at L2. Yellow dots: planets observed in Tier 1. Red dots: planets observed in Tier 2. Green dots: planets observed in Tier 3. (Marc Ollivier, private communication) 5 MRS optimisation for stellar hosts In this section we show another possible selection of the Tier 1 sample that maximises also the diversity of stellar hosts, additionally to other planet pa- rameters. In particular, the stellar metallicity is expected to play an important role in the planet formation process and type of chemistry of the planet (Venot et al., 2015). ARIEL will also collect important data to understand better the relationship between stellar metallicity and planetary characteristics. 5.1 Method We will limit our analysis to those systems which can be studied in up to six visits for each planet (either a transit or an occultation). We chose four physical quantities that define a 4D space to distribute the ARIEL targets. The quantities are: stellar effective temperature (Teff ), metallicity ([Fe/H]), planetary radius (Rpl) and planetary theoretical equi- librium temperature (Tpl). For the metallicity we use the values observed in the solar neighbourhood and reported by Casagrande et al. (2011). We adopt three bins for stellar Teff, [Fe/H] and planetary Rpl, while for the Tpl we use 24 Tiziano Zingales et al. Table 4 Bins of Teff, [Fe/H], Rpl, Tpl defining the 4D parameter space. Stellar Temp.: Teff Labels Metallicity: [Fe/H] Labels Planet Radius: Rpl Labels Planet Temp.: Tpl 3000 < T (K) < 4100 M-Late K [Fe/H] < -0.15 Low [Fe/H] Rpl < 3R⊕ Earths/ Super Earths contiguous bins: [250, 500, 800, 1200, 1600, 2600] K 4100 < T (K) < 5800 −0.15 <[Fe/H]< 0.15 Early K-G Solar 3 < R⊕ < 8 Neptunes T > 5800K F-G [Fe/H]> 0.15 High [Fe/H] Rpl > 8R⊕ Jupiters five bins, as detailed in Table 4. The three Teff bins correspond approximately to the ranges of spectral types M-Late / K stars, Early K-G stars and F-G stars, respectively, as indicated in the labels in Fig 30 to 32. Analogously, we separated the sample in low metallicity, solar metallicity and high metallicity, according to individual temperature values. The binning into 3 intervals of Teff, [Fe/H] and Rpl is a reasonable trade-off between a detailed representa- tion of the sample and a simple visualization of the richness and diversity of the physical configurations of the sample. We inferred from Casagrande et al. (2011) that the metallicities of stars in the solar neighbourhood are consistent with a normal distribution with mean -0.1 and standard deviation sd=0.2. Using such model distribution we simulated the values of [Fe/H] for each star in the ARIEL sample. The 4D space of Teff , [Fe/H], Rpl and Tpl is composed by a total of 3 × 3 × 3 × 5 = 135 cells. We assume that 10 systems are sufficiently reliable to determine the properties of the atmospheres of planets in each cell. 5.2 Results The population of 9545 planets is distributed in the 4-D bins as in Fig 29. From this distribution we selected 1002 exoplanets, requiring altogether 1538 satellite visits. These 1002 planets are distributed in the 4D space as shown in Fig. 30. The 3 × 3 panel grid distributes the sample along the 3 spectral types and the metallicity ranges reported in Table 4. Each panel is a matrix with planetary radii along x-axis and (calculated) equilibrium temperatures along y-axis, as specified in Table 4 and discussed above. The numbers in each box identify the numbers of systems with the corresponding Rpl, Tpl, spectral type, and [Fe/H] values. The 1002 systems in Fig. 30 tend to populate the cells corresponding to F- G-early and K stars orbited by Neptunes/Jupiters size planets (with a number of planets per cell N > 20), as these systems are the easiest to be observed with high signal to noise and, on average, with one or two visits. At the same time, planets around M or late K stars are much less represented in this distribution, especially planets smaller than Neptunes. This issue is addressed by prioritising these targets over the rest of the population. We found that planets around M stars require on average more visits than the analogues around early K, G, The ARIEL Mission Reference Sample 25 Fig. 29 Distribution of the 9545 planets in the 4D space of Teff, [Fe/H], Rpl, Tpl. Above each panels we indicate the spectral type and metallicity. The numbers in each cell are the numbers of planets with the corresponding properties. The colour scale indicates more populated cells (darker orange/brown). Grey cells without any number indicate no objects. and F stars. We managed to select 908 planets and, in particular, 594 of them require only 1 visit (65.4%), 151 planets require 2 visits (16.6%), 83 planets require 3 visits (9.1%), 41 planets require 4 visits (4.5%), and 39 planets require 5 visits (4.4%). The corrected sample is shown in Fig 31, where now ∼ 19% of the population are Earths/Super Earth or Neptunes around M or K stars observable with less than 6 visits. Assuming a total number of visits as in the 1002 planets configuration (approximately 1500 visits), we fixed the maximum number of systems (10 planets in our choice) in each 4D space cell. This choice implies that any additional targets in an "already full" cell will be discarded. In this way we can include planets in the empty or poorly populated parts of the parameter space. The goal is to verify that we can cover with enough statistics most of the 4D parameter space. The distribution of systems selected with such criteria is 0510152050010002000Radius (Rearth)Planet Temp. (K)M−Late K; Low [Fe/H]7421012198111410510152050010002000Radius (Rearth)Planet Temp. (K)M−Late K; Solar7481008281091170510152050010002000Radius (Rearth)Planet Temp. (K)M−Late K; High [Fe/H]314241919180510152050010002000Radius (Rearth)Planet Temp. (K)Early K−G; Low [Fe/H]75622639101220329532972012215420510152050010002000Radius (Rearth)Planet Temp. (K)Early K−G; Solar1254275441552724011839052220315540510152050010002000Radius (Rearth)Planet Temp. (K)Early K−G; High [Fe/H]2858134058102910028401140510152050010002000Radius (Rearth)Planet Temp. (K)F−G; Low [Fe/H]24214637427586139433291134399203010510152050010002000Radius (Rearth)Planet Temp. (K)F−G; Solar62925472107319113146412341625319403810510152050010002000Radius (Rearth)Planet Temp. (K)F−G; High [Fe/H]5638207026361038311133777 26 Tiziano Zingales et al. Fig. 30 Same as Fig. 29 1002 planets of the Mission Reference Sample. shown in Fig. 31. Compared to Fig. 30, we see that we can efficiently cover most of the 4D space in planetary sizes, planetary temperatures, host temperatures and metallicities, apart from those combination of parameters corresponding to not physical or rare systems (e.g., very hot planets around very cool stars). Our selection is composed by 908 unique planets requiring a total of 1504 visits. Among already known systems, 92 of the initial 211 systems are in this new list. This selection is not unique, and depends on our choices, but our exercise shows that we have great freedom on the final choice on how to spend ARIEL observing time, as it can be easily tuned on specific needs. Fig. 32 shows the average number of visits required to cover each cell of the 4D space. The number of visits needed for Jupiters and Neptunes is, typically, one or two, while Earths/Super Earths require from 3 to 5 visits each. To summarise, out of the 908 planets in our selection there are 594 planets requiring only 1 visit (65.4%), 151 planets requiring 2 visits (16.6%), 83 planets requiring 3 visits (9.1%), 41 planets requiring 4 visits (4.5%), and 39 planets requiring 5 visits (4.4%). 0510152050010002000Radius (Rearth)Planet Temp. (K)M−Late K; Low [Fe/H]1210510152050010002000Radius (Rearth)Planet Temp. (K)M−Late K; Solar112210510152050010002000Radius (Rearth)Planet Temp. (K)M−Late K; High [Fe/H]10510152050010002000Radius (Rearth)Planet Temp. (K)Early K−G; Low [Fe/H]11412398825121730510152050010002000Radius (Rearth)Planet Temp. (K)Early K−G; Solar18176197923122060510152050010002000Radius (Rearth)Planet Temp. (K)Early K−G; High [Fe/H]21333112510510152050010002000Radius (Rearth)Planet Temp. (K)F−G; Low [Fe/H]1116626559478145975460510152050010002000Radius (Rearth)Planet Temp. (K)F−G; Solar3212633044104819217335480510152050010002000Radius (Rearth)Planet Temp. (K)F−G; High [Fe/H]62261637524111 The ARIEL Mission Reference Sample 27 Fig. 31 Same as Fig. 30 for the selected sample of 908 known and simulated planetary systems. They have been selected by filling each cell with up to 10 objects and for a budget of total satellite visits of about 1500. As a final comment, we have verified that, by increasing the maximum number of systems per 4D cell while keeping fixed the total number of visits to ∼ 1500, we obtain that the number of observed planets increases (for ex- ample assuming N=15 as maximum systems per cell, we can observe up to 1000 systems), but at the same time the 4D cells of systems with cold/warm Earths/Super Earths would tend to be left empty and thus unexplored. This exercise shows the degree of flexibility offered by ARIEL in the choice of the target sample. 6 Conclusions In this paper we demonstrated that the current ARIEL design enables the observation of 900-1000 planets during its four-year lifetime, depending on the physical parameters of the planet/star systems which one wants to optimise. 0510152050010002000Radius (Rearth)Planet Temp. (K)M−Late K; Low [Fe/H]51010110101010510152050010002000Radius (Rearth)Planet Temp. (K)M−Late K; Solar51010210101100510152050010002000Radius (Rearth)Planet Temp. (K)M−Late K; High [Fe/H]39101410160510152050010002000Radius (Rearth)Planet Temp. (K)Early K−G; Low [Fe/H]610101010101010104101014100510152050010002000Radius (Rearth)Planet Temp. (K)Early K−G; Solar610101010101010105101014100510152050010002000Radius (Rearth)Planet Temp. (K)Early K−G; High [Fe/H]27101010107101028101100510152050010002000Radius (Rearth)Planet Temp. (K)F−G; Low [Fe/H]1010101010101010101010910100510152050010002000Radius (Rearth)Planet Temp. (K)F−G; Solar11010101010101010101010710100510152050010002000Radius (Rearth)Planet Temp. (K)F−G; High [Fe/H]51051010101010710103410 28 Tiziano Zingales et al. Fig. 32 Average number of visits required for the sample selected in Fig. 31. The binning is as in Figs. 30 to 31. The optimal sample of targets fulfils all the science objectives of the mission. While we currently know only ∼200 transiting exoplanets which could be part of the mission reference sample, new space missions and ground-based observatories are expected to discover thousands of new planets in the next decade. NASA-TESS alone is expected to deliver most ARIEL targets. Ackowledgements T. Z. is supported by the European Research CouncilERC projects ExoLights (617119) and from INAF trough the "Progetti PremialiâĂİ funding scheme of the Italian Ministry of Education, University, and Research. I.P and G.M. are supported by Ariel ASI-INAF agreement No. 2015-038-R.0. G.T. is supported by a Royal Society URF. We thank Enzo Pascale and Ludovig Puig for their help in setting up the ESA's Radiometric model. 0510152050010002000Radius (Rearth)Planet Temp. (K)M−Late K; Low [Fe/H]4.942.85.53.82.462.420510152050010002000Radius (Rearth)Planet Temp. (K)M−Late K; Solar54.22.84.83.82.711.40510152050010002000Radius (Rearth)Planet Temp. (K)M−Late K; High [Fe/H]44.42.514.82.332.50510152050010002000Radius (Rearth)Planet Temp. (K)Early K−G; Low [Fe/H]4.33.72.5431.94.33.324.93.51.813.82.20510152050010002000Radius (Rearth)Planet Temp. (K)Early K−G; Solar4.83.62.34.22.92.24.33.21.94.23.32141.80510152050010002000Radius (Rearth)Planet Temp. (K)Early K−G; High [Fe/H]34.22.54.231.9431.93.52.41.852.20510152050010002000Radius (Rearth)Planet Temp. (K)F−G; Low [Fe/H]3.61.94.22.51.73.61.91.43.93.11.53.73.11.40510152050010002000Radius (Rearth)Planet Temp. (K)F−G; Solar5.73.924.22.31.83.72.11.33.92.81.54.731.40510152050010002000Radius (Rearth)Planet Temp. (K)F−G; High [Fe/H]2.21.64.121.9321.64.23.11.82.73.61.3 The ARIEL Mission Reference Sample 29 Appendix A ESA Radiometric Model validation with ExoSim We compare the out-of-transit signal and noise from ESA Radiometric Model (ERM) with that from ExoSim. An early version of ARIEL with a grating design was used for the instrument model in each. We model 55 Cancri and GJ 1214 with the same PHOENIX spectra in each simulator and include only photon noise and the noise floor, Nmin(λ), which is dominated by dark current noise. All the calculations are done per unit time and per spectral bin (R = 30 in Ch1 and R = 100 in Ch0). The noise variance was compared assuming an aperture mask on the final images, and the noiseless signal per unit time was compared assuming no aperture. In the ERM, we use the following expression for Nmin giving the noise variance: Nmin(λ) = 2.44f λ2 mR∆pix 2 Idc (6) where Idc is the dark current per pixel, m is the reciprocal linear dispersion of the spectrum in µm wavelength per µm distance, R is the spectral resolving power and ∆pix is the pixel pitch. The noise variance from ExoSim is given as the average of 50 realizations with a standard deviation (shown as error bars in the following figures). For 55 Cancri e case (Fig 33), over all wavelength bins, the ERM signal is always within 2% of ExoSim, and the averaged noise variance within 5% of the ERM. In 94% of the bins, the ERM noise variance is within the standard deviation from ExoSim. Fig. 33 Comparison between the out-of-transit signal (left) and noise (right) simulated by ExoSim (white points) and the ESA Radiometric Model (blue points) for the star 55 Cancri. Subplots show the percent difference of the ERM from ExoSim. For GJ 1214 (Fig 34), the ERM signal is within 4% of ExoSim over all bins and the averaged noise variance within 6% of ExoSim over all bins. The ERM noise variance is always within the standard deviation from ExoSim over all bins. There is therefore good agreement between the two simulators. 30 Tiziano Zingales et al. Fig. 34 Comparison between the out-of-transit signal (left) and noise (right) simulated by ExoSim (white points) and the ESA Radiometric Model (blue points) for the star GJ 1214. Subplots show the percent difference of the ERM from ExoSim. The ARIEL Mission Reference Sample 31 Appendix B Known planets observable by ARIEL planetary properties P (days) # 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 Planet 55 Cnc e EPIC 204129699 b WASP-52 b HD 189733 A b WASP-77 A b WASP-85 A b WASP-33 b WASP-19 b WASP-95 b WASP-121 b WASP-12 b WASP-35 b HAT-P-30 b WASP-108 b HD 209458 b WASP-122 b WASP-2 A b HAT-P-32 b WASP-43 b WASP-123 b WASP-101 b WASP-74 b WASP-76 b WASP-1 b KELT-10 b KELT-3 b WASP-62 b HD 149026 b WASP-97 b WASP-94 A b HAT-P-8 b WASP-54 b WASP-109 b HAT-P-41 b HAT-P-13 b KELT-15 b KELT-7 b HAT-P-6 b WASP-49 b WASP-15 b WASP-79 b KELT-4A b WASP-17 b WASP-3 b WASP-7 b KELT-8 b HAT-P-22 b WASP-13 b HAT-P-33 b TrES-4 A b R (R⊕) M (M⊕) 1.88 15.47 13.94 12.49 13.28 16.24 15.78 15.21 13.28 19.83 19.05 14.48 14.70 14.09 15.14 21.64 12.26 22.35 11.37 14.56 15.47 17.12 20.08 16.27 15.35 14.90 15.25 7.88 12.40 18.87 14.50 18.14 15.83 18.49 14.05 15.83 16.82 14.59 12.24 15.67 18.65 18.64 21.85 15.96 14.59 20.41 11.85 15.44 20.05 20.17 8.07 563.97 146.24 361.78 559.52 387.85 1459.19 371.32 359.23 376.08 446.34 228.89 226.03 283.57 226.99 436.17 290.57 299.15 646.62 292.47 158.95 302.01 292.47 271.49 215.86 464.78 181.21 113.17 419.64 143.69 405.33 202.19 289.30 254.33 270.22 289.30 406.92 336.03 120.17 172.31 286.12 286.75 154.50 654.89 305.19 275.63 682.55 158.95 242.56 157.05 T (K) 1891 1473 1267 1180 1762 1341 2541 1998 1521 2295 2399 1414 1594 1558 1401 1900 1276 1850 1403 1477 1518 1872 2125 1777 1340 1774 1389 1699 1500 1464 1687 1531 1729 1886 1600 1500 1996 1629 1334 1609 1709 1779 1725 1933 1448 1633 1248 1494 1799 1644 stellar properties T (K) R (R(cid:12)) 5196 0.95 5280 0.91 5000 0.87 4980 0.80 1.00 5500 5685 1.04 7200 1.50 5500 0.97 5630 1.11 6459 1.35 6118 1.35 1.07 5990 6304 1.24 6000 1.17 6075 1.15 5720 1.40 5255 0.89 1.18 6207 4520 0.72 5740 1.21 6400 1.34 5990 1.48 6250 1.46 6160 1.24 1.11 5948 6304 1.28 6230 1.25 6147 1.30 5640 1.12 6170 1.29 1.19 6200 6100 1.15 6520 0.91 6390 1.42 5638 1.22 6003 1.18 6789 1.53 1.29 6570 5600 0.94 6300 1.18 6600 1.56 6206 1.20 6650 1.31 1.24 6400 6400 1.28 5754 1.21 5302 0.92 5989 1.19 6446 1.40 1.45 6295 occultation occultation occultation transit transit transit transit transit transit transit transit transit transit type transit transit transit transit transit transit transit transit transit transit transit transit transit Observation # 1 1 1 1 1 1 1 1 1 1 1 1 1 2 1 1 2 1 1 1 1 1 1 1 1 1 1 1 2 1 1 1 1 1 1 2 1 2 1 1 1 1 1 1 1 1 3 1 1 1 transit transit transit transit transit transit transit transit transit transit transit transit transit transit transit transit transit transit occultation transit occultation occultation occultation occultation 0.74 1.26 1.75 2.22 1.36 2.66 1.22 0.79 2.18 1.27 1.09 3.16 2.81 2.68 3.52 1.71 2.15 2.15 0.81 2.98 3.59 2.14 1.81 2.52 4.17 2.70 4.41 2.88 2.07 3.95 3.08 3.69 3.32 2.69 2.92 3.33 2.73 3.85 2.78 3.75 3.66 2.99 3.74 1.85 4.95 3.24 3.21 4.35 3.47 3.55 32 # 51 52 53 54 55 56 57 58 59 60 61 62 63 64 65 66 67 68 69 70 71 72 73 74 75 76 77 78 79 80 81 82 83 84 85 86 87 88 89 90 91 92 93 94 95 96 97 98 99 100 PTFO 8-8695 b Planet WASP-82 b WASP-31 b HAT-P-45 b KELT-2A b WASP-26 b TrES-2 WASP-50 b WASP-63 b XO-2N b WASP-104 b WASP-41 b HAT-P-40 b WASP-48 b HAT-P-4 b WASP-4 b WASP-103 b WASP-75 b Qatar-1 b WASP-20 b TrES-3 b HAT-P-1 b WASP-90 b HAT-P-46 b WASP-111 b XO-1 b WASP-34 b WASP-88 b HATS-3 b WASP-100 b WASP-68 b CoRoT-2 b HAT-P-49 b HAT-P-56 b HAT-P-7 b WASP-21 b WASP-22 b WASP-24 b WASP-25 b HAT-P-5 b WASP-69 b WASP-87 b HAT-P-24 b HAT-P-39 b WASP-16 b TrES-1 b WASP-64 b WASP-6 b WASP-55 b HAT-P-36 b planetary properties P (days) R (R⊕) M (M⊕) 394.20 18.33 151.96 16.87 283.57 15.65 14.33 472.41 326.81 14.06 398.34 12.83 467.32 12.62 120.80 15.69 189.79 10.68 404.38 12.48 13.28 292.47 195.51 18.98 311.55 18.33 216.18 13.94 356.53 14.96 467.32 17.59 13.94 340.16 422.82 12.77 99.50 16.00 607.20 14.32 953.72 20.96 166.58 13.35 17.89 200.28 156.73 14.09 581.77 15.82 286.12 12.99 187.57 13.39 178.03 18.65 361.78 18.62 18.54 645.35 302.01 13.61 1052.27 16.08 549.98 15.51 693.04 16.09 543.30 16.01 12.75 87.74 186.93 12.71 328.08 12.11 184.39 13.83 336.98 13.74 82.66 11.60 693.04 15.20 13.63 217.77 190.43 17.24 271.81 11.06 241.93 12.06 404.06 13.95 159.91 13.43 14.27 181.21 582.41 13.87 2.71 3.41 3.13 4.11 2.76 2.47 1.96 4.38 2.62 1.76 3.05 4.46 2.14 3.06 1.34 0.93 2.48 1.42 4.90 1.31 0.45 4.47 3.92 4.47 2.31 3.94 4.32 4.95 3.55 2.85 5.08 1.74 2.69 2.79 2.20 4.32 3.53 2.34 3.76 2.79 3.87 1.68 3.36 3.54 3.12 3.03 1.57 3.36 4.47 1.33 Tiziano Zingales et al. T (K) 2127 1502 1605 1671 1618 1458 1354 1496 1312 1476 1278 1719 1980 1653 1818 2430 1660 1347 1345 1654 1884 1259 1791 1413 2065 1216 1131 1716 1757 2143 1447 1484 2072 1791 2141 1298 1383 1611 1209 1477 938 2251 1581 1705 1235 1147 1587 1161 1236 1778 stellar properties T (K) R (R(cid:12)) 6490 1.63 6200 1.16 6330 1.26 1.31 6148 5950 1.12 5850 0.98 5400 0.89 5570 1.32 5332 0.97 5475 1.08 0.95 5450 6080 1.51 5920 1.19 5890 1.26 5500 0.93 6110 1.20 1.14 6100 4910 0.85 5950 1.20 5720 0.88 3470 0.34 5975 1.13 1.55 6430 6120 1.28 6400 1.50 5940 1.00 5700 1.01 6431 1.45 6351 1.30 1.57 6900 5911 1.24 5575 0.97 6820 1.54 6566 1.30 6310 1.59 0.89 5800 6000 1.10 6075 1.13 5750 1.00 5960 1.16 4715 0.83 6450 1.20 1.19 6329 6430 1.40 5550 1.02 5250 0.88 5400 0.98 5450 0.89 1.01 5900 5580 1.02 occultation occultation occultation transit occultation occultation occultation occultation transit occultation occultation transit transit transit transit transit transit transit transit transit type transit transit transit transit Observation # 1 1 2 1 2 4 5 1 2 3 2 1 1 2 1 1 2 5 1 1 1 1 2 1 1 1 1 1 2 1 2 1 1 1 1 1 2 4 1 6 1 1 4 3 4 2 3 1 1 1 transit transit transit transit transit transit transit transit transit transit transit transit occultation occultation transit occultation occultation occultation occultation transit transit occultation occultation occultation occultation occultation The ARIEL Mission Reference Sample 33 planetary properties P (days) Planet XO-4 b # HAT-P-9 b 101 102 HAT-P-14 b 103 WASP-28 b 104 105 WASP-58 b HAT-P-23 b 106 107 Qatar-2 b 108 WASP-5 b 109 WASP-65 b 110 CoRoT-1 b HAT-P-27 b 111 112 KELT-6 b 113 WASP-45 b 114 WASP-72 b 115 HATS-1 b 116 WASP-78 b 117 WASP-96 b 118 HAT-P-28 b 119 WASP-39 b 120 WASP-80 b 121 HATS-2 b 122 WASP-71 b 123 WASP-38 b 124 WASP-110 b 125 HAT-P-3 b 126 WASP-47 b 127 WASP-98 b 128 WASP-46 b 129 HAT-P-25 b 130 WASP-18 b 131 WASP-67 b 132 WASP-14 b 133 WASP-60 b 134 WASP-11 b 135 HAT-P-35 b 136 WASP-36 b 137 HAT-P-50 b 138 WASP-99 b 139 HAT-P-42 b 140 WASP-73 b 141 WASP-135 b 142 WASP-23 b 143 TrES-5 b HAT-P-16 b 144 Kepler-12 b 145 146 Kepler-7 b 147 WASP-44 b 148 149 150 HAT-P-43 b HAT-P-55 b XO-5 b R (R⊕) M (M⊕) 213.00 15.36 699.40 13.17 288.34 13.31 14.70 546.80 282.94 15.03 664.43 15.01 790.63 12.55 520.41 12.85 492.76 12.20 16.35 327.44 197.10 11.19 140.51 12.95 320.13 12.73 448.25 11.08 589.72 14.29 368.77 19.20 13.17 152.60 199.01 13.30 89.01 13.94 176.12 10.45 427.58 12.82 712.75 16.02 11.96 861.53 162.13 13.58 187.88 9.07 336.98 12.84 263.86 12.07 667.92 14.38 13.06 180.22 3315.77 12.78 133.52 15.36 2333.75 14.06 163.40 9.44 146.24 11.47 14.62 335.07 724.51 13.93 429.17 14.13 883.78 12.07 309.96 14.01 597.66 12.73 604.02 14.27 10.56 281.03 565.24 13.27 1332.98 13.06 136.70 19.20 137.65 17.71 276.26 11.00 11.30 342.39 209.82 14.08 12.97 185.02 T (K) 1490 1525 1429 1418 1242 1997 1256 1693 1446 1839 1161 1284 1165 1819 1332 2136 1251 1345 1088 794 1528 1987 1218 1113 1115 1240 1149 1615 1172 2345 1000 1834 1261 1002 1537 1655 1805 1438 1389 1736 1673 1099 1433 1527 1341 1584 1275 1206 1322 1278 stellar properties T (K) R (R(cid:12)) 6350 1.28 6600 1.39 6150 1.02 1.32 5700 5800 0.94 5905 1.13 4645 0.74 5700 1.00 5600 0.93 0.95 6298 5300 0.92 6272 1.13 5140 0.91 6250 1.23 5870 0.99 6100 2.02 1.06 5540 5680 1.02 5400 0.93 4145 0.57 5227 0.88 6050 1.56 1.23 6150 5400 0.89 5224 0.92 5576 1.04 5525 0.69 5620 0.96 1.01 5500 6400 1.24 5200 0.87 6462 1.21 5900 0.51 4974 0.82 1.24 6096 5881 1.02 6280 1.27 6180 1.48 5743 1.18 6036 1.34 5675 0.98 0.78 5150 5171 0.88 6140 1.22 5953 1.09 5933 1.36 5410 0.92 0.88 5510 5645 1.05 1.01 5808 occultation transit occultation occultation transit occultation transit occultation occultation occultation occultation occultation type transit transit transit transit transit transit transit transit Observation # 4 3 6 6 2 1 9 2 7 1 3 1 6 2 14 1 3 6 1 1 5 1 10 1 4 8 8 3 3 1 1 1 6 1 12 3 3 6 13 3 3 10 10 3 2 4 28 16 12 10 transit transit transit transit transit transit transit transit transit transit transit transit transit transit transit transit transit occultation occultation occultation occultation occultation occultation occultation occultation occultation occultation occultation occultation occultation 3.92 4.63 3.41 4.13 5.02 1.21 1.34 1.63 2.31 1.51 3.04 7.85 3.13 2.22 3.45 2.18 3.43 3.26 4.06 3.07 1.35 2.90 6.87 3.78 2.90 4.16 2.96 1.43 3.65 0.94 4.61 2.24 4.31 3.72 3.65 1.54 3.12 5.75 4.64 4.09 1.40 2.94 1.48 2.78 4.44 4.89 2.42 4.19 3.33 3.58 34 Tiziano Zingales et al. Planet occultation occultation occultation occultation occultation planetary properties P (days) HAT-P-52 b HAT-P-20 b WASP-120 b HATS-9 b CoRoT-19 b WASP-32 b HAT-P-29 b WASP-10 b Kepler-6 b HD219134b HATS-13 b HAT-P-51 b HAT-P-34 b WASP-37 b WASP-56 b WASP-66 b WASP-112 b HAT-P-44 b HAT-P-37 b Gliese 436 b WASP-29 b HD 219134 b HAT-P-12 b Kepler-13 A b HAT-P-19 b CoRoT-11 b Kepler-8 b HATS-10 b WTS-2 b # 151 152 153 154 155 156 157 158 159 160 161 162 163 164 165 166 167 168 169 170 171 172 173 174 175 176 177 178 179 180 181 182 183 184 185 186 187 188 189 190 191 192 193 194 195 196 197 198 199 200 201 202 203 204 205 206 207 208 209 210 occultation Table 5 List of known planets observable by ARIEL. The former to last column represents the number of transits/eclipses necessary to fulfil the ARIEL Tier 1 goals. R (R⊕) M (M⊕) 1144.46 13.06 247.33 12.15 972.79 11.85 213.00 14.27 1.61 4.47 172.62 13.30 98.23 14.19 1057.99 13.14 539.17 12.47 181.52 11.98 15.25 737.54 279.76 13.07 124.62 14.05 371.63 12.93 23.11 4.13 77.57 8.69 1.57 3.81 67.08 10.52 2571.87 15.43 92.83 12.17 791.59 15.25 187.57 15.58 167.22 10.63 14.96 356.06 260.05 11.07 2303.55 9.51 1592.71 16.62 266.09 11.69 352.88 15.91 216.18 18.87 11.85 158.95 654.89 13.61 690.18 11.74 471.77 14.46 713.38 11.39 420.59 11.19 18.11 435.53 638.99 14.92 213.63 10.05 148.46 15.23 1014.44 12.02 298.51 14.54 3748.12 13.35 11.20 87.58 2.77 6.20 13.73 3.80 25.75 4.96 1.62 1.16 0.66 0.28 18.76 6.20 62.31 10.39 2.34 7.55 169.76 11.08 222.53 10.70 101.41 10.95 27.02 7.93 4.13 1.76 75.34 10.01 6.95 2.10 10.44 stellar properties T (K) R (R(cid:12)) 6100 1.10 6087 1.21 4675 0.71 5640 1.05 0.78 4699 5523 0.96 5449 0.98 6442 1.39 5800 0.85 5600 1.03 1.30 6600 5610 0.81 5295 0.94 5500 0.93 3684 0.45 4800 0.82 4699 0.79 0.73 4650 7200 1.72 4990 0.84 6440 1.27 6251 1.13 5880 1.10 0.82 5000 5131 0.89 4595 0.76 6450 1.45 5366 1.03 6090 1.21 6075 1.28 0.88 5200 6250 1.22 6065 1.22 5956 1.09 5600 1.03 5403 1.00 0.76 5493 6409 1.20 5600 1.01 6100 1.00 6079 1.27 5750 1.17 6429 1.21 1.13 6040 3250 0.18 3652 0.51 4780 0.81 3270 0.18 3684 0.45 0.82 5079 4870 0.77 5119 0.77 5246 0.86 5280 0.85 3724 0.57 5743 1.12 1.02 5513 5304 0.94 6018 1.04 1.36 6290 T (K) 1507 1224 1009 1354 934 1212 1159 1440 1293 1117 1754 1349 1092 1166 695 970 931 932 2389 982 1686 1528 1369 1495 1184 946 1842 1769 1616 1554 969 1509 1343 1624 906 1282 908 2074 1430 1315 816 1780 1665 997 552 635 848 529 813 967 818 729 758 780 693 596 764 998 496 1443 2.72 5.72 3.09 3.23 3.09 3.04 4.22 5.45 3.58 4.62 4.09 3.04 4.30 2.80 2.64 3.92 3.09 3.21 1.76 4.01 2.99 3.52 3.31 1.02 2.75 2.88 3.61 1.92 3.90 3.10 4.98 3.86 5.01 1.96 8.16 2.52 7.79 1.54 2.84 4.04 21.22 1.72 3.19 10.02 1.58 3.34 4.89 1.63 1.37 4.23 5.51 9.49 10.34 8.52 3.33 42.36 10.95 4.76 45.15 5.63 WASP-117 b Gliese 1214 b Gliese 3470 b HAT-P-11 b GJ 1132 b Gliese 436 c HAT-P-26 b HAT-P-18 b HD 97658 b HAT-P-17 b WASP-84 b HATS-6 b WASP-42 b WASP-61 b HAT-P-31 b HAT-P-53 b WASP-8 b HATS-4 b Kepler-447 b Kepler-76 b WASP-57 b CoRoT-5 b HD 17156 b Kepler-412 b occultation occultation occultation occultation occultation KOI-142 b HATS-5 b Kepler-51 b HAT-P-2 b occultation occultation occultation occultation occultation type transit transit transit transit transit transit transit transit transit transit transit transit transit transit transit transit transit transit transit Observation # 5 8 37 15 1 10 2 10 43 4 3 28 2 37 1 1 1 1 1 1 5 20 28 16 43 97 1 9 16 16 1 17 43 7 6 97 2 2 43 15 9 12 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 transit transit transit transit transit transit transit transit transit transit transit transit transit transit transit transit transit transit transit transit transit EPIC 203771098 c occultation occultation occultation occultation OGLE-TR-10 b XO-3 b 2778.51 The ARIEL Mission Reference Sample 35 References Baraffe, I., Chabrier, G., Allard, F., and Hauschildt, P. H. (1998). Evolutionary models for solar metallicity low-mass stars: mass-magnitude relationships and color-magnitude diagrams. A&A, 337:403 -- 412. Casagrande, L., Schönrich, R., Asplund, M., Cassisi, S., Ramírez, I., Melén- dez, J., Bensby, T., and Feltzing, S. (2011). New constraints on the chemical evolution of the solar neighbourhood and Galactic disc(s). Improved astro- physical parameters for the Geneva-Copenhagen Survey. A&A, 530:A138. Chen, J. and Kipping, D. M. (2016). Probabilistic Forecasting of the Masses and Radii of Other Worlds. ArXiv e-prints. Cohen, M., Wheaton, W. A., and Megeath, S. T. (2003). Spectral Irradiance Calibration in the Infrared. XIV. The Absolute Calibration of 2MASS. AJ, 126:1090 -- 1096. Fressin, F., Torres, G., Charbonneau, D., Bryson, S. T., Christiansen, J., Dress- ing, C. D., Jenkins, J. M., Walkowicz, L. M., and Batalha, N. M. (2013). The False Positive Rate of Kepler and the Occurrence of Planets. ApJ, 766:81. Mulders, G. D., Pascucci, I., and Apai, D. (2015). An Increase in the Mass of Planetary Systems around Lower-mass Stars. ApJ, 814:130. Mulders, G. D., Pascucci, I., Apai, D., Frasca, A., and Molenda-Zakowicz, J. (2016). A Super-Solar Metallicity For Stars With Hot Rocky Exoplanets. ArXiv e-prints. Öberg, K. I., Boogert, A. C. A., Pontoppidan, K. M., van den Broek, S., van Dishoeck, E. F., Bottinelli, S., Blake, G. A., and Evans, II, N. J. (2011). The Spitzer Ice Legacy: Ice Evolution from Cores to Protostars. ApJ, 740:109. Pascale, E., Waldmann, I. P., MacTavish, C. J., Papageorgiou, A., Amaral- Rogers, A., Varley, R., Coudé du Foresto, V., Griffin, M. J., Ollivier, M., Sarkar, S., Spencer, L., Swinyard, B. M., Tessenyi, M., and Tinetti, G. (2015). EChOSim: The Exoplanet Characterisation Observatory software simulator. Experimental Astronomy, 40:601 -- 619. Piskorz, D., Knutson, H. A., Ngo, H., Muirhead, P. S., Batygin, K., Crepp, J. R., Hinkley, S., and Morton, T. D. (2015). Friends of Hot Jupiters. III. An Infrared Spectroscopic Search for Low-mass Stellar Companions. ApJ, 814:148. Puig, L., Isaak, K., Linder, M., Escudero, I., Crouzet, P.-E., Walker, R., Ehle, M., Hübner, J., Timm, R., de Vogeleer, B., Drossart, P., Hartogh, P., Lovis, C., Micela, G., Ollivier, M., Ribas, I., Snellen, I., Swinyard, B., Tinetti, G., and Eccleston, P. (2015). The phase 0/A study of the ESA M3 mission candidate EChO. Experimental Astronomy, 40:393 -- 425. Ribas, I. and Lovis, C. (2013). Echo targets: the mission reference sample and beyond [echo-sre-sa-phasea-001_mrsv2.4], european space agency. Sarkar, S., Papageorgiou, A., and Pascale, E. (2016). Exploring the poten- tial of the ExoSim simulator for transit spectroscopy noise estimation. In Space Telescopes and Instrumentation 2016: Optical, Infrared, and Millime- ter Wave, volume 9904, page 99043R. Sarkar, S. and Pascale, E. (2015). ExoSim: a novel simulator of exoplanet 36 Tiziano Zingales et al. spectroscopic observations. European Planetary Science Congress 2015, held 27 September - 2 October, 2015 in Nantes, France, Online at <A href="http://meetingorganizer.copernicus.org/EPSC2015/EPSC2015"> http://meetingorganizer.copernicus.org/EPSC2015</A>, 187, 10:EPSC2015 -- 187. id.EPSC2015- Sullivan, P. W., Winn, J. N., Berta-Thompson, Z. K., Charbonneau, D., Dem- ing, D., Dressing, C. D., Latham, D. W., Levine, A. M., McCullough, P. R., Morton, T., Ricker, G. R., Vanderspek, R., and Woods, D. (2015). The Transiting Exoplanet Survey Satellite: Simulations of Planet Detections and Astrophysical False Positives. ApJ, 809:77. Triaud, A. H. M. J., Lanotte, A. A., Smalley, B., and Gillon, M. (2014). Colour-magnitude diagrams of transiting Exoplanets - II. A larger sample from photometric distances. MNRAS, 444:711 -- 728. Valencia, D., Sasselov, D. D., and O'Connell, R. J. (2007). Detailed Models of Super-Earths: How Well Can We Infer Bulk Properties? ApJ, 665:1413 -- 1420. Venot, O., Hébrard, E., Agúndez, M., Decin, L., and Bounaceur, R. (2015). New chemical scheme for studying carbon-rich exoplanet atmospheres. A&A, 577:A33.
1605.04145
1
1605
2016-05-13T12:05:31
Near-infrared spatially resolved spectroscopy of (136108) Haumea's multiple system
[ "astro-ph.EP" ]
The transneptunian region of the solar system is populated by a wide variety of icy bodies showing great diversity. The dwarf planet (136108) Haumea is among the largest TNOs and displays a highly elongated shape and hosts two moons, covered with crystalline water ice like Hamuea. Haumea is also the largest member of the sole TNO family known to date. A catastrophic collision is likely responsible for its unique characteristics. We report here on the analysis of a new set of observations of Haumea obtained with SINFONI at the ESO VLT. Combined with previous data, and using light-curve measurements in the optical and far infrared, we carry out a rotationally resolved spectroscopic study of the surface of Haumea. We describe the physical characteristics of the crystalline water ice present on the surface of Haumea for both regions, in and out of the Dark Red Spot (DRS), and analyze the differences obtained for each individual spectrum. The presence of crystalline water ice is confirmed over more than half of the surface of Haumea. Our measurements of the average spectral slope confirm the redder characteristic of the spot region. Detailed analysis of the crystalline water-ice absorption bands do not show significant differences between the DRS and the remaining part of the surface. We also present the results of applying Hapke modeling to our data set. The best spectral fit is obtained with a mixture of crystalline water ice (grain sizes smaller than 60 micron) with a few percent of amorphous carbon. Improvements to the fit are obtained by adding ~10% of amorphous water ice. Additionally, we used the IFU-reconstructed images to measure the relative astrometric position of the largest satellite Hi`iaka and determine its orbital elements. An orbital solution was computed with our genetic-based algorithm GENOID and our results are in full agreement with recent results.
astro-ph.EP
astro-ph
Astronomy & Astrophysics manuscript no. Gourgeot_Haumea October 4, 2018 c(cid:13)ESO 2018 Near-infrared spatially resolved spectroscopy of (136108) Haumea's multiple system(cid:63) F. Gourgeot1, 2, B. Carry3, 4, C. Dumas1, F. Vachier3, F. Merlin5, 6, P. Lacerda7, M. A. Barucci4, and J. Berthier3 1 European Southern Observatory, Alonso de Córdova 3107, Vitacura, Santiago, Chile. 2 Observatório Nacional, COAA, Rua General José Cristino 77, 20921-400, Rio de Janeiro, Brazil. e-mail: 6 1 0 2 y a M 3 1 . ] P E h p - o r t s a [ 1 v 5 4 1 4 0 . 5 0 6 1 : v i X r a [email protected] 3 IMCCE, Observatoire de Paris, PSL Research University, CNRS, Sorbonne Universités, UPMC Univ Paris 06, Univ. Lille, France 4 Laboratoire Lagrange, Université de Nice-Sophia Antipolis, CNRS, Observatoire de la Côte d'Azur, France 5 Observatoire de Paris-Meudon / LESIA, Meudon, France. 6 Université Denis Diderot, Paris VII, France. 7 Max Planck Institute for Solar System Research, Justus-von-Liebig-Weg 3, 37077 Göttingen, Germany October 4, 2018 ABSTRACT Context. The transneptunian region of the solar system is populated by a wide variety of icy bodies showing great diversity in orbital behavior, size, surface color, and composition. Aims. The dwarf planet (136108) Haumea is among the largest transneptunian objects and is a very fast rotator (∼3.9h). This dwarf planet displays a highly elongated shape and hosts two small moons that are covered with crystalline water ice, similar to their central body. A particular region of interest is the Dark Red Spot (DRS) identified on the surface of Haumea from multiband light-curve analysis (Lacerda et al. 2008). Haumea is also known to be the largest member of the sole TransNeptunian Objects (TNO) family known to date, and an outcome of a catastrophic collision that is likely responsible for the unique characteristics of Haumea. Methods. We report here on the analysis of a new set of near-infrared Laser Guide Star assisted observations of Haumea obtained with the Integral Field Unit (IFU) Spectrograph for INtegral Field Observations in the Near Infrared (SINFONI) at the European Southern Observatory (ESO) Very Large Telescope (VLT) Observatory. Combined with previous data published by Dumas et al. (2011), and using light-curve measurements in the optical and far infrared to associate each spectrum with its corresponding rotational phase, we were able to carry out a rotationally resolved spectroscopic study of the surface of Haumea. Results. We describe the physical characteristics of the crystalline water ice present on the surface of Haumea for both regions, in and out of the DRS, and analyze the differences obtained for each individual spectrum. The presence of crystalline water ice is confirmed over more than half of the surface of Haumea. Our measurements of the average spectral slope (1.45 ± 0.82 % by 100 nm) confirm the redder characteristic of the spot region. Detailed analysis of the crystalline water-ice absorption bands do not show significant differences between the DRS and the remaining part of the surface. We also present the results of applying Hapke modeling to our data set. The best spectral fit is obtained with a mixture of crystalline water ice (grain sizes smaller than 60 µm) with a few percent of amorphous carbon. Improvements to the fit are obtained by adding ∼10% of amorphous water ice. Additionally, we used the IFU- reconstructed images to measure the relative astrometric position of the largest satellite Hi'iaka and determine its orbital elements. An orbital solution was computed with our genetic-based algorithm GENOID and our results are in full agreement with recent results. Key words. Transneptunian object - 2003 EL61 (136108) Haumea - Surface composition high angular resolution - Infrared imaging spectroscopy - Dynamics - Hi'iaka - Namaka 1. Introduction The object (136108) Haumea is one of the most remarkable transneptunian objects (TNO). Its visible light curve indicates a very short (∼3.91 h) rotation period and an ellipsoidal shape (1000 x 800 x 500 km; Rabinowitz et al. 2006), while its gravity is determined to be sufficient for having relaxed into hydrostatic equilibrium, making Haumea a dwarf planet. Dynamically, it is considered a classical TNO, with an orbital period of 283 years, a perihelion at 35 AU, and an orbital inclination of 28◦ (Rabi- nowitz et al. 2006). Haumea is one of the bluest TNOs (Tegler et al. 2007) known to date. Its surface is covered by almost pure water ice (Trujillo et al. 2007; Merlin et al. 2007), though its (cid:63) Based on observations collected at the European Organisation for Astronomical Research in the Southern Hemisphere, Chile, Program ID: 60.A-9235. high density (∼2.6-3.3 g.cm−3, Rabinowitz et al. 2006; Lacerda & Jewitt 2007; Carry et al. 2012) indicates a more rocky interior. It possesses two satellites (Brown et al. 2005, 2006), the largest of the two is also coated with crystalline water ice (Barkume et al. 2006; Dumas et al. 2011). Such dynamical and physical characteristics, added to the fact that Haumea is the largest body of a population of small water-ice TNOs, all sharing the same orbital parameters (Schaller & Brown 2008), point to the idea that Haumea is the remnant body of an ancient (>1 Gyr) catas- trophic collision (Ragozzine & Brown 2007; Snodgrass et al. 2010; Carry et al. 2012). High time-resolution observations, combined with multi- color photometry, provide evidence for a localized surface fea- ture that is redder and darker than the surrounding material (Lac- erda et al. 2008; Lacerda 2009). The presence of such a fea- ture (perhaps of collisional origin) makes Haumea the second Article number, page 1 of 10 A&A proofs: manuscript no. Gourgeot_Haumea TNO (after Pluto) displaying strong surface heterogeneities. In this study, we combine two sets of observations taken four years apart to analyze the spectral behavior of the Dark Red Spot (DRS) and its surrounding areas and the ice properties of the surface material. Finally, we measure the astrometric positions of Hi'iaka directly from our set of hyperspectral images, and determine the Keplerian elements of its orbit using Genoid, a genetic-based algorithm (Vachier et al. 2012). 2. Observations Haumea was observed in H and K bands on 2007 March 15 and 2011 April 09 UT, using the Laser Guide-Star Facility (LGSF) and the Spectrograph for INtegral Field Observations in the Near Infrared (SINFONI) instrument installed at the 8-m "Yepun" telescope of the European Southern Observatory (ESO) Very Large Telescope (VLT). SINFONI is an integral-field spectrome- ter working in the 1.0 -- 2.5 µm range. It is equipped with an adap- tive optics (AO) system offering Natural Guide Star (NGS) and Laser Guide Star (LGS) channels. The use of this instrument for observing the large TNOs Haumea, Eris, and Orcus has been de- scribed in earlier papers (Merlin et al. 2007; Barucci et al. 2008; Dumas et al. 2007, 2011; Carry et al. 2012), and more informa- tion about SINFONI can be found in Eisenhauer et al. (2003) and Bonnet et al. (2004). The observations for both epochs were car- ried out using the AO system and its LGS facility. The laser pro- duces an artificial visible-light star of R mag∼ 13.4 near the line of sight of Haumea (V-mag = 17.4), thus providing a gain of four magnitudes for characterizing the high orders of the wavefront (in comparison to non-laser observations). Haumea was used as the on-axis tip-tilt reference source, thus delivering optimal cor- rection by the AO-LGS system. The atmospheric conditions were extremely good during the 2007 observations with a seeing varying between 0.81" and 1.13". On March 15, between 6h34 UT and 7h24 UT, six ex- posures of 300 s each were obtained on Haumea (total integra- tion time of 0.5 h), interspersed with three exposures of 300s to record the sky background. The results of these data were already published (Dumas et al. 2011), but we decided to re- duce them again using the same data processing and the latest data reduction tools for both epochs. On 2011 April 09, be- tween 3h57 UT and 5h40 UT, six exposures of 600 s each were obtained on Haumea (total integration time of 1.15 h), inter- spaced by sky background measurements. However, as a result of relatively poor meteorological conditions (degraded sky trans- parency due to clouds and wind above 15 m/s after 6h00 UT), it was not possible to obtain the full mapping of Haumea during the 2011 observations (see Fig. 1). Still, combining the 2007 (sam- pling almost exactly the DRS longitudinal span) and the 2011 data allowed us to obtain a substantial longitudinal coverage of Haumea and explore any variations of its surface properties. The H+K spectral grating (resolving power of ∼1500), covering both H and K bands simultaneously, as well as a plate scale of 100 mas/pixel (3(cid:48)(cid:48) × 3(cid:48)(cid:48) field of view), were used on each observation date. Using the same settings (though in NGS), we also observed standard solar-analog stars that are close in time to observations of Haumea and at similar airmass. These stars were used to cor- rect our spectra from the solar response and telluric absorption features (see Table 1). Article number, page 2 of 10 Fig. 1. Positions of all observed epochs (six for 2007 and seven for 2011, see Table 1 for notations) onto Haumea's light curve. The location of the Dark Red Spot (Lacerda 2009) is represented by the red-shaded region. The combination of the data obtained on both dates returns the 230 degrees of rotational coverage of Haumea. The 2007 observations published in (Dumas et al. 2011) can be used to characterize the DRS region. The rotational phase (longitude) is also represented in degrees along the X-axis. The size of the diamond is proportional to the times- pan between the start and end of each observation. with typical exposure times of 300 s in 2007 and 600 s in 2011. Haumea's light curve is taken from Lellouch et al. (2010). The longitude uncertainties are not repre- sented because they are too small (less than 0.13% and 0.07% over the period covering the 2007 and 2011 data). 3. Data reduction The science and calibration data were reduced using the ESO pipeline version 2.3.2 (Modigliani et al. 2007). We first corrected all raw frames from the noise pattern of dark and bright horizon- tal lines introduced when reading the detector. We then used the ESO pipeline to produce all master calibration files needed by the data reduction process, such as the bad pixel masks, master darks, flat-field frames, and the wavelength and distortion cali- bration files, which are needed to associate a wavelength value with each pixel, respectively, and to reconstruct the final image cubes. From each object (science or calibrator) frame, we subtracted the sky frame recorded closest in time. The quality of the sky subtraction was further improved by enabling the correction of sky residuals in the pipeline, i.e., by subtracting the residual me- dian value of the sky from each image slice of the spectral-image cube. We extracted all individual spectra one by one and com- bined them after correction of any remaining bad pixels, which were replaced by the median of all frames at the correspond- ing wavelength. Airmass correction was applied by dividing the spectra of Haumea with their corresponding solar analogs. The final spectra were then normalized to unity at 2.20 µm. We de- termined the rotational phase of each of our spectra from the accurate knowledge of Haumea's rotation period (3.915341h± 0.000005h), based on the closest photometric observations ac- quired in December 2009 (Lellouch et al. 2010) and June 2007 (Lacerda 2009). The 2007 observations were acquired while the DRS region was facing the Earth (see Figure 1). An arbitrary origin for the rotational phase of Haumea was chosen to corre- spond to the lowest of its two light-curve maxima. A detailed F. Gourgeot et al.: Near-infrared spatially resolved spectroscopy of (136108) Haumea's multiple system Table 1. Observation logs for Haumea, with its associated central longitude at each epoch, and the solar analog calibrators. Longitude*** Exp. Time of the frames Name H07-0 H07-1 H07-2 H07-3 H07-4 H07-5 H11-0 H11-1 H11-2 H11-3 H11-4 H11-5 H11-6 HD142093 (G2V) HD107087* (G1V) HD136776* (G5) BD+192951* (G0) HD146233 (G2V) Date & Time** (UT) [yyyy-mm-ss] [hh:mm:ss] 2007-03-15 06:37:27 2007-03-15 06:48:40 2007-03-15 06:54:10 2007-03-15 07:05:24 2007-03-15 07:10:53 2007-03-15 07:16:33 2011-04-09 03:52:03 2011-04-09 04:32:30 2011-04-09 04:53:41 2011-04-09 05:04:12 2011-04-09 05:25:26 2011-04-09 05:35:55 2011-04-09 07:15:36 2007-03-15 07:40:10 2011-04-09 01:55:34 2011-04-09 06:02:45 2011-04-09 06:22:18 2011-04-09 06:43:25 MJD [days] 54174.2760 54174.2838 54174.2876 54174.2954 54174.2992 54174.3032 55660.1612 55660.1893 55660.2040 55660.2113 55660.2260 55660.2333 55660.3025 N.A. N.A. N.A. N.A. N.A. [◦] 30 47 55 72 81 89 58 120 153 169 201 217 10 N.A. N.A. N.A. N.A. N.A. [s] 300 300 300 300 300 300 600 600 600 600 600 600 600 0.83 1 1 2 0.83 Airmass 1.398-1.400 1.398-1.398 1.399-1.398 1.405-1.402 1.409-1.405 1.415-1.410 1.531-1.498 1.425-1.409 1.396-1.387 1.386-1.381 1.379-1.380 1.380-1.385 1.589-1.636 1.406-1.408 1.578-1.639 1.410-1.431 1.397-1.405 1.083-1.084 Seeing ["] 0.81-1.13 0.86-1.05 0.86-1.05 0.84-1.00 0.83-0.98 0.81-0.90 1.21-1.36 0.91-1.25 1.19-1.08 1.03-0.90 0.80-0.80 0.80-1.10 0.89-1.11 0.68-0.83 1.24-1.50 0.77-0.95 0.85-1.26 0.85-0.86 * Solar analog present in four frames for each 2011 observation. ** Middle time of the observation. *** The sub-Earth longitude of Haumea. The origin corresponds to the lowest maximum of the light curve (in the DRS region). analysis of the combined 2007 and 2011 observations is shown in the next section. 4. Spectral analysis 4.1. Identification of the spectral features and comparison Figure 2.A shows the average spectra obtained in 2007 and 2011 in the 1.4 -- 2.4 µm range. We also present the spectra correspond- ing to the different regions (DRS and no-DRS). These four spectra look broadly similar with no major differ- ences between the DRS and the no-DRS regions, and the main spectral features are that of water ice, with deep absorption bands at 1.5 µ and 2 µm. The shapes of these bands, and their relative depths between 1.50 and 1.57 µm, and around 1.65 µm, confirm that most of the H2O ice is in its crystalline state (amorphous ice lacks the 1.57 and 1.65 µm absorption bands, while the 1.5 and 2.0 µm band shapes remain; see Grundy & Schmitt 1998). This predominance of crystalline ice is not surprising, since it is the most thermodynamically stable phase of water ice and has been widely reported on the surface of other icy outer solar sys- tem objects (e.g., satellites of giant planets and transneptunian objects). It is important to note that the spectral difference observed in the 1.63-1.8 µm range is due to a feature of the solar analog (HD142093) used for our 2007 observations. It is clearly visi- ble in Fig. 2.A, between 1.63-1.80 µm (gray region in Fig. 2.B), where we present the ratio between both dates for Haumea and the solar analog spectra. We decided not to remove this artifact in the spectra to conserve all the observed properties, but we ex- cluded this wavelength region from our analysis of the spectral slope. We could not extract the spectra of the satellites because of their high proximity with Haumea. 4.2. Study of the spectral slope As a result of the above analysis, we only considered the spec- tral ranges between 1.51-1.63 µm in H and 2.03-2.30 µm in K (represented in blue in Fig. 2.B) to measure the overall spectral slope. We determined a slope of 0.81 (± 0.60) %/100 nm for the ratio between the data acquired in 2007 and in 2011, and a slope of 1.49 (± 0.60) %/100 nm for the ratio between the data taken inside and outside the DRS. These values confirm thereby a redder surface for the DRS region. This behavior was so far only known for the visible part of the spectrum; present mea- surements are the first detection of this redder slope in the near infrared. We refined the measurements by computing the ratio of each spectrum with the average spectrum of the no-DRS region (H11_x observations, with x = [2,3,4,5]). The spectral slopes are presented in Fig. 3 and were obtained from a simple linear fit of our data over the same wavelength range as Fig. 2. The obtained slope coefficients are presented in Fig. 4, which shows the evolution of the spectral slope versus time. The disper- sion of the data points is larger for the DRS region, which is also shown by their larger error bars. Still, the average slope obtained from all DRS measurements point to a redder surface in this re- gion. Similarly, the points with the maximum slope values (3.25 and 3.26 %/100 nm) are located at Haumea's longitudes 30◦ and 81◦, respectively, i.e., within the DRS region. Even though better data are needed to improve our measurements, we can confirm from this analysis that the DRS region is spectrally redder than the rest of the surface of Haumea. 4.3. Analysis of the 2.0 µm band depth The depth of an absorption feature is computed as Dλ[%] = 100 × (1 − fλ), where fλ is the normalized reflectance at the wavelength λ. This formula gives the percentile value of the absorption depth of a (1) Article number, page 3 of 10 A&A proofs: manuscript no. Gourgeot_Haumea Fig. 2. A) Average spectra of Haumea obtained for each date (top) and for both regions: the DRS (including all the 2007 observations with H11_0 and H11_6) and no-DRS (all combined H11_x with x=[1..5]). See Table 1 and Fig. 1 for notations. The standard deviation at 1 σ is also represented for each spectrum. Water-ice features are clearly visible at 1.5 µm, 1.65 µm (crystalline), and with the large band at 2.0 µm, but no significant differences were observed between each regions. All spectra were normalized at 2.20 µm (dashed line). B) Spectral ratios between the two data sets of 2007 and 2011: Top: division of the two average spectra of Haumea and that of the two solar analogs. We note the bump observed in the 1.63-1.80 µm range (gray region), which is due to an artifact of the solar analog calibration star (see Dumas et al. 2011 for more explanation). Bottom: the ratio of the spectra 2007/2011 and DRS/no-DRS. The blue parts of the spectra correspond to the wavelengths range (1.51-1.63 µm in H and 2.03-2.30 µm in K) used to fit the polynomial regressions (red curves). The slopes are [0.81 ± 0.60] %/100 nm and [1.49 ± 0.60] %/100 nm for the 2007/2011 and DRS/no-DRS ratios, respectively, and they confirm the redder profile of the DRS region. given band, which informs us about the relative quantity of the absorber and a possible indication of the path length traveled by the light. To compute this parameter we assume the ratio be- tween the median values of the bottom band [2.00-2.05 µm] and the top band [2.20-2.25 µm], where all the spectra were normal- ized (at 2.225 µm). The results of the band depth are shown in Figure 5. We see that the values of the absorption band depth are in the range 40% to 60% with no significant difference be- tween DRS and no-DRS. However, the smallest water-ice ab- sorption (41.5%) seems to coincide with the smallest longitude (10◦, close to the lowest maximum of the light curve). As a re- sult, the surface material present on the DRS region, even though dominated by crystalline water ice, could correspond to a less hydrated mixture than the remaining surface of Haumea. 4.4. Analysis of the area of H complex band However, the most unambiguous characterization of the surface properties of the DRS might come from the spectral analysis of the H-band area. For this, we investigated the behavior of all Haumea spectra around the 1.50-1.63 µm absorption band (sadly, we could not study the crystalline water-ice band at 1.65 microns because of the artifact introduced by the solar analog Article number, page 4 of 10 used in our 2007 data). The integrated areas of the 1.5 µm water- ice absorption complex were computed by fitting the continuum level on either side of the absorption band (1.50 to 1.63 µm). Then we determined the value corresponding to "one minus the normalized spectrum" over the band interval, following a method proposed by Grundy et al. (2006). The results are shown in Fig- ure 6 where we see the variations of this integrated area with time (and rotational phase). The shallow 1.56 µm band, still included in our wavelength range of study, is also of interest to inves- tigate possible differences between amorphous and crystalline ice, as was suggested in the laboratory (Mastrapa et al. 2008). Figure 6 shows that the strongest absorptions in the H-band re- gion are located within the DRS with a band area almost twice larger (at longitude 10◦) than for the region outside the DRS. Indeed, contrary to the 2.0 µm band behavior (see previously), the highest absorption seems to coincide with the smallest lon- gitude (10◦, close to the lowest maximum of the light curve). This result points to the fact that, although the DRS might repre- sent the less hydrated region of the surface, it could still display a higher concentration of crystalline water ice than the rest of the surface. This result itself is confirmed by the analysis of our spectral modeling explained in the next section. F. Gourgeot et al.: Near-infrared spatially resolved spectroscopy of (136108) Haumea's multiple system Table 2. Model mixtures Slope χ2 red [1.45-2.45 µm] Models 2007 Model 1 Compounds Mixtures (%) Grain size (µm) Normalization Cr. H2O Cr. H2O Am. carbon Cr. H2O Am. H2O Am. carbon Cr. H2O 2007 Previous Model 1 Am. H2O (Dumas et al. 2011) 95.1 2.9 2.0 87 12.8 0.2 73 25 2 2007 Model 2 0.60 at 1.65 µm N/A 0.75 at 1.75 µm 0.22 0.75 at 1.75 µm 0.16 1.83 1.66 N/A 40 20 10 31 100 10 9 10 10 55 50 10 57 46 53 10 2011 Model 1 2011 Model 2 Titan Tholin Cr. H2O Cr. H2O Am. carbon Cr. H2O Cr. H2O Am. H2O Am. carbon 86.2 2.9 10.9 82.3 1.3 5.9 10.5 0.70 at 1.75 µm 0.25 1.23 0.70 at 1.75 µm 0.25 1.21 "Cr." and "Am." mean crystalline and amorphous, respectively. Fig. 3. Comparison of the spectra ratios obtained by dividing each spectrum by the average spectrum of the no-DRS region (H11_x with x=[2,3,4,5]). The slope is calculated considering the wavelength ranges 1.51-1.63 µm and 2.03-2.30 µm (red and blue parts for 2007 and 2011, respectively) and a simple linear regression. All spectra have been nor- malized at 2.20 µm (dashed line). The variation of the spectral slopes obtained for each spectra as a function of time is represented in Figure 4. 5. Compositional modeling 5.1. Methods We use the spectral model developed by Hapke (1981, 1993) to investigate the chemical properties of the surface of Haumea. This approach allows us to model the reflectance spectrum and the albedo of a medium, from the physical properties of the dif- ferent chemical compounds present on the surface. The albedo is approximated using Eq. (44) of Hapke (1981): Alb = r0 (0.5 + r0 6 ) + w 8 ((1 + B0) P(0) − 1), (2) Fig. 4. Representation of the spectral slopes (%/100nm) from Figure 3 as a function of time (in MJD-55660 days). The rotational phase (lon- gitude) is also represented on the top axis in degrees. The 2007 data have been shifted by an integer number of rotational period so that they can be overplotted on the 2011 data (see Table 1). The red circles and blue triangles correspond to the 2007 and 2011 data, respectively. A null slope corresponding to the no-DRS part (dashed line) and the black square represents the average of the slope values in the DRS equal to [1.46 ± 0.82] %/100 nm, highlighting the redder spectral slope for this region of the surface of Haumea. The error bars correspond to a standard deviation of 1 σ. Haumea's light curve (solid line) is also represented for clarity. where w is the single-scattering albedo and r0 the bi- hemispherical reflectance, which is purely single-scattering albedo dependent, i.e., 1 − w 1 − w , (3) 1 − √ √ 1 + r0 = where w depends on the optical constants and the size of the particles and is computed for a multicomponent surface that is assumed to be intimately or geographically mixed; see Poulet et al. (2002). The parameter B0 is the ratio of the near-surface contribution to the total particle scattering at zero phase angle Article number, page 5 of 10 Ratios 2007H07_0H07_1H07_2H07_3H07_4H07_5H11_0H11_1H11_2H11_3H11_4H11_5H11_61.41.61.82.02.22.4Wavelength (µm)024681012Reflectances (shifted by 2.0 for clarity)Ratios 20111.41.61.82.02.22.4Wavelength (µm)02468101214 A&A proofs: manuscript no. Gourgeot_Haumea Fig. 5. Variation of the depth of the 2.0 µm absorption band (in %) cal- culated as a function of time (MJD - 55660 days). The 2007 data have been shifted by an integer number of rotational period so that they can be overplotted on the 2011 data (see Table 1). The red circles and blue triangles correspond to the 2007 and 2011 data, respectively. All spec- tra were normalized at 2.225 µm for this study. The error bars are 1-σ uncertainty for both data. Haumea's light curve (solid line) is also repre- sented for clarity. Values corresponding to the absorption band depth are in the [40 -- 60%] range with a smallest water-ice absorption at 10◦ close to the lowest minimum of the light curve with no significant difference between DRS and no-DRS. and P is the phase function. The method follows the instruction provided by Merlin et al. (2010), assuming an albedo approx- imation model with a phase angle equal to 0, and B0 close to 0.67 for icy objects (Verbiscer & Helfenstein 1998). The surface roughness and interference were neglected in this work. 5.2. Results for intimate mixtures The free parameters are the asymmetry parameters of the phase function (which is approximated by a single Henyey-Greenstein function), the particle sizes, and the abundances of the various compounds. In order to take the blue component in our model into account, we add two other parameters, a and b, to adjust the blue continuum of the spectrum with a second degree polyno- mial curve. We note x the wavelength, given by x = (λ[i]-1.75) µm, to normalize the spectra at 2.24 µm with an albedo of 0.67. In each case, the model is adjusted to the spectra using this con- tinuum. The result is shown in Figure 7. The normalization at 1.75 µm was correlated for each date (geometrical albedo: 0.75 and 0.702 for 2007 and 2011, respectively) taking into account the solar analog artifact described earlier and affecting the 2007 data. A previous model for the 2007 observations (Dumas et al. 2011) is also represented with a normalization at 0.60 at 1.75 µm. We considered two sets of data: all 2007 observations in a first part (DRS) and all 2011 data (no-DRS, and excluding H11_6, which corresponds to the DRS region) in the second part. For each date, two models are used: one considering only crys- talline water ice, and a second including some amorphous water ice. To constrain the free parameters, we use a best-fit model based on the Levenberg-Marquardt algorithm to minimize the Article number, page 6 of 10 Fig. 6. Variation of the integrated area of the [1.50-1.63] µm H2O ice complex band as a function of time (MJD - 55660 days). The 2007 data have been shifted by an integer number of rotational period so that they can be overplotted on the 2011 data (see Table 1). The red circles and blue triangles correspond to the 2007 and 2011 data, respectively. The errors represented are 1 σ for both data. Haumea's light curve (solid line) is also represented for clarity. We see that the strongest H-band absorptions are localized in the DRS region. reduced χ2. The minimization is applied outside the telluric ab- sorption bands, which can affect the results. Best models are based on an intimate mixture of crystalline water ice (optical constants at 50K from Grundy & Schmitt 1998), amorphous car- bon (from Zubko et al. 1996), and amorphous water ice (at low temperature, from Mastrapa et al. 2008). For the 2007 data, the [1.63-1.80] µm range was considered but with a lower weight than the rest of the spectra to minimize the contamination of the solar analog artifact. All the results are shown in Table 2. Crystalline water ice with grain sizes smaller than 60 µm as the major component, with a few percent of amorphous carbon, are sufficient to fit the data. Some improvements of the fit can be obtained by adding ∼10% of amorphous water ice. Indeed, we see that, for the 2007 data, the presence of amorphous water ice (with large-size 0.1 mm grains) allows us to better fit the spec- trum of the DRS region, as measured by the small difference of the reduced χ2. For the 2011 data, we obtain a nearly similar reduced χ2 between both models (with and without amorphous ice), although we get lower values than for 2007 owing to the analog artifact affecting our older data. If we focus on Model 2 for both epochs, we confirm our earlier result that a higher concentration of crystalline water ice is seen for the DRS region (87.0% for 2007 vs 83.6% for 2011), considering all grain sizes. Finally, we note that the different amounts of amorphous carbon needed to fit the 2011 (more than 10%) and 2007 (less than 2%) also points to differences in composition between the DRS and no-DRS surfaces. It is also important to note that Hershel could not detect any "signature" associated with the DRS in its analy- sis of Haumea's light curve (Lellouch et al. 2010), emphasizing the difficulty of quantifying any variation of surface properties at the spot's location. F. Gourgeot et al.: Near-infrared spatially resolved spectroscopy of (136108) Haumea's multiple system Fig. 8. Comparison of H+K band of SINFONI broadband images of Haumea obtained with the LGS-AO corrected mode for 2007 (left) and 2011 (right). North is up and east is left. The spatial and intensity scales are similar. The color represents the intensity given in ADU units. The improved contrast and spatial resolution of the AO image make the de- tection of the two faint satellites possible: Namaka (the faintest, just below Haumea) and Hi'iaka (the brightest, at the bottom of image) in 2007, only Hi'iaka (faint source northeast of Haumea) in 2011. Fig. 7. Spectra of Haumea (in gray) taken on 2007 (DRS, top figure) and 2011 (no-DRS, bottom figure; excluding data point H11_6, which is right on top of the DRS). The results of our spectral modeling is shown with red and blue lines. Crystalline water ice with grain sizes smaller than 60 µm as the major component, with only a few percent of amorphous carbon, are sufficient to fit our data nicely (Model 1, in blue dashed line). However, slight improvements to the fit can be obtained by adding ∼10% of amorphous water ice (Model 2, in solid red line). See Table 2 for more details on the models. 6. Orbit of Hi'iaka and Namaka 6.1. Astrometric positions Taking advantage of the imaging capabilities of SINFONI, we extracted the relative astrometric position of Haumea and its largest satellite Hi'iaka from the hyperspectral cubes of 2172 x 64 x 64 pixels. We created a broadband image by stack- ing the cube along the wavelength, avoiding wavelength ranges affected by telluric absorption. The resulting image, shown in Fig. 8, contains all the slices in the wavelength intervals 1.524 -- 1.764 µm and 2.077 -- 2.332 µm. We used Moffat 2-D functions to determine the coordinates of the photocenters of the components, following the procedure described in Carry et al. (2011). The spatial accuracy of this mea- surement is about a quarter of a pixel. Owing to the rectangular shape of SINFONI pixels, it is twice as large in the north-south direction (100 mas) than along the east-west direction (50 mas). The relative positions of Hi'iaka with respect to Haumea photo- center are listed in Table 3, and shown in Fig. 9. The brightness difference between Hi'iaka and Haumea, which is normalized to Fig. 9. Relative positions of Hi'iaka on 2011 April 09. Left: The seven positions on the plane of the sky listed in Table 3 are shown as open cir- cles, with their uncertainty (15x40 mas). The bold diamond represents the average position. Right: Evolution of the position with time (top: ∆RA and bottom: ∆DEC). the average brightness of Haumea, that is, removing the effect of Haumea's high-amplitude light curve, is also reported as a mag- nitude difference (∆Mag) for each epoch. With the exception of the last epoch, the magnitude difference ranges from 2.6 to 3.1, with an average of 2.76± 0.18, in agreement with the value of 3.0 reported in Lacerda (2009). The estimation of the flux of the components is rather crude, and is reported here only for in- formation. Namaka is not detected in 2011 image (see Fig. 8), although its detection is within SINFONI capabilities (see Fig. 8 and Dumas et al. 2011). It is likely that Namaka was angularly too close to Haumea at the time of the observations, as suggested by its predicted position at (0(cid:48)(cid:48).008, -0(cid:48)(cid:48).110), based on the orbit by Ragozzine & Brown (2009) and our own computations below, thus less than 2 pixels from the center of Haumea. Article number, page 7 of 10 1.41.61.82.02.22.4Wavelength (µm)0.00.20.40.60.8Geometric Albedo1.41.61.82.02.22.4Wavelength (µm)0.00.20.40.60.8Geometric Albedo2007 Data2007 Model 12007 Model 22011 Data2011 Model 12011 Model 20.150.200.250.30Time (MJD − 55660 Days)0.400.450.500.550.600.65∆DEC (")0.150.200.250.30Time (MJD − 55660 Days)0.300.320.340.360.380.40∆RA (")0.300.320.340.360.380.40∆RA (")0.400.450.500.550.600.65∆DEC (") A&A proofs: manuscript no. Gourgeot_Haumea Fig. 10. Observed positions and model positions of Hi'iaka and Namaka. Figure similar to Fig. 2 of Ragozzine & Brown (2009), with similar scale to ease comparison. From top to bottom, the curves represent the model on-the-sky position of Hi'iaka in the x-direction (i.e., the negative offset in right ascension), Hi'iaka in the y-direction (i.e., the offset in declination), Namaka in the x-direction, and Namaka in the y-direction, all in arcseconds. Symbols represent astrometric observations. We note the much larger time coverage provided by the use of both Keck and VLT/SINFONI data set. Table 4. Dynamical parameters of Hi'iaka and Namaka reported in EQJ2000 and derived mass of Haumea. Hi'iaka Namaka Parameter P a e i Ω ω tp (JD) Mass Value 49.031527 49502.940 0.05260 259.48 192.99 276.14 2452190.3944 3.999 1-σ 0.008980 741.272 0.00599 0.71 0.21 10.40 1.383 0.179 Value 18.323535 25147.829 0.15543 88.83 206.73 143.44 2452167.5299 3.754 1-σ 0.003016 634.162 0.03127 1.13 0.67 11.07 0.763 0.284 Units day km -- deg. deg. deg. day ×1021 kg Comment Orbital period Semi-major axis Eccentricity Inclination Longitude of ascending node Argument of the pericenter Time of pericenter System mass Table 3. Relative positions of Hi'iaka on 2011 April 09. The uncer- tainty is 15 mas in right ascension, and 40 mas in declination. UT 03:52:03 04:32:30 04:53:41 05:04:12 05:25:26 05:35:55 07:15:36 ∆RA ((cid:48)(cid:48)) ∆DEC ((cid:48)(cid:48)) 0.370 0.358 0.340 0.355 0.364 0.372 0.332 0.597 0.530 0.474 0.541 0.573 0.505 0.583 Flux ratio ∆Mag 0.0956 0.0865 0.0699 0.0618 0.0590 0.0619 0.0177 2.9 3.1 2.6 2.6 2.6 2.7 4.1 6.2. Orbital solutions Ragozzine & Brown (2009) built a dynamic solution for Haumea's system on the base of about 30 observations spanning Article number, page 8 of 10 a period of 1260 days from January 2005 to May 2008, acquired with the Hubble Space Telescope (HST). These authors used a dynamic three-body model in which the gravitational interaction between the satellites causes non-Keplerian perturbations on Na- maka's orbit on timescales much longer than a month. However, they found that the perturbation of Namaka on Hi'iaka's orbit was negligible, and that its orbit is described well by a Keple- rian motion. Furthermore, they found no evidence for an effect of Haumea gravitational quadrupole (J2). Using the astrometric positions of Hi'iaka and Namaka measured on SINFONI dat- acubes in 2007, Dumas et al. (2011) also studied the orbits of both satellites and discussed heating that could be generated by tidal effects. We took advantage of another astrometric position, taken 431 days (about nine revolutions) after the latest position reported by Ragozzine & Brown (2009) to determine the orbital elements of F. Gourgeot et al.: Near-infrared spatially resolved spectroscopy of (136108) Haumea's multiple system Hi'iaka. For that, we used the genetic-based algorithm Genoid (Vachier et al. 2012), which relies on a metaheuristic approach to find the best-fit set of orbital elements in a two-body prob- lem. This approach can be used to search for Keplerian orbits, as well as to explore a more complex problem, including the gravitational field of the central body up to the fourth order. To define the set of orbital elements that best fit the data, Genoid minimizes a fitness function, fp, defined as a χ2 minimization function (eq. 4) (cid:32) xo n(cid:88) i=1 i − xc xe i i fp = χ2 = (cid:33)2 + (cid:32) yo i − yc ye i i (cid:33)2 , (4) where n is the number of observations, and xi and yi are the relative position between Hi'iaka and Haumea along the right ascension and declination, respectively. The exponents o and c stand for observed and computed positions, and exponent e stands for measured error. For convenience, in the following text, we express the fitness function as the quadratic mean of the residuals. The main advantage of this fitness function is to provide a link between the quality of the fitted orbit and the uncertainties of the astrometric positions provided by the ephemeris. A definition and discussion about estimating uncertainties in astronomy can be found, for example, in Andrae (2010), Andrae et al. (2010), and references therein. We used a total of 35 astrometric positions, spread over 2264 days or 46.2 orbital periods, to determine the orbit of Hi'iaka. The orbit of Namaka is based on 30 observations spread over 1175 days or 64.1 orbital periods. We added to the set of HST observations, the positions measured on SINFONI spectral cubes in 2007 and 2011 (see above) and the set of observations obtained with NIRC2 on the W. K. Keck II telescope (pixel scale of 9.963 mas) between 2005 and 2008 reported by Brown et al. (2005), Brown et al. (2006), and Ragozzine & Brown (2009). Albeit the intrinsic precision of Keck images is deemed cruder than that of HST, which benefits from highly stable PSF, and the absolute link between HST and Keck astrometry is unknown (pixel scale and field orientation), we favor orbital solutions based on a longer time baseline that are more sensitive to departures from a pure Keplerian motion. Because of these points, however, the goodness of fit is expected to be lower. The addition of Keck data here corresponds to an increase of 350 days of time coverage. Compared to the case using only HST data, as carried out by Ragozzine & Brown (2009), this represents 44% of the temporal baseline. This baseline is longer with the addition of the SINFONI measurements in 2011, the leverage provided by Keck data is of 20% (see Fig. 10). 6.3. Results We run two separate orbital fits: one assuming the central ob- ject (Haumea) to be a point-like mass and the satellites to be massless test particles (i.e., Genoid-Kepler), and one considering the zonal harmonic coefficients J2 and J4 of the central object, its size, and the coordinates of its pole of sidereal rotation (i.e., Genoid-ANIS). Our best-fit solution is obtained using a pure Keplerian motion with a fitness function fp = 11.3 mas for Hi'iaka and fp = 17.3 mas for Namaka. This corresponds to reduced χ2 of 1.61 and 2.5, respectively. This level of accuracy is typical of the pixel size of the Keck NIRC2 camera of ≈10 mas, which is 5 and 10 times smaller than the pixel size of HST WFPC2 Fig. 11. Residuals (observed minus computed satellite -- primary posi- tions) normalized by the positional uncertainty σ for both Hi'iaka and Namaka (filled and open symbols, respectively). The three circles rep- resent the 1-, 2-, and 3 -- σ contours. The color enlightens the different epochs of observation. The redder dot corresponds to the observation by SINFONI in 2011. (≈50 mas) and ACS (≈100 mas) cameras, respectively. These re- duced χ2 are to be compared with the reduced χ2 of 1.1 reported by Ragozzine & Brown (2009), based on a fit of both HST and Keck data. Our calculations show clearly that there is no signature for non-Keplerian motion that would be caused by Haumea gravi- tational quadrupole, confirming Ragozzine & Brown (2009) re- sults. Although the orbits determined here are restrained to the two-body assumption, compared with the three-body integration by Ragozzine & Brown (2009), they still allow us to predict the position of both satellites to a high accuracy over the pe- riod 2005-2020. Our Virtual Observatory web service Miriade1 (Berthier et al. 2009) allows anyone to compute the ephemeris of Hi'iaka and Namaka for any arbitrary epoch and observer loca- tion, and can be used to plan future observations of the system. Figure 11 shows the residual positions (between observed and computed) obtained normalized by the positional uncer- tainty. We list in Table 4 the orbital parameters of Hi'iaka and Namaka obtained with Genoid-Kepler. They correspond to the same geometry as the orbits presented by Ragozzine & Brown (2009), albeit Namaka eccentricity is found to be lower (0.15 here vs 0.25). We note the existence of a numerical solution with similar residuals ( fp =10.7 mas) in which the orbital motion of Hi'iaka is reversed. We do not consider this solution here. 7. Conclusion We presented a rotationally resolved spectroscopic study of the surface of the transneptunian object Haumea derived from the 2007 and 2011 near-infrared (1.4 -- 2.5 µm) observations obtained with the integral-field spectrograph SINFONI at the ESO VLT. Crystalline water ice is confirmed to be the major compound present over the entire surface of Haumea. We detected a steeper spectral slope in the near-infrared associated with the so-called 1 http://vo.imcce.fr/webservices/miriade/?ephemcc Article number, page 9 of 10 A&A proofs: manuscript no. Gourgeot_Haumea DRS region over which the visible spectrum is redder, as re- ported by Lacerda (2009). We find that the reddening of the spec- trum extends up to 2 microns. A detailed analysis of the depth of the 2.0 µm absorption band shows that the weakest water-ice absorption is measured for 10◦ of rotation phase, i.e., near the center of the DRS re- gion. Similarly, the analysis of the water absorption band around 1.6 µm reveals that the concentration of crystalline water ice is highest at the DRS location. Hapke modeling of our spectra also shows that amorphous carbon appears to be depleted in the DRS region in comparison to the rest of the surface. Although no major compositional differences have been found between the DRS region and the remaining surface, all the evidence hints at a slightly different composition for this region. New astrometric positions of Hi'iaka were reported and an orbital solution was computed with the genetic-based algorithm Genoid. The results are in agreement with the orbits originally presented by Ragozzine & Brown (2009) and ephemeris genera- tion is proposed to the community. Acknowledgements. We thank Darin Ragozzine for providing the expected po- sitions of both satellites from his modeling at the time of our observations. References Andrae, R. 2010, ArXiv e-prints : 1009.2755 Andrae, R., Schulze-Hartung, T., & Melchior, P. 2010, ArXiv e-prints : 1012.3754 Barkume, K. M., Brown, M. E., & Schaller, E. L. 2006, ApJ, 640, L87 Barucci, M. A., Merlin, F., Guilbert, A., et al. 2008, A&A, 479, L13 Berthier, J., Hestroffer, D., Carry, B., et al. 2009, in European Planetary Science Congress 2009, 676 Bonnet, H., Abuter, R., Baker, A., et al. 2004, The Messenger, 117, 17 Brown, M. E., Bouchez, A. H., Rabinowitz, D., et al. 2005, ApJ, 632, L45 Brown, M. E., van Dam, M. A., Bouchez, A. H., et al. 2006, ApJ, 639, L43 Carry, B., Hestroffer, D., DeMeo, F. E., et al. 2011, Astronomy and Astrophysics, 534, A115 544, A137 Carry, B., Snodgrass, C., Lacerda, P., Hainaut, O., & Dumas, C. 2012, A&A, Dumas, C., Carry, B., Hestroffer, D., & Merlin, F. 2011, A&A, 528, A105 Dumas, C., Merlin, F., Barucci, M. A., et al. 2007, A&A, 471, 331 Eisenhauer, F., Abuter, R., Bickert, K., et al. 2003, in Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series, Vol. 4841, Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series, ed. M. Iye & A. F. M. Moorwood, 1548 -- 1561 Grundy, W. M. & Schmitt, B. 1998, J. Geophys. Res., 103, 25809 Grundy, W. M., Young, L. A., Spencer, J. R., et al. 2006, Icarus, 184, 543 Lacerda, P. 2009, AJ, 137, 3404 Lacerda, P., Jewitt, D., & Peixinho, N. 2008, AJ, 135, 1749 Lacerda, P. & Jewitt, D. C. 2007, AJ, 133, 1393 Lellouch, E., Kiss, C., Santos-Sanz, P., et al. 2010, A&A, 518, L147 Mastrapa, R. M., Bernstein, M. P., Sandford, S. A., et al. 2008, Icarus, 197, 307 Merlin, F., Barucci, M. A., de Bergh, C., et al. 2010, Icarus, 208, 945 Merlin, F., Guilbert, A., Dumas, C., et al. 2007, A&A, 466, 1185 Modigliani, A., Hummel, W., Abuter, R., et al. 2007, ArXiv Astrophysics e- prints Icarus, 160, 313 Poulet, F., Cuzzi, J. N., Cruikshank, D. P., Roush, T., & Dalle Ore, C. M. 2002, Rabinowitz, D. L., Barkume, K., Brown, M. E., et al. 2006, ApJ, 639, 1238 Ragozzine, D. & Brown, M. E. 2007, AJ, 134, 2160 Ragozzine, D. & Brown, M. E. 2009, AJ, 137, 4766 Schaller, E. L. & Brown, M. E. 2008, ApJ, 684, L107 Snodgrass, C., Carry, B., Dumas, C., & Hainaut, O. 2010, A&A, 511, A72 Tegler, S. C., Grundy, W. M., Romanishin, W., et al. 2007, AJ, 133, 526 Trujillo, C. A., Brown, M. E., Barkume, K. M., Schaller, E. L., & Rabinowitz, D. L. 2007, ApJ, 655, 1172 Vachier, F., Berthier, J., & Marchis, F. 2012, A&A, 543, A68 Verbiscer, A. & Helfenstein, P. 1998, in Astrophysics and Space Science Library, Vol. 227, Solar System Ices, ed. B. Schmitt, C. de Bergh, & M. Festou, 157 Zubko, V. G., Mennella, V., Colangeli, L., & Bussoletti, E. 1996, MNRAS, 282, 1321 Article number, page 10 of 10
1807.08799
1
1807
2018-07-23T19:33:46
WEIRD: Wide-orbit Exoplanet search with InfraRed Direct imaging
[ "astro-ph.EP", "astro-ph.SR" ]
We report results from the Wide-orbit Exoplanet search with InfraRed Direct imaging (WEIRD), a survey designed to search for Jupiter-like companions on very wide orbits (1000 to 5000 AU) around young stars ($<$120 Myr) that are known members of moving groups in the solar neighborhood ($<$70 pc). Sharing the same age, distance, and metallicity as their host while being on large enough orbits to be studied as "isolated" objects make such companions prime targets for spectroscopic observations and valuable benchmark objects for exoplanet atmosphere models. The search strategy is based on deep imaging in multiple bands across the near-infrared domain. For all 177 objects of our sample, $z_{ab}^\prime$, $J$, [3.6] and [4.5] images were obtained with CFHT/MegaCam, GEMINI/GMOS, CFHT/WIRCam, GEMINI/Flamingos-2, and $Spitzer$/IRAC. Using this set of 4 images per target, we searched for sources with red $z_{ab}^\prime$ and $[3.6]-[4.5]$ colors, typically reaching good completeness down to 2Mjup companions, while going down to 1Mjup for some targets, at separations of $1000-5000$ AU. The search yielded 4 candidate companions with the expected colors, but they were all rejected through follow-up proper motion observations. Our results constrain the occurrence of 1-13 Mjup planetary-mass companions on orbits with a semi-major axis between 1000 and 5000 AU at less than 0.03, with a 95\% confidence level.
astro-ph.EP
astro-ph
DRAFT VERSION JUNE 16, 2021 Typeset using LATEX twocolumn style in AASTeX62 8 1 0 2 l u J 3 2 . ] P E h p - o r t s a [ 1 v 9 9 7 8 0 . 7 0 8 1 : v i X r a WEIRD: Wide-orbit Exoplanet search with InfraRed Direct imaging FRÉDÉRIQUE BARON,1 ÉTIENNE ARTIGAU,1 JULIEN RAMEAU,1 DAVID LAFRENIÈRE,1 JONATHAN GAGNÉ,2, 3 LISON MALO,1 LOÏC ALBERT,1 MARIE-EVE NAUD,1 RENÉ DOYON,1 MARKUS JANSON,4 PHILIPPE DELORME,5 AND CHARLES BEICHMAN6 1Institut de Recherche sur les Exoplanètes, Département de Physique, Université de Montréal, Montréal, QC H3C 3J7, Canada 2Carnegie Institution of Washington DTM, 5241 Broad Branch Road NW, Washington, DC 20015, USA 3NASA Sagan Fellow 4Department of Astronomy, Stockholm University, SE-106 91 Stockholm, Sweden 5Université de Grenoble Alpes, CNRS, IPAG, F-38000 Grenoble, France 6NASA Exoplanet Science Institute, California Institute of Technology, Jet Propulsion Laboratory, Pasadena, CA 91125, USA ABSTRACT We report results from the Wide-orbit Exoplanet search with InfraRed Direct imaging (WEIRD), a survey designed to search for Jupiter-like companions on very wide orbits (1000 to 5000 AU) around young stars (<120 Myr) that are known members of moving groups in the solar neighborhood (<70 pc). Sharing the same age, distance, and metallicity as their host while being on large enough orbits to be studied as "isolated" objects make such companions prime targets for spectroscopic observations and valuable benchmark objects for exoplanet atmosphere models. The search strategy is based on deep imaging in multiple bands across the near-infrared domain. For all 177 objects of our sample, z(cid:48) ab, J, [3.6] and [4.5] images were obtained with CFHT/MegaCam, GEMINI/GMOS, CFHT/WIRCam, GEMINI/Flamingos-2, and Spitzer/IRAC. Using this set of 4 images per target, we searched for sources with red z(cid:48) ab and [3.6] − [4.5] colors, typically reaching good completeness down to 2 MJup companions, while going down to 1 MJup for some targets, at separations of 1000−5000 AU. The search yielded 4 candidate companions with the expected colors, but they were all rejected through follow-up proper motion observations. Our results constrain the occurrence of 1 -- 13 MJup planetary- mass companions on orbits with a semi-major axis between 1000 and 5000 AU at less than 0.03, with a 95% confidence level. Keywords: Exoplanets -- Direct Imaging -- Brown Dwarfs 1. INTRODUCTION Since the first detection of an exoplanet around a main se- quence star by Mayor & Queloz (1995), thousands of ex- oplanets have been discovered, revealing planetary system architectures vastly different from that of the Solar System. The most successful methods to detect exoplanets are the transit and radial velocity methods, which are more effective for planets close to their host star (up to 15 AU). The outer part of planetary systems can be probed by direct imaging. The first detection of a giant planet by direct imaging was made in 2004, with the discovery of a 4 MJup planet orbit- ing the brown dwarf 2MASSW J1207334-393254 (Chauvin et al. 2004), and the search for directly imaged planets has continued since then. A good number of direct imaging surveys for planetary mass objects on wide obits were carried out in the last decade. Some targeted only low mass stars, such as Bowler et al. (2015), Lannier et al. (2016), and Naud et al. (2017), while others surveyed higher mass stars (Vigan et al. 2012; Nielsen et al. 2013; Rameau et al. 2013) or all spectral types (Lafrenière et al. 2007b; Heinze et al. 2010; Biller et al. 2013; Chauvin et al. 2015). Bowler (2016) did a meta-analysis us- ing data from the most complete studies that surveyed all types of star (Lafrenière et al. 2007b; Janson et al. 2011; Vi- gan et al. 2012; Biller et al. 2013; Janson et al. 2013; Nielsen et al. 2013; Wahhaj et al. 2013; Brandt et al. 2014; Bowler et al. 2015) using 384 stars with spectral types B2 to M6. He obtained an overall planet occurrence rate for BA, FGK, and M stars of respectively 2.83.7 −2.3 %, <4.1%, and <3.9% for 5-13 MJup planets at separations of 30 to 300 AU. Direct imaging surveys have typically targeted young stars, which are prime targets since their planets are still contract- ing and are thus warmer and brighter than their older coun- terparts, for a given mass. The number of known young stars near the Sun has dramatically increased in the last few years, as a result of a growing interest for young stellar moving groups (e.g., Zuckerman & Song 2004; Torres et al. 2008). A moving group is composed of stars that were formed together less than a few hundreds of Myr ago, and therefore still share similar UVW galactic velocities, enabling their identifica- 2 BARON ET AL. tion. In the recent years, a significant effort has been made to identify a large number of low mass stars, brown dwarfs and isolated planetary-mass objects that are members of known young moving group (Lépine & Simon 2009; Shkolnik et al. 2009; Kiss et al. 2011; Schlieder et al. 2010; Rodriguez et al. 2011; Liu et al. 2013; Schlieder et al. 2012a,b; Shkolnik et al. 2012; Malo et al. 2013; Moór et al. 2013; Rodriguez et al. 2013; Kraus et al. 2014; Malo et al. 2014; Riedel et al. 2014; Gagné et al. 2014; Binks et al. 2015; Gagné et al. 2015b). Planets found on wide orbits around young stars are very interesting because they can be characterized much better than their closer-in counterparts. First, a planet bound to a well-studied star shares some properties with it, like its age, distance from Earth, and metallicity. Furthermore, when a planet is on an large enough orbit, it can be studied as if it were an isolated object, that is without adaptive optics (Naud et al. 2014; Gauza et al. 2015), and a very high res- olution spectrum can then be acquired, which is very hard to obtain for closer-in planets. Also, the large separation to the host enables direct studies that are very challenging with high-contrast imaging (e.g., accurate spectro-photometry, in- termediate resolution spectroscopy, optical imaging, time- variability). Such planetary mass objects are also prime tar- gets for JW ST follow-up. Widely separated systems are of further interest because they challenge formation processes. Theories predict that gi- ant planets form either by core accretion or gravitational in- stability, or like brown dwarfs by cloud fragmentation. The former process describes a way of forming planets by first building a 5 to 20 MEarth core of rocks and ices, in a proto- planetary disk (Alibert et al. 2009; Inaba et al. 2003; Pollack et al. 1996). The core then accretes gas very rapidly to form a giant planet. This method explains very well the formation of planets on close-in orbit (< 10 AU, Mordasini et al. 2012), but struggles to explain the formation of planets on wide or- bits. The second process suggests that planets form from the fragmentation of a gravitationally unstable disk (Boss 2011), which forms clumps that then can accrete gas and dust to be- come planets (Stamatellos et al. 2007; Bate 2012). However, this mechanism also has difficulties forming planets on wide orbits, as shown for example by Nayakshin (2017) and Vigan et al. (2017). The last process predicts that planets on wide orbits form from the direct collapse of the molecular cloud (Padoan & Nordlund 2004). A fragment of a few Jupiter mass is formed which then accretes gas from the cloud to form a higher mass object. However, Bate et al. (2002) and Bate (2012) have shown that the accreation process can be stopped at a low mass if the companion is ejected away from the dense part of the envelope or if the envelope is depleted at the formation time. However this formation process tends to form preferentially equal mass binaries and does not seem to produce systems with the high mass ratios needed to match the observed planetary systems at wide separations. Dynam- ical instabilities are a tantalizing alternative to explain the detected planets at large separations (Chatterjee et al. 2008; Veras et al. 2009; Baruteau & Masset 2013). Mutual gravita- tional perturbations and close encounters among the planets occur and pump the eccentricity and the semi-major axis of the less massive giant planets up to 100 - 100 000 AU (Veras et al. 2009), but close-in scatterers are yet to be discovered (Bryan et al. 2016). We report here the results from the Wide-orbit Exoplanet search with InfraRed Direct imaging (WEIRD). The WEIRD survey started in 2014 with the aim to detect Jupiter-like com- panions on very wide orbits (at separations 1000 -- 5000 AU) around all the known members of young moving groups within 70 pc. We gathered a large dataset to try to con- struct the SED of such objects through of deep [3.6] and [4.5] imaging from Spitzer/IRAC and deep seeing-limited J and z(cid:48) ab imaging from CFHT and Gemini-South of all 177 known (at the time) young (<120 Myr) objects in a volume-limited sample of 70 pc of the Sun. Using these data, planetary com- panions can be revealed through their distinctively red z(cid:48) ab − J and [3.6]−[4.5] colors. The range of separations studied here has been barely probed by previous direct imaging surveys as they were limited by the field of view of high contrast im- agers, with the exception of Naud et al. (2017), which was much less sensitive than the present survey, and limited to low-mass stars. The selection of the sample of young stars and the observing strategy and data reduction are described in Section 2. Section 3 presents the results of our search, while Section 4 discusses the statistical analysis of the sur- vey. 2. SAMPLE AND OBSERVATIONS 2.1. Sample The best targets to find giant planets on very wide orbits are young stars in the solar neighborhood because giant plan- ets are more luminous when they are younger and they be- come fainter with time. Therefore, observations of younger stars are sensitive to lower-mass planets. A sample was thus created by selecting all stars within 70 pc that are mem- bers of the following young moving groups or associations (see Table 1): TW Hydrae (de la Reza et al. 1989; Kastner et al. 1997), β Pictoris (Zuckerman et al. 2001a), AB Do- radus (Zuckerman et al. 2004), Tucana Horologium (Torres et al. 2000; Zuckerman et al. 2001b), Carina (Torres et al. 2008), Columba (Torres et al. 2008) and Argus (Makarov & Urban 2000). The members of these groups have ages in the range 10 -- 150 Myr. The age of the Argus moving group is not well constrained, likely because current membership lists suffer from significant contamination from unrelated field-aged stars (Bell et al. 2016). To be considered bona fide members of one group and included in our sample, the WEIRD Table 1. Young Moving Groups 3 Name Distance Nβ-Pictoris AB-Doradus Argus Carina Columba Tucana-Horologium TW Hydrae (pc) 9-73 37-77 8-68 46-88 35-81 36-71 28-92 Age (Myr) 24±3 149+51 −19 30-50 45+11 −7 42+6 −4 45±3 10±3 Number of members detected Ref. 51 58 10 6 15 50 16 Shkolnik et al. (2017) Bell et al. (2016) Torres et al. (2008) Bell et al. (2016) Bell et al. (2016) Bell et al. (2016) Bell et al. (2016) stars must have a trigonometric parallax and a radial veloc- ity measurement, XY ZUVW values consistent with the mov- ing group membership, as well as independent signatures of youth, e.g. spectroscopic signs of low-gravity, strong X-ray or UV emission or lithium absorption (see Soderblom 2010). The sample was constructed from Gagné et al. (2014); Kiss et al. (2011); Lépine & Simon (2009); Malo et al. (2013); Schlieder et al. (2010) ; Shkolnik et al. (2009) ; Shkolnik et al. (2011) ; Shkolnik et al. (2012) ; Song et al. (2003); Tor- res et al. (2000) ; Torres et al. (2008); Zuckerman & Webb (2000); Zuckerman et al. (2004); Zuckerman et al. (2001a) ; Zuckerman (2001) ; Zuckerman et al. (2011). We note that these publications also proposed a larger sample of strong candidates but they lacked one or more measurements to be confirmed members; these objects were not included in our sample. Our complete sample includes 177 objects. Multiple systems were not excluded from the sample as the presence of a lower or equal mass object does not exclude the possibility of having a planetary mass object on a wide orbit. For example, Ross 458 (AB)c is a triple system comprising a tight M0.5+M7 binary orbited by an 11 MJup object (Gold- man et al. 2010) and 2MASS J01033563-5515561(AB)b, a 12 -- 14 MJup object, orbits a pair of young late-M stars at 84 AU(Delorme et al. 2013). Also, Wang et al. (2015) have shown that stellar multiplicity does not influence the pres- ence of planets on wide (100 to 2000 AU) orbits in the sys- tem. In our sample of targets, 68 are multiple systems, 2 host brown dwarf companions (Schneider et al. 2004; Chau- vin et al. 2005) and 4 host known planets (Chauvin et al. 2004; Marois et al. 2008; Lagrange et al. 2009; Macintosh et al. 2015). The properties of the 177 objects in our sample is presented in Table 2. They have spectral classes in the range A -- M, with a majority of M dwarfs, are located at distances of 7 to 70 pc, are located all over the sky, and have relatively high proper motions (see Figure 1). The median star has a proper motion of 100 mas/yr, a distance of 42 pc and an age of 45 Myr. Table 3 lists all the systems in our sample along with their radial velocities, distances and the association they are a member of. 2.2. Observing strategy Figure 2 presents the typical z(cid:48) ab − J and [3.6] − [4.5] colors as a function of spectral type for objects ranging from spec- tral types L to T, a range relevant for the companions sought here. It shows that both L and T dwarfs have very red z(cid:48) ab − J colors, with the color of an L dwarf being between 2.5 and 3 mag, and the color of a T dwarf between 3 and 4.5 mag. Beyond those types, as shown by Lodieu et al. (2013), the z(cid:48) ab − J colors of Y dwarfs remain red but vary much more, ranging from from 2.5 to 5 mag. In the mid-infrared, start- ing at around T0, the [3.6]−[4.5] color becomes increasingly red with spectral type, reaching values larger than 1.5 mag for late-T's. Young objects, with larger radii and correspond- ingly lower surface gravities, would have slightly redder col- ors compared to the colors of field dwarfs shown in the fig- ure. The strategy used in the current survey builds on these markedly red colors of very late-type dwarfs across these four spectral bands, which enables distinguishing them eas- ily from earlier-type objects and most other astrophysical sources. In addition, as shown in Figure 3, these bands are also optimal to maximize the flux of the objects sought over the temperature range of interest. The ground-based component of our survey is optimized to find companions up to spectral type ∼T9, while the Spitzer component is optimized for later types. At a distance of 42 pc (the median distance of our sample), the expected J mag- nitude of a T9 dwarf is about 21 mag. We thus designed our observations in J-band to reach 21 mag. As T dwarfs later than T0 are expected to have z(cid:48) ab − J > 3 mag, we de- signed our observations to reach z(cid:48) ab = 24 mag as they can be identified either through detection in both bands or as z(cid:48) ab dropouts. For the this part of the survey, we used the same detection criteria as for the CFBDSIR survey (Delorme et al. 2008, 2010; Albert et al. 2011). That survey was a wide- field search for T dwarfs and early-type Y dwarfs, and the candidates were identified through their very-red z(cid:48) ab − J > 3 mag colors if they were detected in both bands, or through z(cid:48) abdropouts. The CFBDSIR survey returned only 64 candi- dates over the 280 square degrees observed, of which 17 were actual field T dwarfs. The strategy of searching for very low- 4 BARON ET AL. Figure 1. Distributions of associations, distances (pc), spectral types and proper motions (mas/yr) for all the stars in the sample. mass objects using NIR colors has also been employed by the PSYM-wide survey (Naud et al. 2017) to probe nearby young M dwarfs for planetary mass companions. The survey discovered a new planetary mass object (9 -- 13 MJup) orbiting at 2000 AU around the M3V star Gu Psc, a highly proba- ble member of the AB Doradus moving group (Naud et al. 2014)1 . The Spitzer/IRAC observations were designed so that they reach a sufficient depth to identify point-sources in the field down to ∼ 0.5 mag of the confusion limit and have their color measured with accuracy to identify them at > 5σ level compared to the bulk of background objects. We perform the point-source detection in [3.6], which provides deeper images for flat-spectrum sources, and use the [4.5] pho- tometry to constrain colors. Our observations are designed to reach depths of 21.2 mag (5σ) and 20.7 mag (3σ) in [3.6] and [4.5], respectively. Planetary-mass objects close 1 Using the parallax of 21.0019 ± 0.0721 mas (Lindegren et al. 2018; Gaia Collaboration et al. 2018) for GU Psc from the Gaia DR2 release, and the web tool BANYAN Σ from Gagné et al. (2018b), we infer that the probability of Gu Psc to be a member of the AB Doradus moving group is 99.1 %, which confirms the membership of the star. ab- and J-band dropouts (z(cid:48) to the detection limit, with masses below 3-5 MJup will have [3.6] − [4.5] > 2 mag (see right-hand side Figure 2) and will therefore be detected at a higher signal-to-noise ratio in [4.5]. Because they would have MJ > 18 mag (or J > 21 mag for a typical target) or a spectral type (cid:38)T8.5, such objects would be z(cid:48) ab − [4.5] > 6 mag and J −[4.5] > 3.5 mag). Given that background objects typically have [3.6] − [4.5]∼ 0.0± 0.4 mag, such planetary-mass ob- jects will differ from the bulk of background objects at the >5σ-level. However, by itself this part of the dataset is in- sensitive to more massive (> 3-5 MJup) companions as their colors don't differ enough from those of background objects. 2.3. Observations and Data Reduction All targets in our sample were observed with deep seeing- limited J and z(cid:48) ab imaging at either the Canada-France- Hawaii Telescope (CFHT) with WIRCam (Puget et al. 2004) and Megacam (Boulade et al. 1998), or at Gemini-South with GMOS-S (Hook et al. 2004; Gimeno et al. 2016) and Flamingos-2 (Eikenberry et al. 2012), as well as with Spitzer/IRAC (Fazio et al. 2004) in the [3.6] and [4.5] bands. Stars with a declination < −35◦ were observed from the ground at the Gemini-South Observatory while the others TWBPICARGCOLTUCCARABDAssociations0102030405060Number fo stars102440424545149Age (Myr)10203040506070Distance (pc)0102030405060Number fo starsBAFGKMSpectral Types0204060Number fo stars02004006008001000Proper motion (mas/year)020406080Number fo stars WEIRD 5 were observed at the CFHT. Throughout this work, all the J-band magnitudes are in the Vega system while all the z(cid:48) ab magnitudes are in the AB system. For a median star in our sample with a distance of about 42 pc and an age of 45 Myr, the limiting magnitude in both bands corresponds to MJ = 17.9 mag and Mz = 20.9 mag, or to an effective temper- ature of about 385 K according to models from Baraffe et al. (2003). 2.3.1. Gemini Observations The observations were made from 2014 to 2017 at Gemini- South (GS-2014B-Q-2, GS-2015A-Q-71, GS-2015B-Q- 57, GS-2016A-Q-69, GS-2016B-Q-33, GS-2017A-Q-58, PI Frederique Baron). We obtained deep imaging of our southern sub-sample with Flamingos-2 with the J filter (J_G0802, 1.255 µm) and the Gemini Multi-Object Spectro- graph (GMOS) in the z(cid:48) ab filter (z_G0328, >848 µm). Objects beyond 30 pc, the vast majority of our targets, are sufficiently far for the entire projected 5000 AU sphere around them to fit within the GMOS/F2 FOV. Flamingos-2 is a near-infrared wide-field imager and multi-object spectrometer with a 6.19 arcmin2 circular field of view and a 0.18(cid:48)(cid:48)pixel scale. We obtained at least 600s of integration time on each target, divided into a different num- ber of expositions (at least 9) depending on the magnitude of the star and the observing conditions. A small random dither pattern was used to mitigate detector defects. The exposition time was selected to reach a limiting magnitude of J = 21 mag at a 7σ level. Each observation was about 20 minutes long, including all overheads. GMOS has three 2048×4608 CCDs which, when com- bined, have a field of view of 5.5x5.5 arcmin2 and a pixel scale of 0.073(cid:48)(cid:48). We obtained 8 expositions of 65 s for each target of the sample. A dither pattern of 17 (cid:48)(cid:48)was used for all observations. The exposure time was chosen to reach a lim- iting magnitude of z = 24 mag at 3σ. The observations were each about 20 minutes long, including all overheads. The J-band images from F2 were reduced using a custom IDL pipeline. The individual images were reduced by sub- tracting dark images, dividing by flat field images, and cor- recting the residual gradient from the vignetting of the Pe- ripheral Wavefront Sensor (PWFS). This step was done by first normalizing the image to its median value, then masking regions with values significantly over the median to get rid of the stars. This image was in turn used to create a gradient image where each pixel is the median of a 128x128 pixel box of the masked image. A polynomial fit of degree 3 was then applied to the gradient image. This was divided from the F2 image to correct for the vignetting by the PWFS. The astro- metric correction was then computed by anchoring the star positions on the Gaia DR1 catalog (Gaia Collaboration et al. 2016). A radial profile about the bright target star was then subtracted to help search for sources at smaller separations. Lastly, a low-pass filter was created by median binning the image by 4x4 pixels, applying a 15x15-pixel median filter, and then resampling at the original image size. This low- pass filter was subtracted from the image to facilitate the de- tection of point sources. The individual images for a given target were then combined by taking their median, after as- trometric registration, to produce the final J-band image. For GMOS, the images were also reduced using a custom IDL reduction pipeline. Each CCD was first processed sepa- rately. First, a sky correction was applied by subtracting the median of all images taken on a given night. When needed, any detector region affected by the on-instrument wave front sensor was masked. Most of the time the wave front sen- sor was off the detector, but sometimes it was not possible to find a guide star outside of the FOV. The astrometric solution was found for each CCD by anchoring the field to the F2 re- duced image of the same target. A high-pass filter was then applied by subtracting a median-filtered image with a width of 15 pixels. A one-dimensional median-filter with a width of 61 pixels was also subtracted from each line of the image to correct for the saturation banding. The 3 CCDs were then combined to form a complete image, to which the astrometric solution was applied again. All the images for a given target were then aligned and stacked by taking their median to get the resulting reduced image. 2.3.2. CFHT Observations Deep imaging of our northern sub-sample was obtained at the CFHT from 2014 to 2017 using WIRCam with the J fil- ter and MegaCam with the z(cid:48) ab filter (14BC016, 15AC032, 15BC012, 16AC021, 16BC018, 17AC23; PI Frédérique Baron). WIRCam (Puget et al. 2004) is a near infrared wide-field imager with a field-of-view of 20 arcmin2 and a pixel scale of 0.3(cid:48)(cid:48). It uses a mosaic of 4 detectors with a small gap between each. We used the J-band (1.253 µm) filter and a homemade dither pattern of 16 60-s expositions, arranged so that the target does a small dither of 28(cid:48)(cid:48)around a pixel sit- uated 64(cid:48)(cid:48)from the corner of one detector near the center of the field, for a total of 1120 s of on-target integration time. A different dither pattern was used to mitigate the saturation effects of stars brighter than J = 7. In those cases, the bright target was put in a gap between quadrants at each position of the dither pattern. With a seeing between 1(cid:48)(cid:48)and 1.2(cid:48)(cid:48), the exposure time is sufficient to reach a SNR=7 at a limiting magnitude of J ≈ 21. MegaCam (Boulade et al. 1998) is a wide-field optical im- ager with a 1 square degree field-of-view and a pixel scale of 0.187(cid:48)(cid:48). We used the z(cid:48) ab filter (z_G0328, >848 µm) and a dither pattern with 4 positions offset by 15(cid:48)(cid:48), which is twice the size of the standard dither pattern. The total integration 6 BARON ET AL. Figure 2. On the left, z(cid:48) ab − J vs spectral type for L to T dwarfs from Hawley et al. (2002) for the L dwarfs and Albert et al. (2011) for the T dwarfs. The L to T dwarfs are characterized by red z(cid:48) ab − J colors. The red dot represents Gu Psc b, the planetary mass object discovered by Naud et al. (2014), representative of the kind of objects we are seeking in this work. On the right, [3.6]-[4.5] for L to T dwarf from Dupuy & Liu (2012). We see that late T dwarfs can be identified both by their red [3.6]-[4.5] > 1.5 and z(cid:48) time per target varies between 311 s and 476 s. The higher integration time is for targets with a declination in the range −35 to −30, to accommodate the higher airmass and maintain a good SNR. With a seeing between 0.55(cid:48)(cid:48)and 0.65(cid:48)(cid:48), this en- sures a SNR of 3 for all our z(cid:48) ab-band observations with a limiting magnitude of z(cid:48) Our complete sample was observed with Spitzer/IRAC. Nine of our targets had previously been observed with IRAC with an exposition time that meets our requirements. The others targets were observed between 2015 and 2016 (Spitzer proposal 11092) in both IRAC [3.6] and [4.5], with a total in- tegration time of 2160 s in each band (per-visit total of 5221 s with overheads). More precisely, we used 30 s individual ex- posure time, two exposures per dither position per band, and a 36-exposure reuleaux dither pattern. ab − J > 4 colors. The Spitzer/IRAC pipeline reduced images were further processed with custom IDL routines. First, the different im- ages of a given target were oversampled on a 0.5(cid:48)(cid:48) pixel grid and median-combined using the pipeline-provided astrome- try and polynomial distortion. Then, to preserve the PSF morphology orientation in the image, in view of the PSF sub- traction routines to be applied, we registered all images to a common PSF rotation angle. Further data reduction involved the subtraction of the stel- lar point spread function (PSF) to reveal embedded and close- in sources. Since the Spitzer observations were uniform, we used the Reference Differential Imaging technique to sub- tract the PSF from a reference library. The strategy is similar to the re-analysis of Hubble imaging data through the ALICE project (Soummer et al. 2016; Choquet et al. 2015), and to previous analysis of archival Spitzer data (Janson et al. 2015; Durkan et al. 2016). The library of reference PSF was created out of the newly obtained data, using the PSF of the observed stars. Satu- rated stars, very crowded fields, and low-contrast (< 1) vi- sual binaries were removed from the library, resulting in a total of 111 PSFs out of the 168 targets observed. Each im- age was registered on a common center based on the fit of a two-dimensional Moffat function. It was then normalized in brightness from the flux measured in an aperture of a radius of seven pixels centered on the PSF core. Point sources were ab=24. The WIRCam images were reduced using the method de- scribed in Albert et al. (2011). First, they were preprocessed by CFHT using their 'I'iwi pipeline version 2.0. Next, a low-pass filter was created by median binning the image by 4x4 pixels, applying a 5x5-pixel median filter, and then re- sampling at the original image size. This low-pass filter was subtracted from the image to preserve only high spatial fre- quencies. After this, the different images were stacked using the Bertin software suite. First, SExtractor (Bertin & Arnouts 1996) builds a catalog of objects in each image. This catalog is in turn read by Scamp (Bertin 2010a), which also com- putes the astrometric and photometric solutions by anchor- ing on the J-band data of the 2MASS catalog. Swarp (Bertin 2010b) then stacks the images together. Data from MegaCam were first processed by CFHT's Elixir pipeline. Then, the astrometric solution from each of the 40 CCDs was found by anchoring the field on the po- sitions from the USNO-B1 catalog. A high-pass filter was applied, as before, by subtracting an image created by me- dian binning the image by 4x4 pixels, applying a 7x7-pixel median filter, and then re-sampling at the original image size. Then the images from the different CCDs were combined to form an image of size matching that of the WIRCam im- ages, as the field of view of MegaCam is much wider than WIRCam's. The different images of a given target obtained on a given night were then median-combined to obtain the final reduced z(cid:48) ab image. 2.3.3. Spitzer/IRAC observations Gu Psc bL0L5T0T5Spectral type2.53.03.54.0zAB' − J (mag)L0L5T0T5Spectral type0.00.51.01.52.0[3.6]-[4.5] (mag) WEIRD 7 Figure 3. BT-Settl spectral energy distribution of young objects (logg =4 and solar metallicity) with effective temperatures of 500 K, 800 K, 1000 K, and 1200 K. The transmission functions of the four bandpasses used for our observations (z(cid:48) ab, J, [3.6], and [4.5]) are overlaid. These bandpasses provide distinctive red colors while maintaining a high flux level across the temperature range. identified as 3σ outliers from the noise (calculated with a ro- bust sigma estimator) after an initial PSF subtraction from the median of the reference library, excluding the given image. They were subsequently masked out in the original image. We used classical RDI to subtract stellar PSFs in each Spizter image for both filters. Because of the very large num- ber of point sources in our deep data, advanced techniques such as LOCI or PCA (Lafrenière et al. 2007a; Soummer et al. 2012) suffered from too many pixels that were masked out, reducing the effective number of reference images. They tended to oversubtract the target PSF and other point sources in the field. We therefore opted for a classical median sub- traction, a trade-off between the quality of the PSF subtrac- tion and source preservation. Following this strategy, the pro- cessed image under consideration was excluded from the li- brary of references, the median of which was then taken as the reference PSF for subtraction of this image. The position of the star was estimated by fitting a Moffat function and the reference PSF was shifted to this position. The reference PSF was normalized to the target brightness within the same aper- ture and subtracted from the image. This three-step process was repeated for any low-contrast (< 1) visual companion of our target present in the field. For tight binaries or triple systems, the subtraction of all PSFs was done at once by it- erating over the position and flux of each component in order to minimize the residuals in a box of width of 30 pixels. Sat- urated stars were processed like binaries to optimize the star registration and flux normalization. They still suffered from poorly subtracted wings and bright vertical stripes escaping from the core. Therefore, a new library was built from resid- ual images of similarly saturated stars, following the same cleaning processes as for the original library. The median of this new library was used to subtract these residuals. The images were finally high-pass filtered by subtracting a median filter of width 15 pixels. 2.3.4. Archival Spitzer/MIPS 24 µm data A search through the Spitzer archive revealed that 141 of our 177 targets were observed with MIPS at 24 µm as part of surveys to find infrared excess indicative of debris disk. A flux measurement (or upper limit) at such a longer wave- length can be useful to better constrain the SED of our can- didate objects identified. Thus for those 141 targets we re- trieved the MIPS data and built our own combined image using the Enhanced BCD images (EBCD), as they have a superior flat fielding than the BCD image. A high-pass fil- Teff=500KTeff=800KTeff=1000KTeff=1200K12345l (mm)05101520lflF2's J filterWIRCam's J filterGMOS-S's z' filterMegaCam's z' filterIRAC [3.6] filterIRAC [4.5] filter 8 BARON ET AL. ter was then applied by subtracting a double-pass median- filtered version of the image, respectively of widths 32 and 12. We obtained the limiting flux in Jy/sr by doing aperture photometry at several random positions over the whole im- age and then evaluating the robust standard deviation of the resulting flux distribution. We converted this limiting flux in Jy/sr to magnitude and, over all images, obtained a median limiting magnitude of 12.5 mag. 2.4. Photometric calibration Our GEMINI observations were all acquired with a spec- ification for observing conditions of up to 70% cloud cover, or patchy clouds. Under those conditions, a variation of up to 0.3 mag can be expected. We assessed if significant vari- ations were present or not from the data themselves. For a sequence of observations of a given field, we calculated the standard deviation of the flux variations of the 20 brightest stars, as compared to a reference image from the sequence. If this variation was higher than 3%, then we considered that the images of that target were not taken in photometric con- ditions ('patchy clouds' in table 4 and table 5); otherwise we considered that the images were taken under photometric conditions ('phot' in table 4 and table 5). All of our CFHT images were taken in good conditions, with seeing around 0.6 for MegaCam and 1.1 for WIRCam, but we still checked the flux variations between images for a given target on a given night to make sure that the observa- tions were acquired in photometric conditions. Our final, stacked images in z(cid:48) ab and J were calibrated in flux by comparison with the Sloan Digital Sky Survey cat- alog (SDSS DR9; Ahn et al. 2012) or the 2MASS All-Sky Catalog of Point Sources (Cutri et al. 2003), respectively. If SDSS data were not available for a given field, we used either PanSTARRS (Chambers & Team 2018) zp1 data (available for 55 of our targets with dec >−30) or SkyMapper (Wolf et al. 2018) z(cid:48) data (available for 79 of our targets). The PanSTARRS filters (gp1, rp1, ip1, zp1, yp1, wp1) are not the same as the SDSS filters, so we used the Tonry et al. (2012) color corrections to convert the Pan-STARSS magnitudes to SDSS magnitudes. For the J band, if too few stars in our im- ages were in the 2MASS PSC, we used deeper data from the VISTA Hemisphere Survey (VHS; McMahon et al. 2013) or archival observations from the Observatoire du Mont Mégan- tic obtained using the Spectrographe infrarouge de Montréal (SIMON) (Albert 2006). For a given image, the magnitude that produces one count per second on the detector, or the zero point, was calcu- lated for each individual point source in common between our image and the catalog, based on the difference between the magnitude extracted from our image and the magnitude taken from the catalog. Then, the zero point of the image was taken to be the median of the individual zero points, and the error was computed by taking the standard deviation of those individual zero points divided by the square root of the number of sources. When no catalog data were available for a given image, or when less than 5 objects with a magnitude measurement were available in the field of view, we used as a zero point the me- dian zero point for the given observing condition ('phot' or 'patchy clouds'), instrument and filter of the image. This oc- curred for 16 of our J-band images and 48 of our z(cid:48) ab-band In the J band, we obtained a zero point of 22.4 images. ± 0.7 and 22.1 ± 0.9 with Gemini/F2, and 22.5 ± 0.8 and 22.7 ± 0.6 with CFHT/WIRCam, for photometric and non- photometric conditions respectively. In the z-band, we cal- culated a median zero point of 24.5 ± 0.3 and 24.6 ± 0.5 with CFHT/MegaCam and 29.7 ± 2 and 28.9 ± 2.3 with Gemini/GMOSS, for photometric and non-photometric con- ditions respectively. 2.5. Follow-up observations Our search for planetary mass object revealed a number of candidates (see Section 3.2.3) that motivated us to obtain follow-up observations. An astrometric follow-up was carried out between 2016 and 2017 in the J-band, with either CFHT/WIRCam or Gemini-South/Flamingos2. Only J-band images were ob- tained as it is in this band that the SNR of the candidate is highest. We used the same observation parameters as for the first epoch observations. We obtained proper motions follow- up in the J band for 4 candidate companions. 3. RESULTS The ground-based observations described earlier were de- signed to reach a limiting magnitude of z = 24 mag at 3σ and J=21 mag at 7σ. In practice, we achieved a median [AB] limiting magnitude in the z-band of 23.4±1.2 mag with CFHT/MegaCam and 23.7±1.2 mag with Gemini/GMOS-S, at 3σ. In the J-band, we achived a 7σ median Vega limit- ing magnitude of 21.2± 0.5 mag with CFHT/WIRCam and 21.0± 0.8 mag with Gemini/F2. For the Spitzer/IRAC ob- servations, we reached a median magnitude limit of 18.5± 0.9 mag at [3.6] at 5σ and 18.5± 0.8 mag at [4.5] at 3σ. 3.1. Detection Limits The sensitivities to companions, in terms of limiting mag- nitudes, were evaluated for each J-band stacked image and [4.5] image as a function of the radial distance from the tar- get star. For each radius from the central star, aperture pho- tometry was performed by obtaining the flux inside 100 aper- tures of radius of 1 FWHM and a sky annulus between 4 to 6 FWHM. The limiting flux at each radius is the standard devi- ation of these 100 fluxes and it was then converted into mag- nitudes to get these 7σ limiting magnitudes. Theses results WEIRD 9 are presented in Figure 4 for the J-band images and in Fig- ure 5 for the [4.5] images. They show that the limiting magni- tudes initially grow with increasing distance from the central star, and then reach a plateau. The black region of the plots contain 50% of the detection limit curves while 80% of the curves fall inside the grey region. For the J-band images, the plateau is reached at ∼30(cid:48)(cid:48) at a magnitude of J ∼ 21.5 mag for 50% of our target stars (the black region). The curves are truncated at about 180(cid:48)(cid:48), which corresponds to the limit of the field of view of the Flamingos-2 images. The plateau is reached at a projected physical separation of 1000 AU for an average star of our sample. For our [4.5] images, the plateau at magnitude ∼18.5 is reached at a radius of ∼50(cid:48)(cid:48)for 50% of the stars of the sample (the black region). We used the same cut-off as the J-band images. Tables 6 and 7 present, respec- tively for the J-band and [4.5] images, the 7σ detection limits for each target over the plateau along with the minimum and maximum separations (in arcsec and AU) where these limits are valid (defined as the range for which the detection limit is at most 1 magnitude brighter than the plateau value given, to accommodate for small fluctuations with separations). The limiting magnitudes can be converted to limiting masses using evolutionary models at the ages of our targets (which range between 10 and 150 Myr). We used the COND models from Baraffe et al. (2003) to infer the masses. These models assume a hot start, which as described by Bowler (2016), corresponds to idealized initial conditions and an ar- bitrarily large initial radius. This model is thus optimistic as it represents more luminous planets than cold start models. The mass limit reached over the sensitivity plateau for each target is indicated in Table 6 and 7. 3.2. Candidate Search We searched for and identified candidates in our imaging based on their z(cid:48) ab − J and [3.6] − [4.5] colors. We started by identifying all point sources in the J-band images using the IDL find procedure (from Astrolib) and then fitted a 2D Gaussian function to each of them to get a more pre- cise position. At this step, sources with an elongated PSF were rejected, as a first attempt to exclude extra galactic con- taminants. We also rejected sources too close to the edge of the field (for F2 or GMOS-S) and sources that were sat- urated in either band. We used coordinates measured in our J-band images to identify sources in the z(cid:48) ab-band images. In both bands, we used aperture photometry with a radius of 1 FWHM and a sky sampling annulus extending between 2 and 3 FHWM to retrieve the instrumental flux of each source. We kept only point sources detected at 7σ in J, 5σ in [4.5] and 3σ in [3.6]. At a distance of > 20 pc, which is the case for 90% of the stars in our sample, a radius of 5000 AU fits in the field of view of the Spitzer/IRAC images. For that reason, we searched for candidates only inside a projected separation of 5000 AU from the target stars. We found the center of the target star by fitting a 2D Gaus- sian to the PSF, for stars that were not saturated. However, most of our targets were saturated in our J-band images. Thus, we used the Gaia DR1 DR1 catalog (Gaia Collabora- tion et al. 2016) to find an approximated position of the star. We then used the known proper motion of the star to com- pute the position at the time the image was obtained. If Gaia data were not available, we fitted a 2D Gaussian to the PSF, where all of the saturated pixels were given the maximum value possible for a pixel. For the IRAC images, the center of the stars was obtained during the PSF removal process. 3.2.1. Colors For the ages of our target stars, 10 -- 150 Myr, the tran- sition between brown dwarfs and planets happens between L1/L2 and L5/L6 based on AMES.Cond models (Baraffe et al. 2003). As mentioned above (and see Figure 2), early- ab − J color (cid:38)2.5 mag. Considering type L dwarfs have a z(cid:48) our errors on magnitudes and zero points, we selected only sources with z(cid:48) ab − J > 2.2. Per the above discussion, this same color cut is also sensitive to T and Y dwarfs, which can be identified either through detection in both bands or as z(cid:48) ab dropouts (in the cases without detection in z(cid:48) ab, we get only a lower limit on the z(cid:48) ab −J color).Thus at this stage, we kept all sources with z(cid:48) ab dropouts. As a second step, we removed any source that has a coun- terpart in the Gaia DR1 DR1 catalog (Gaia Collaboration et al. 2016). Gaia can detect objects with magnitude as low as G=20. Based on the relationship between G − J and the spectral type of L dwarfs from Smart et al. (2017), using the expected J magnitude of L dwarfs from Faherty et al. (2016), and assuming a distance of 42 pc (the median distance of our sample), the cut in Gaia magnitudes rejects objects earlier than ∼L2. ab −J > 2.2 mag, including all the z(cid:48) Then, we compared the z(cid:48) ab − J colors and [3.6] − [4.5] colors of our candidates to typical colors of ultracool field dwarfs (Dupuy & Liu 2012), see Figure 7. This figure shows all point sources in a radius of 5000 AU in the J-band im- age for an average target of the sample for which there was no candidate detected. The solid black line represents the expected colors for L to T dwarfs according to Dupuy & Liu (2012). We kept as candidates only the detections with [3.6] − [4.5] ∼ 0.1 to 2 mag, as this is the expected interval for T dwarf's colors. We also kept as candidate source with MJ < 16 mag and [3.6] − [4.5] < 1 mag or MJ > 16 mag and [3.6] − [4.5] > 1 mag. Figure 6, right, presents the flowchart of the candidate selection for the candidates detected in the J band. In some cases, a source was detected at 5σ in our IRAC ab imaging, data but we found no counterpart in our J or z(cid:48) 10 BARON ET AL. Figure 4. Detection limits for all of our stacked J-band images observed with Flamingos-2 at Gemini-South or WIRCam at CFHT. The left panel shows limiting apparent magnitudes as a function of the projected separation from the target star in arcseconds. The right panel shows the corresponding absolute magnitudes at the distance of the star as a function of projected separation from the star in AU (with a cutoff at 5000 AU). 50% of the detection limit curves fall inside the black region while the grey region contains 80% of the curves. Figure 5. Same as 4 for the Spitzer/IRAC observations. respectively at 7 and 3σ. Unambiguous IRAC-only detec- tion of planets is possible only if [3.6] − [4.5] > 2, which corresponds to our detection limits of ∼21 in the J band, or MJ ∼ 18 (T8.5) at 50 pc according to AMES.Cond mod- els. However, most of our IRAC-only detection had 0.5 < [3.6] − [4.5] < 2. As the color in those bands for young 2 Mjup objects is rather uncertain, we decided to follow- up these sources anyway. Thus from the IRAC-only detec- tion, we selected only sources with [3.6] − [4.5] > 0.5 and no Gaia detection. In addition, as the absolute magnitudes of young planetary mass objects analog to T dwarfs are not well known, we kept only sources with a [4.5] absolute magnitude within 0.75 mag from the typical values of field T dwarfs, see Figure 8. This method has uncovered 79 candidates with the expected colors of T dwarfs. Figure 6, on the right, presents the flowchart of the candidate selection for the candidates not detected in the J band. The color criteria above yielded typically a few candidates per field. However, most were easily discarded by either looking at the stacked images or the individual frames: some had an elongated PSF that escaped our automatic cut, some fell out of the detector in one or more frames of the dither pattern biasing their photometry, some were due to a persis- tence signal from a bright star that was on the same part of the detector in a previous frame (for the WIRCam images), and some fell over the spider diffraction spikes of the host star. After these initial verifications, our search yielded 4 candi- dates with J-band detection and 48 candidates with IRAC- only detections. 3.2.2. Cross-match with the 2MASS calatog The detection method described earlier is not sensitive to companions with spectral type earlier than early L dwarf. In- stead, the latest M to early L-type dwarf companions can be identified through a search for common proper motion based on a comparison of our J-band images with 2MASS images, given the ∼15 years baseline between them. We per- formed such a proper motion comparison for all sources with J<16.5 mag. 050100150Radius from center star (arcsec)18192021222324J Apparent Mag010002000300040005000Distance from center star (AU)15161718192021J Absolute Mag50100150Radius from center star (arcsec)14161820[4.5] Apparent Mag010002000300040005000Distance from center star (AU)12141618[4.5] Absolute Mag WEIRD 11 Figure 6. Flowchart of the candidate selection for candidates detected in the J-band on the left and the IRAC-only candidates on the right. DetectedinJand[4.5]Detectedat[4.5]andnotdetectedinJDetectedinGAIA?Rejectedz(cid:48)−J>2.2mag0.1<[3.6]-[4.5]<2magMJ>16magand[3.6]-[4.5]<1mag4candidatesYesDetectedinGaia?[3.6]-[4.5]>0.5magDetectedinMIPS17candidatesNoYesYesNoYesNoNoYesYesNoYesYesNoNoYes1 12 BARON ET AL. companion of TWA30 -- an M5 dwarf member of the TW Hydrae association -- at a separation of 3400 AU and which was discovered previously by Looper et al. (2010). 3.2.3. Follow-up of candidates Figure 7. Color-color diagram for HIP 26453, a known member of Columba. The dots represent all sources detected in our J-band imaging, and without detection in Gaia, within a radius of 5000 AU from the target star. The solid line shows the expected color se- quence for spectral types L to T from Dupuy & Liu (2012). The box represents the expected colors for early Y dwarfs. No candi- dates were detected in this field. Figure 8. [3.6]−[4.5] color of sources detected in our Spizter imag- ing of HIP 11152 versus their [4.5] absolute magnitude at the dis- tance of the target star. The solid red line corresponds to the colors of M6 to T9 dwarf from Dupuy & Liu (2012). The dotted lines on either sides represented a spread of 0.75 magnitude. The dots are all the point sources presents in a sphere of 5000 AU around the central star for which there is no detection in the optical. One point source has colors consistent with a late T dwarf at the right abso- lute magnitude. This point source is not detected in the z(cid:48) ab nor J images. While it is expected for a planetary mass companion to be undetected in z(cid:48) ab, it should have been detected in J images, given our detection limits. It is thus likely that the candidate is in fact an extragalactic contaminant. This search identified one candidate with a proper motion consistent with a target star. It is TWA30B, an M4V dwarf Figure 9. Photometric data for one candidate that has a large [3.6]− [4.5] color but no detection in z(cid:48) ab and J. The data are compared to the model spectrum of an object with a Teff = 1100 K, logg = 4 and z = 0 from BT-Settl (purple) and to the spectrum of a featureless AGN with a redshift of 0.7 and a DL=4300 Mpc (magenta, from Kirkpatrick et al. 2012a). We see that the detection at 24 µm makes it very easy to untangle between a mid-T dwarf and a AGN. The follow-up of our candidate companions includes 3 dif- ferent types of observations. First of all, IRAC-only detec- tions were studied in greater details by using MIPS data. A photometric follow-up was obtained to try to identify puz- zling objects with very red [3.6] − [4.5] colors and no detec- tion the z(cid:48) ab and J bands. Lastly, a proper motion follow-up was obtained for all candidates detected in J that survived the color cuts and verifications. Our search for candidates in the Spitzer/IRAC images yielded 48 candidates with [3.6] − [4.5] > 0.5 and no detec- tion in z(cid:48) or J. Figure 8 shows all the point sources detected in [3.6] and [4.5] in a given field and for which no visible counterpart was found (from the Gaia DR1 catalog). Faint, red objects like this certainly constitute interesting planetary mass candidates, as indeed it is expected for such objects to be z(cid:48) dropouts. Yet given our limits it is unexpected for them to be unseen in J. Other astrophysical sources that may have similar photometric properties include galaxies and ac- tive galactic nuclei (AGNs). Figure 9 shows the expected SED of a low-mass object with an effective temperature of 1100 K compared to the SED of a featureless AGN. As the Figure illustrates, it is difficult to untangle AGNs from planetary candidates using [3.6] and [4.5] photometry alone, but photometry at 24 µm is a very L0T9Y1234z'−J (mag)-0.50.00.51.01.52.0[3.6]-[4.5] (mag)6810121416M4.5 (mag)-0.50.00.51.01.52.02.5[3.6]-[4.5] (mag)110l (mm)-10-8-6-4-20log Fn (Jy)Featureless AGN, redshift=0.7, DL=4300MpcBT-Settl, T=1100K, log(g)=4, z=0 WEIRD 13 Figure 10. Colors of our 17 Spitzer/IRAC-only candidates remaining after the MIPS detection cut (triangles, upper limits in J-band). [3.6] − [4.5] colors versus absolute J magnitude are shown on the upper left while [3.6] − [4.5] colors versus absolute [4.5] magnitudes are displayed on the upper right. The the lower left shows absolute [4.5] magnitudes vs absolute J magnitudes. Colors for M6 to T9 dwarfs from Dupuy & Liu (2012) are shown with a black line. The red curves represent the Ames.Cond models (Baraffe et al. 2003) at 10, 20, 120 and 5000 Myr, using respectively the solid, dotted, dashed and dash-dotted line. Also shown are models from Beichman et al. (2014) in cyan, Mordasini et al. (2012) in yellow and Ames.Dusty in green. Photometric data for 3 young T dwarfs are also shown by an orange star for Gu Psc b (Naud et al. 2014), a purple star for SDSS1110+0116 (Gagné et al. 2015a) and a red orange star for 2MASS1324+6358 (Gagné et al. 2018a). While the candidate companions have similar [3.6]-[4.5] colors versus [4.5] as the young T dwarfs, they are too faint in the J-band to be considered planetary objects. good discriminator. We used the MIPS 24 µm images men- tioned above, reaching a limiting magnitude of 12.5 at 1σ in that band for most targets, to see if our candidates were detected at that wavelength, which would be incompatible with a planetary mass object. This enabled us to reject 31 of our remaining IRAC-only candidates and to identify them as extra-galactic contaminants. We checked archives to see if those MIPS detection are associated with X-ray or radio emission, but none of them are already known as AGN. After this cut, 17 IRAC-only candidates remain. Fig- ure 10 shows the colors and magnitudes of the candidates compared to different models as well as to photometric data from known young T dwarfs. Of those 17 candidates, 4 were observed by MIPS but not detected. These candidates have [3.6] − [4.5] = 0.7 to 0.9 mag and [4.5] magnitudes between between 15.8 mag and 17.5 mag. Using only their IRAC color and assuming that they are T dwarfs and that the BT- Settl/Ames.Cond model are valid, one would expect them to have MJ ∼ 15 mag, which would have been detected by our survey. As these candidates show no detection in our J-band imaging, we rejected those 4 candidates. The last 13 can- didates were not observed by MIPS. Those candidates have 1214161820MJ (mag)-0.50.00.51.01.52.02.53.0[3.6]-[4.5] (mag)111213141516M4.5 (mag)01234111213141516M4.5 (mag)1314151617181920MJ (mag)9101112131415M4.5 (mag)AMES.CondAMES.DustyBeichmanMorley2012Gu Pcs bSDSS1110+01162MASS1324+6358Candidates 14 BARON ET AL. [3.6] − [4.5] colors between 0.6 mag and 1.3 mag and [4.5] magnitudes between 15.2 and 16.9 mag. Using the same thought process as for the candidates not seen in MIPS, we see that those candidates also should have been seen in the J band, but they were not detected. We thus reject the last 13 candidates such that no IRAC-only candidate remain. How- ever, we decided to list those 17 rejected candidates in the interest of completeness, as we cannot identify the nature of the candidates at this stage, and because models might not reproduce accurately the colors of young late T to early Y dwarfs. Table 8 lists them all, with their RA, DEC, associ- ated host star, limiting magnitude in z(cid:48) and J, apparent mag- nitude in [3.6] and [4.5], separation in AU from the host star, and distance of the host star in pc. These unknown objects are possibly Ultra Luminous Galaxies (ULIRGS). ULIRGS are identified by their red [3.6]-[4.5] > 0.5 colors meaning that they share colors with T dwarfs. Daddi et al. (2007) have shown that ULIRGS from the GOODS sample, with 0.7<z<1.3 have a space density of 2x10−5Mpc−3. At a lu- minosity distance corresponding to a redshift of z=1, about 3 ULIRGS should have been found per Spitzer/IRAC field. As ULIRGS have Fν ∼ 10µJy for z ∼ 1 to 2 (Kirkpatrick et al. 2012b), they are expected to be detected in our images. A proper motion follow-up was obtained for all 4 candi- dates identified through their z(cid:48) − J and [3.6] − [4.5] colors. It was carried out between 2016 and 2017 both at CFHT and at Gemini-South. Table 9 lists the candidates with their RA, DEC, host stars, Mz(cid:48), MJ, M3.6, M4.5, separation in AU, pmra, pmdec and the number of sigma at which the proper motion of the candidate differ from the host star's proper motion. The candidates are rejected at 3σ or higher. 4. ANALYSIS AND DISCUSSION 4.1. Sensitivity and completeness For each image of our survey, the sensitivity to planets of a given semi-major axis and mass can be determined using the limiting magnitude reached as a function of the projected separations from the star and the corresponding fraction of pixels where a companion could have been detected. In com- puting these detection completeness maps for all stars in our sample, we adopted an approach similar to that of Nielsen et al. (2008) and Naud et al. (2017), relying on a Monte Carlo simulation. First, for a given image and a given separation from the star, the fraction of clean pixels, i.e., pixels where a compan- ion could have been detected if indeed it was present, was simply determined by counting pixels at that separation that were not flagged as bad, not saturated, and not affected by the presence of a star. Figures 11 and 12 show this fraction as a function of separation from the star for the J-band images and the [4.5]-band images, respectively. In most cases, at 10(cid:48)(cid:48) the fraction reaches 0.9 for the J-band images and 0.98 for the Figure 11. Fraction of clean pixels where a companion could be detected as a function of the separation from the target star in the J-band images. 50% of the stars have a fraction of pixel that is included in the black area while the grey area represents 80% of the stars. For most stars, the fraction of clean pixels reaches 90% at 10(cid:48)(cid:48). Figure 12. Same as 11 for Spitzer/IRAC observations at [4.5]. For most stars, the fraction of clean pixels reaches 98% at 10(cid:48)(cid:48). [4.5]-band images. In a few cases, the target star is in the galactic plane, making the detection of a companion harder and the fraction lower. Huge variations in fpixel at smaller separations come from the different magnitude of the central stars, and the associated different areas affected by satura- tion. Some stars of the sample are very saturated and thus fpixel is very low at small separation while the M dwarfs of our sample are not saturated and thus a higher fpixel is reached at smaller separations. In general, the fraction of pixel for an individual target can be fitted by a logistic function with the shape of 1/(e−a0(x−a1) + ea2), where a0 is the steepness of the curve, a1 is the x-values of the mid point and a2 is typically close to 0. Table 9 and Table 10 show the values of the 3 parameters for each target of the sample for the J-band and 110100Radius from center star (arcsec)0.00.20.40.60.81.0fpixel10100Radius from center star (arcsec)0.50.60.70.80.91.0fpixel WEIRD 15 [4.5] images respectively (a0 varies from -10 to 40, a1 goes from 0 to 14, and a2 is close to 0). Next, we defined a grid of masses and semi-major axes, with the masses equally spaced in logarithmic scale between 0.5 and 15 MJup and the semi-major axes equally spaced in logarithmic scale between 100 and 5000 AU. For each point of the grid, we simulated 104 planets. Each planet has an ec- centricity taken randomly from the eccentricity distribution reported in Kipping (2013), which is taken from the eccen- tricity from RV planets. Then we used the method of Bran- deker et al. (2006) and Brandt et al. (2014) to find the in- stantaneous projected separation of each planet, given their eccentricity, semi-major axis, and some random inclination and time of observation. The projected separation in AU was finally converted to a projected angular separation in arcsec by dividing by the star distance, which is sampled uniformly within its interval of uncertainty. For each grid point, we converted the mass into a J-band absolute magnitude using the AMES.Cond evolution models (Baraffe et al. 2003) and the ages of the targets from Table 1 and Table 3. We randomly sampled the age of each gener- ated planet uniformly between the uncertainties given for the appropriate moving group (see Table 1). We then used the known distance of the star to convert the planets' absolute magnitudes to apparent magnitudes, and compared these to the detection limits found earlier to assess the detectability of each planet. If a planet was brighter than the detection limit, we used the fraction of clean pixels found earlier at that sep- aration as the detection probability; otherwise the planet was assigned a detection probability of zero. This was repeated for each simulated planet, and the results were averaged to find the probability of detection at each point of the grid. This procedure was repeated for all targets of the sample. The sensitivity of the whole survey was calculated by tak- ing the median of all the detection probability maps. Two completeness maps were made this way, one for the J-band images (Figure 13, left) and one for the [4.5]-band images (Figure 13, right). The ground-based survey is mostly sensi- tive to objects with masses higher than 2 MJup with a semi- major axis of more than 1000 AU while the Spitzer survey is sensitive to planets slightly less massive (down to 1 MJup) at larger separations. The completeness maps for each star of the sample and for both J and [4.5] bands were combined to build the overall completeness map of the survey. For each star at each point of the grid, the highest probability was taken between the completeness map of the J-band images and the [4.5] images. The two-band combined completeness maps were then aver- aged over all stars to obtain the overall survey completeness maps, see Figure 14. Figure 15 shows the mean detection probability as a function of semi-major axis for planetary ob- jects with masses of 1 MJup, 2 MJup, 3 MJup and 13 MJup, taken from the overall completeness map of the survey. The max- imal probabilities of detection are respectively 64%, 95%, 98%, and 99%. Our survey is mostly sensitive to planets with masses of 2 MJup and above, as the detection probability falls very rapidly between 2 and 1 MJup. Our results probe an area of the semi-major axis -- mass di- agram that has not been studied before. Figure 14 shows our completeness map compared to the regions probed by the fol- lowing other studies: the PSYM-WIDE survey (Naud et al. 2017), aiming at discovering planetary mass objects on wide orbits around K5-L5 dwarfs, the PALMS survey (Bowler 2016), a deep coronagraphic study of 78 single young nearby (<40 pc) M dwarfs, the GPDS survey (Lafrenière et al. 2007b), a survey of young stars searching for giant planets on large orbits, the NaCo Survey of Young Nearby Dusty Stars (Rameau et al. 2013), which targeted 59 young nearby AFGK stars, the NaCo-LP survey (Chauvin et al. 2015), which fo- cused on 86 young, bright, and primarily FGK stars, the IDPS-AF survey (Vigan et al. 2012), which observed 42 AF stars, the MMT L(cid:48) and M-band Survey of 54 nearby FGK stars (Heinze et al. 2010), the Gemini NICI Planet-finding Campaign (Biller et al. 2013), which targeted 230 young stars of all spectral types, MASSIVE (Lannier et al. 2016), which targeted 58 young and nearby M-type dwarfs, the IDPS sur- vey Galicher et al. (2016), which combine results for 292 young nearby stars and Durkan et al. (2016) who studied 121 nearby stars observed with SPIT ZER/IRAC. On the whole, this survey is a good complement to AO imaging surveys, being mostly sensitive at separations of several hundreds of AU but insensitive at semi-major axes of less than ∼ 150 AU, where AO imaging surveys are most sensitive. 4.2. Constraints on additional companions in systems with known directly imaged companions At least one planetary mass or brown dwarf companion was previously found around 6 stars in our sample; most of these companions were found using high-contrast AO imag- ing. Our search, being sensitive to much wider separations and reaching lower masses, adds valuable constraints on the presence of additional companions in these systems. We pro- vide in Figure 16 the individual completeness maps from our survey for these six systems. The companion Pz Tel B, a 36± 6 MJup brown dwarf orbit- ing at 16.4±1 AU from a pre-main sequence G9 star mem- ber of the β-Pictoris association, was found by Biller et al. (2010) using VLT/NACO. We put constraints on the pres- ence of companions at larger orbits (see Figure 16, top left). At a confidence level of more than 90%, we can reject a com- panion with masses as low as 1 -- 2 MJup at 2000 -- 5000 AU. The companion 2M1207 b, a 4±1 MJup object (Chauvin et al. 2004) orbiting at 46 +37 −15 AU (Blunt et al. 2017) around the young brown dwarf TWA27, member of the TW Hydrae 16 BARON ET AL. Figure 13. Completeness map for the J-band images on the left and for the [4.5] images on the right. They show the probability of detecting a planet with a mass between 1 and 13 MJup as a function of the separation from the host star. Curves for 10%, 50% and 90% are shown. association at 52 pc, was discovered using VTL/NACO. Our survey put strong constraints on the presence of > 10 MJup objects in the system, as they should have been detected at separations from 100 to 5000 AU. Moreover, at a distance of 1000 AU, the detection probability of 1 MJup object is about 80%. Our survey covers quite well the regime of separations > 1000 AU and masses > 1 MJup (see Figure 16, top right). No companion was detected by our survey. Chauvin et al. (2005) found a 13.5 ± 0.5 MJup object at 250 AU of AB Pic, a K2V star member of the Tucana- Horologium association, by using VLT/NACO. Figure 16, middle left, presents the completeness reached by our survey. We put strong constraints on the presence of companions of 2 MJup or more at separations higher than 1000 AU. −3 and 9+4 Marois et al. (2008, 2010) used AO observations with −2, 10+3 Keck/NIRC2 and Gemini/NIRI to find 4 planets of 7+4 −3, −4 MJup at respectively ∼68, 43, 27 and 17 AU 10+3 from HR 8799(Wertz et al. 2017), an A5V star member of the Columba association. We probed a region in mass that is equivalent to the planets already known, but at much larger semi-major axes. We put good constraints on the presence of companion with ≥ 4 MJup and semi-major axis greater than 1500 AU. Lagrange et al. (2009) found a 12.7±0.3 MJup (Morzin- ski et al. 2015) planet at 9.2+1.5 0.4 AU (Millar-Blanchaer et al. 2015) orbiting β Pictoris, an A6V star member of the β Pic- toris association, using high-contrast VLT/NACO observa- tions. Our observations put strong constraints on the exis- tence of objects of 1 MJup or more at semi-major axes of >1000 AU. 2 (Rajan et al. 2017) object orbiting 51 Eri at ∼14 AU, an F0IV star, was found by Macintosh et al. (2015) using Gemini/GPI. 51 Eri is part of a triple system, bound A 1 -- 2 MJup 2 This mass was inferred from hot start model from (Marley et al. 2007). It is also possible that the mass is anywhere between 2 -- 12 MJup according to the cold start model from Fortney et al. (2008). to and separated by ∼2000 AU from GJ3305AB, an M+M binary of unresolved spectral type M0 (Montet et al. 2015). Our survey put strong constraints on the presence of compan- ions of mass > 1 MJup at semi-majors axes between 100 and 5000 AU. 4.3. Planet frequency Based on the null result of our survey, and our complete- ness limits calculated in section 4.2, we evaluated an up- per limit to the frequency of occurrence of planets at large semi-major axis (1000 -- 5000 AU), following the method de- veloped by Lafrenière et al. (2007b). If we have N=177 stars enumerated from j=1 to N, and we survey an interval of mass going from 1 to 13 MJup and an in- terval of semi-major axis of 1000 to 5000 AU, then we define f to be the fraction of stars with at least one companion in the intervals and p j the probability of detecting such a com- panion. This probability is computed from the completeness map calculated previously by taking the mean of the proba- bility at each point of the 100×100 grid. Since the grid is uniform in logarithmic space, this amounts to assuming that the semi-major axis and the mass are distributed uniformly in log. The detections in the survey are characterized by the set {d j}, and in our case, since the survey gave a null result (all known companions around our targets were too close-in to be seen in our data), d j=0 for all j. The probability of ob- serving the set {d j} in our survey is given by the following binomial likelihood, N(cid:89) L({d j} f ) = (1 − f p j)1−d j( f p j)d j . (1) j=0 Then according to Bayes theorem, the posterior distribu- tion for f , in light of our results, is given by, (cid:82) 1 L({d j} f )p( f ) 0 L({d j} f )p( f )df p( f{d j}) = , (2) 0.10.50.910010005000Semi-major axis (AU)110Mass (MJup)0.10.50.910010005000Semi-major axis (AU)110Mass (MJup) WEIRD 17 Figure 14. Overall completeness map for our survey. Our results are shown in shades of magenta and the contours correspond to the probability of detecting a planet of a giving mass and semi-major axis. The solid green box is the PSYM-WIDE survey (Naud et al. 2017), the solid brown box is the survey of Durkan et al. (2016), and the dashed-dotted boxes correspond to high contrast direct imaging surveys: PALMS in blue (Bowler 2016), GPDS in red (Lafrenière et al. 2007b), NaCo Survey of Young Nearby Dusty Stars (Rameau et al. 2013) in brown, NaCo-LP in yellow (Chauvin et al. 2015), IDPS-AF in orange (Vigan et al. 2012), MMT L(cid:48) and M-band Survey of Nearby Sun-like Stars (Heinze et al. 2010) in purple, Gemini NICI Planet-finding Campaign (Biller et al. 2013) in turquoise, MASSIVE in lime green (Lannier et al. 2016) and IDPS in olive green (Galicher et al. 2016). Our observations probe larger semi-major axes than AO imaging surveys, but are insensitive to semi-major axes where AO observations are mostly sensitive. careful in the choice of the prior, and here we elected to use a non-informative Jeffrey's prior (see Berger et al. 2009) , given by, 1√ f 1√ 1 − f P( f ) = 1 π . (3) For our survey with no detection, the posterior distribution of f peaks at 0, and we can only set an upper limit on f (by integrating the posterior from 0 to the fraction f that give a probability matching the desired confidence level). We obtained an upper limit for the fraction of stars with at least one planet of fmax = 0.03 at a 95% confidence level, for planets with masses between 1 and 13 MJup and semi-major axis between 1000 and 5000 AU distributed uniformly in log space. 5. CONCLUSIONS A sample of 177 young stars, bona fide members of mov- ing groups, were observed between 2014B and 2017B by CFHT's MegaCam in the z(cid:48) ab-band and WIRCam in the J- band, or Gemini GMOS-S in the z(cid:48) ab-band and Flamingos-2 in the J-band, as well as with Spitzer/IRAC at [3.6] and [4.5] Figure 15. Mean detection probability for 1 MJup (dash), 2 MJup (dot), 3 MJup (dash-dot) and 13 MJup (solid) companions as a func- tion of the semi-major axis in AU. where p( f ) is the prior probability on f , reflecting our state of knowledge independently of our new data. One has to be 1010010005000Semi-major axis (AU)110Mass (MJup)Lafreniere et al. 2007Heinze et al. 2010Vigan et al. 2012Biller et al. 2013Rameau et al. 2013Chauvin et al. 2015Bowler et al. 2016Galicher et al. 2016Lannier et al. 2016Durkan et al. 2016Naud et al. 20170.10.90.51001000Semi-major axis (AU)0.00.20.40.60.81.0Mean detection probability (%)1 MJup2 MJup3 MJup13 MJup 18 BARON ET AL. Figure 16. Contrast curves for Pz Tel, 2M1207, AB Pic, HR 8799, β Pictoris and 51 Eri. Known companions are shown as black points with error bars, using masses from hot start models. See text for references for the masses. to search for planetary mass companions on very wide orbits (up to 5000 AU). The survey made use of the very red z(cid:48) − J and [3.6]-[4.5] colors intrinsic to such objects and reached good sensitivities down to objects of 1 MJup. Four candi- dates were identified through colors selection but proper mo- tion follow-up obtained a year after the first epoch rejected the candidates. No planet was found. This null result al- lowed us to set an upper limit of 0.03 for the fraction of stars with at least one planet with mass between 1 and 13 MJup and semi-major axis between 1000 and 5000 AU, at a 95% 0.10.50.9Pz Tel1010010005000Semi-major axis (AU)110Mass (MJup)0.10.50.92M12071010010005000Semi-major axis (AU)110Mass (MJup)0.10.50.9AB Pic1010010005000Semi-major axis (AU)110Mass (MJup)0.10.50.9HR 87991010010005000Semi-major axis (AU)110Mass (MJup)0.10.50.90.9β Pictoris1010010005000Semi-major axis (AU)110Mass (MJup)0.10.50.951 Eri1010010005000Semi-major axis (AU)110Mass (MJup) WEIRD 19 ence Institute, the National Aeronautics and Space Adminis- tration under Grant No. NNX08AR22G issued through the Planetary Science Division of the NASA Science Mission Directorate, the National Science Foundation under Grant No. AST-1238877, the University of Maryland, and Eotvos Lorand University (ELTE) and the Los Alamos National Lab- oratory. This work has made use of data from the European Space Agency (ESA) mission Gaia (https://www.cosmos.esa.int/ gaia), processed by the Gaia Data Processing and Analysis Consortium (DPAC, https://www.cosmos.esa.int/web/gaia/ dpac/consortium). Funding for the DPAC has been provided by national institutions, in particular the institutions partici- pating in the Gaia Multilateral Agreement. Facility: Gemini-South (Flamingos-2,GMOS-S), CFHT (WIRCam, MegaCam), Spitzer (Irac) Software: SExtractor (Bertin & Arnouts 1996), Scamp (Bertin 2010a), Swarp (Bertin 2010b), CFHT'S Elixir pipeline confidence level, assuming logarithmically uniform distribu- tions in planet mass and semi-major axis.While it was not the main objective of the survey, our data also constrain the frequency of brown dwarfs to be less than 2.2% for objects with masses between 13 and 80 MJup and for semi-major axis between 1000 and 5000 AU. As mentioned above, the formation process by which Jupiter-like objects on wide orbits form has been the sub- ject of an ongoing debate. The very low occurrence rate for planets at 1000-5000 AU found by our survey indicates that neither core accretion nor disk instability is actually efficient at forming gas giants at these large separations. It is possible that the few known instances of planets at such large separa- tions from their host star represent the low-mass tail end of distribution of brown dwarf companions that form like stars, rather than objects that form like planets. More quantitative implications of our results on the properties of the overall distribution of planets around stars, as well as on the forma- tion mechanism of very distant companions will be explored in a forthcoming paper, where we will further incorporate the results of AO surveys. Based on observations obtained at the Gemini Observa- tory through programs number GS-2014B-Q-2, GS-2015A- Q-71, GS-2015B-Q-57, GS-2016A-Q-69, GS-2016B-Q-33, GS-2017A-Q-58 and GS-2017B-Q-34. The Gemini Obser- vatory is operated by the Association of Universities for Re- search in Astronomy, Inc., under a cooperative agreement with the National Science Foundation (NSF) on behalf of the Gemini partnership: the NSF (United States), the Na- tional Research Council (Canada), CONICYT (Chile), the Australian Research Council (Australia), Ministério da Ciên- cia, Tecnologia e Inovação (Brazil), and Ministerio de Cien- cia, Tecnología e Innovación Productiva (Argentina). Based on observations obtained with MegaPrime/MegaCam, a joint project of CFHT and CEA/DAPNIA, at the Canada- France-Hawaii Telescope (CFHT) which is operated by the National Research Council (NRC) of Canada, the Institut National des Science de l'Univers of the Centre National de la Recherche Scientifique (CNRS) of France, and the Uni- versity of Hawaii. The Pan-STARRS1 Surveys (PS1) have been made pos- sible through contributions of the Institute for Astronomy, the University of Hawaii, the Pan-STARRS Project Office, the Max-Planck Society and its participating institutes, the Max Planck Institute for Astronomy, Heidelberg and the Max Planck Institute for Extraterrestrial Physics, Garching, The Johns Hopkins University, Durham University, the Univer- sity of Edinburgh, Queen's University Belfast, the Harvard- Smithsonian Center for Astrophysics, the Las Cumbres Ob- servatory Global Telescope Network Incorporated, the Na- tional Central University of Taiwan, the Space Telescope Sci- 20 BARON ET AL. Table 2. Properties of the sample of bona fide members Name RA DEC SpT J H K W1 W2 HIP 490 HIP 560 HIP 1113 HIP 1134 HIP 1481 HIP 1910 AB HIP 1993 GJ 2006A HIP 2484 B HIP 2578 HIP 2729 HIP 3556 HD 4277 A HIP 4448 A G 132-51 B HD 6569 AB 2MASS J01112542+1526214 CD-12 243 2MUCD 13056 HIP 6485 G 269-153 A HIP 6856 2MASS J01351393-0712517 G271-110 HIP 9141 AB HIP 9685 HIP 9892 AB HIP 9902 HD 13482 A HIP 10602 A HIP 10679 HIP 11152 HIP 11360 HIP 11437 A 1RXSJ022735.8+471021 HIP12394 HIP12413 HIP 12545 AB AF Hor HIP 12635 HIP12925 HD 17332 A HIP 13209 HIP 14551 IS Eri HIP 14807 HIP 14913 A HIP 15247 (J2000.0) (J2000.0) 00 05 52.54 00 06 50.08 00 13 53.01 00 14 10.25 00 18 26.12 00 24 08.98 00 25 14.66 00 27 50.23 00 31 32.67 00 32 43.91 00 34 51.20 00 45 28.15 00 45 50.89 00 56 55.46 01 03 42.11 01 06 26.15 01 11 25.42 01 20 32.27 01 23 11.26 01 23 21.25 01 24 27.68 01 28 08.66 01 35 13.93 01 36 55.17 01 57 48.98 02 04 35.12 02 07 18.06 02 07 26.12 02 12 15.41 02 16 30.59 02 17 24.74 02 23 26.64 02 26 16.24 02 27 29.25 02 27 37.26 02 39 35.36 02 39 47.99 02 41 25.89 02 41 47.31 02 42 20.95 02 46 14.61 02 47 27.24 02 49 59.03 03 07 50.85 03 09 42.29 03 11 12.33 03 12 25.75 03 16 40.67 -41 45 11.0 -23 06 27.1 -74 41 17.8 -07 11 56.8 -63 28 39.0 -62 11 04.3 -61 30 48.3 -32 33 06.4 -62 57 29.6 -63 01 53.4 -61 54 58.1 -51 37 33.9 54 58 40.2 -51 52 31.9 +40 51 15.8 -14 17 47.1 15 26 21.5 -11 28 03.7 -69 21 38.0 -57 28 50.7 -33 55 08.6 -52 38 19.1 -07 12 51.8 -06 47 37.9 -21 54 05.3 -54 52 54.1 -53 11 56.5 -59 40 45.9 23 57 29.5 -51 30 43.8 28 44 30.4 22 44 06.7 06 17 33.2 30 58 24.6 47 10 04.5 -68 16 01.0 -42 53 30.03 05 59 18.4 -52 59 30.7 38 37 21.5 +05 35 33.3 19 22 18.5 27 15 37.8 -27 49 52.1 -09 34 46.6 22 25 22.7 -44 25 10.8 -03 31 48.9 G0V F3V G8V F5V F8V M0V M0V M3.5V A2V A0V K5V M3V F8V K3V M2.6V K1V M5V G0V M7.5V G6V M4.3V K1V M4.3V M3.5V G3V F2V G7V F8V K1V B8IV G2V M3V F4IV K8V M4.6V B9V A1V K6V M2V K3.5V F8V G0V B8V A5V G0V K6V A8V+F3V F6V 6.464±0.011 5.451±0.017 7.406±0.013 6.402±0.015 6.462±0.007 8.385±0.019 8.615±0.021 8.882±0.032 4.664±0.254 5.061±0.033 7.337±0.007 8.481±0.011 6.645±0.009 7.040±0.021 9.372±0.036 7.909±0.017 9.082±0.019 7.026±0.011 12.320±0.029 7.241±0.013 9.203±0.034 7.405±0.009 8.964±0.017 9.707±0.022 6.856±0.015 5.696±0.041 7.347±0.017 6.534±0.011 6.203±0.009 4.026±0.298 6.570±0.013 8.182±0.007 6.028±0.011 7.870±0.029 10.306±0.021 4.443±0.296 4.678±0.266 7.904±0.021 8.481±0.027 8.377±0.015 6.859±0.031 5.868±0.011 3.657±0.294 5.891±0.011 7.156±0.023 8.358±0.017 5.118±0.025 6.457±0.013 6.189±0.017 5.331±0.045 7.087±0.025 6.170±0.035 6.248±0.033 7.708±0.031 7.943±0.037 8.236±0.038 4.677±0.075 5.156±0.075 6.721±0.031 7.867±0.019 6.399±0.021 6.522±0.045 8.839±0.046 7.427±0.031 8.512±0.033 6.654±0.039 11.710±0.033 6.946±0.031 8.659±0.045 6.944±0.021 8.387±0.023 9.137±0.026 6.555±0.035 5.489±0.023 6.986±0.039 6.304±0.029 5.827±0.007 3.951±0.262 6.355±0.021 7.561±0.015 5.863±0.019 7.235±0.011 9.733±0.018 4.433±0.270 4.620±0.075 7.234±0.027 7.851±0.034 7.904±0.051 6.632±0.049 5.564±0.011 3.803±0.238 5.851±0.051 6.794±0.037 7.789±0.027 4.931±0.021 6.209±0.021 6.117±0.013 5.240±0.019 6.962±0.017 6.073±0.015 6.149±0.009 7.494±0.015 7.749±0.021 8.012±0.033 4.481±0.033 4.985±0.013 6.533±0.011 7.623±0.023 6.361±0.011 6.358±0.019 8.513±0.029 7.340±0.017 8.208±0.029 6.549±0.015 11.320±0.031 6.847±0.025 8.240±0.030 6.834±0.017 8.078±0.029 8.862±0.021 6.472±0.021 5.448±0.011 6.894±0.017 6.204±0.013 5.727±0.007 4.127±0.268 6.262±0.009 7.346±0.011 5.822±0.015 7.080±0.021 9.461±0.017 4.254±0.033 4.460±0.019 7.069±0.027 7.641±0.027 7.762±0.019 6.517±0.033 5.517±0.015 3.864±0.033 5.772±0.011 6.701±0.021 7.652±0.021 4.827±0.013 6.099±0.015 6.043±0.053 5.245±0.072 6.888±0.035 6.049±0.046 6.141±0.048 7.354±0.026 7.606±0.025 7.720±0.021 4.604±0.087 5.010±0.068 6.427±0.044 7.509±7.428 6.245±0.047 6.340±0.043 8.092±0.022 7.258±0.027 8.004±0.023 6.533±0.041 11.060±0.023 6.753±0.037 7.895±0.022 6.765±0.036 7.975±0.024 8.684±0.022 6.391±0.045 5.393±0.069 6.865±0.034 6.208±0.046 5.610±0.063 3.881±0.111 6.221±0.039 7.264±0.027 5.757±0.045 6.991±0.032 9.288±0.023 4.201±0.090 4.425±0.091 6.946±0.034 7.374±0.028 7.735±0.024 6.445±0.043 5.117±0.069 3.842±0.094 5.722±0.055 6.644±0.038 7.585±0.025 4.773±0.096 6.031±0.050 6.053±0.023 5.013±0.036 6.932±0.020 5.999±0.024 6.102±0.023 7.306±0.019 7.594±0.020 7.541±0.017 4.104±0.042 4.657±0.038 6.443±0.019 7.329±0.026 6.267±0.020 6.345±0.021 7.937±0.019 7.332±0.020 7.791±0.020 6.581±0.021 10.808±0.021 6.809±0.021 7.720±0.020 6.813±0.020 7.795±0.020 8.522±0.020 6.440±0.019 5.237±0.031 6.908±0.019 6.147±0.022 5.582±0.028 3.336±0.059 6.251±0.021 7.239±0.021 5.646±0.025 7.039±0.020 9.113±0.019 3.707±0.058 4.066±0.051 6.943±0.020 7.336±0.019 7.763±0.021 6.466±0.020 5.014±0.024 3.296±0.063 5.620±0.028 6.681±0.021 7.594±0.018 4.403±0.063 5.972±0.025 Table 2 continued Name RA DEC SpT J H K W1 W2 WEIRD Table 2 (continued) 21 HIP 15353 CD-35 1167 CD-44 1173 V577 Per 2MASS J03350208+2342356 HIP 16853 AB HIP 17248 HIP 17695 HIP 17764 HIP 17782 AB HIP 17797 HIP 18714 AB HD 25457 HD 25953 1RXS J041417.0-090650 HIP 21547 HIP 21632 HIP 21965 HIP 22295 BD+01 2447 CD-56 1032 A HIP 23179 HIP 23362 HIP 23200 HIP 23309 HIP 23418 ABCD GJ 3331 A HIP 24947 HD 35650 AB HIP 25486 V* AB Dor B HIP 26309 HIP 26369 A HIP 26453 HIP 26990 HIP 27321 HIP 28036 HIP28474 AP Col 2MASS J06085283-2753583 CD-35 2722 SCR 0613-2742AB HIP 29964 HIP 30030 HIP 30034 A HD 45270 AB AK Pic AB CD-61 1439 A HIP 32104 HIP 32235 HIP 32435 (J2000.0) (J2000.0) 03 17 59.07 03 19 08.66 03 31 55.64 03 33 13.49 03 35 02.09 03 36 53.40 03 41 37.24 03 47 23.43 03 48 11.47 03 48 23.00 03 48 35.88 04 00 31.99 04 02 36.75 04 06 41.53 04 14 17.30 04 37 36.13 04 38 43.94 04 43 17.20 04 48 05.17 04 52 24.41 04 53 30.54 04 59 15.43 04 59 15.43 04 59 34.83 05 00 47.12 05 01 58.79 05 06 49.91 05 20 38.05 05 24 30.17 05 27 04.76 05 28 44.47 05 36 10.29 05 36 55.10 05 37 39.62 05 43 35.80 05 47 17.09 05 55 43.16 06 00 41.30 06 04 52.15 06 08 52.84 06 09 19.21 06 13 13.31 06 18 28.21 06 19 08.05 06 19 12.91 06 22 30.94 06 38 00.37 06 39 50.02 06 42 24.31 06 43 46.25 06 46 13.54 -66 55 36.7 -35 07 00.3 -43 59 13.5 46 15 26.5 23 42 35.6 -49 57 28.9 55 13 06.8 -01 58 19.9 -74 41 38.8 52 02 16.3 -37 37 12.5 -41 44 54.4 -00 16 08.1 01 41 02.1 -09 06 54.4 -02 28 24.8 -27 02 01.8 -23 37 42.0 -80 46 45.3 -16 49 21.9 -55 51 31.7 37 53 25.1 37 53 25.1 01 47 00.7 -57 15 25.4 09 58 59.3 -21 35 09.1 -39 45 17.8 -38 58 10.7 -11 54 03.4 -65 26 46.3 -28 42 28.9 -47 57 48.1 -28 37 34.6 -39 55 24.6 -51 03 59.4 -38 06 16.3 -44 53 50.0 -34 33 36.0 -27 53 58.4 -35 49 31.2 -27 42 05.5 -72 02 41.4 -03 26 20.3 -58 03 15.6 -60 13 07.1 -61 32 00.2 -61 28 41.5 17 38 43.0 -71 58 35.6 -83 59 29.5 A3V K7V K6V G5V M8.5V G2V M0.5V M2.5V F3V G8V A1V G3V F5V F5V M4.3V F0V G3V F2V F7V M3V M3V A1V A1V M0V M0.5V M3V M1V F6V K6V F7V M3V A2V K6V F3V G0V A5V F7V G8V M5V M8.5V M1V M4V K4V G0V K1V G1V G1.5V K7V A2V G6V F5V 5.782±0.015 8.576±0.027 8.300±0.018 6.836±0.013 12.250±0.017 6.492±0.021 8.347±0.021 7.804±0.019 6.367±0.013 7.222±0.021 3.900±1.054 7.203±0.009 4.712±0.236 6.892±0.019 9.630±0.024 4.744±0.033 7.273±0.015 6.288±0.011 7.170±0.013 6.176±0.021 7.197±0.027 4.903±0.470 4.903±0.470 7.117±0.011 7.095±0.013 7.212±0.015 7.046±0.013 6.416±0.015 6.702±0.005 5.268±0.021 5.316±0.009 5.958±0.017 7.448±0.019 6.470±0.017 7.056±0.017 3.669±0.236 6.494±0.011 7.730±0.011 7.742±0.021 13.595±0.026 7.920±0.015 8.002±0.034 7.530±0.009 6.848±0.013 7.576±0.017 5.433±0.031 5.079±0.272 7.301±0.011 5.026±0.033 7.693±0.023 6.553±0.023 5.752±0.029 7.919±0.031 7.679±0.015 6.457±0.003 11.655±0.020 6.264±0.035 7.649±0.023 7.174±0.049 6.224±0.043 6.859±0.031 4.626±9.996 6.939±0.009 4.342±0.075 6.695±0.041 9.056±0.024 4.770±0.075 6.970±0.029 6.068±0.031 6.991±0.021 5.605±0.033 6.623±0.055 4.980±0.015 4.980±0.015 6.450±0.027 6.429±0.025 6.657±0.025 6.391±0.047 6.218±0.031 6.105±0.021 5.087±0.021 4.845±0.029 5.936±0.029 6.828±0.037 6.288±0.015 6.845±0.033 3.544±0.200 6.308±0.047 7.433±0.021 7.183±0.011 12.897±0.024 7.283±0.031 7.432±0.071 6.984±0.031 6.591±0.013 7.088±0.015 5.156±0.029 4.747±0.091 6.643±0.019 5.070±0.015 7.380±0.029 6.396±0.027 5.691±0.021 7.723±0.023 7.470±0.021 6.368±0.017 11.261±0.014 6.137±0.013 7.499±0.017 6.933±0.019 6.136±0.011 6.747±0.015 4.824±0.007 6.875±0.023 4.181±0.033 6.582±0.017 8.755±0.023 4.537±0.019 6.866±0.005 6.023±0.017 6.868±0.025 5.311±0.023 6.338±0.021 4.922±0.021 4.922±0.021 6.261±0.009 6.244±0.019 6.370±0.015 6.117±0.009 6.144±0.019 5.921±0.011 4.926±0.015 4.686±0.007 5.864±0.011 6.607±0.009 6.277±0.013 6.756±0.015 3.526±0.222 6.206±0.017 7.321±0.045 6.866±0.015 12.371±0.024 7.046±0.011 7.145±0.024 6.814±0.025 6.552±0.013 6.981±0.019 5.045±0.011 4.544±0.021 6.500±0.013 5.011±0.013 7.278±0.039 6.299±0.015 5.643±0.059 7.576±0.026 7.434±0.026 5.785±0.047 11.044±0.023 6.020±0.054 7.436±0.026 6.810±0.037 6.112±0.050 6.707±6.715 4.763±0.031 6.802±0.042 9.654±-9.000 6.503±0.040 8.586±8.432 4.486±0.081 6.861±0.035 5.931±0.058 6.788±0.038 6.765±6.602 5.837±5.289 4.956±0.110 4.956±0.110 6.173±0.046 6.129±0.050 6.180±0.048 0.000±0.000 6.109±6.057 5.854±0.053 4.924±0.070 4.598±0.121 5.919±0.053 6.544±0.035 6.235±0.050 6.718±0.036 3.663±0.100 6.175±0.048 7.257±0.030 6.642±0.039 11.976±0.024 6.929±0.033 7.042±0.035 6.679±0.040 6.408±0.041 6.888±0.035 5.088±0.073 4.492±0.081 6.424±0.045 5.022±0.064 7.276±0.030 6.312±0.046 5.540±0.024 7.607±0.020 7.426±0.020 6.102±0.019 10.767±0.020 6.022±0.022 7.448±0.021 6.684±0.019 6.095±0.022 6.687±0.038 4.304±0.021 6.827±0.020 6.076±0.114 6.563±0.021 8.319±0.023 4.085±0.049 6.899±0.022 5.934±0.023 6.821±0.019 6.532±0.038 5.200±0.042 4.653±0.073 4.653±0.073 6.079±0.023 6.093±0.022 5.977±0.024 0.000±0.000 6.101±0.053 5.806±0.024 4.543±0.043 4.189±0.057 5.776±0.026 6.524±0.019 6.213±0.023 6.728±0.019 3.003±0.040 6.153±0.022 7.290±0.019 6.404±0.021 11.623±0.021 6.877±0.019 6.851±0.020 6.692±0.020 6.485±0.021 6.906±0.020 4.748±0.038 4.026±0.044 6.488±0.021 4.718±0.040 7.319±0.020 6.273±0.021 Table 2 continued 22 BARON ET AL. Table 2 (continued) Name RA DEC SpT J H K W1 W2 HIP 33737 V* V429 Gem HIP 36349 C HIP 36948 HIP 47135 TWA21 HIP 50191 TWA 22 B TWA 1 TWA 2 A TWA 12 TWA 13 A TWA 4 AC TWA 5 A TWA30 TWA 8 B TWA 26 TWA 9 A HIP 57632 TWA 23 TWA 27 AB TWA 25 TWA 11 C GJ 490 A PX Vir GJ 1167 AB HIP 68994 HIP 74405 HIP 76629 A HD 139751 A HIP 79797 HIP 79881 HIP 81084 HD 152555 HIP 83494 HIP 84586 HIP 84642 A HD 160934 AB HIP 88399 A HIP 88726 A HIP 92024 A HIP 92680 HIP 94235 AB HIP 95261 HIP 95347 HIP 98495 HIP 99273 2MASSJ20100002-2801410 HIP 99770 HIP 100751 HIP 102141B (J2000.0) (J2000.0) 07 00 30.46 07 23 43.59 07 25 51.18 07 35 47.47 09 36 17.83 10 13 14.78 10 14 44.16 10 17 26.89 11 01 51.91 11 09 13.81 11 21 05.50 11 21 17.24 11 22 05.29 11 31 55.26 11 32 18.31 11 32 41.16 11 39 51.14 11 48 23.73 11 49 03.66 12 07 27.38 12 07 33.47 12 15 30.72 12 35 48.94 12 57 40.30 13 03 49.65 13 09 34.95 14 07 29.29 15 12 23.43 15 38 57.55 15 40 28.39 16 17 05.40 16 18 17.90 16 33 41.61 16 54 08.14 17 03 53.58 17 17 25.51 17 18 14.65 17 38 39.63 18 03 03.41 18 06 49.90 18 45 26.90 18 53 05.87 19 10 57.85 19 22 51.21 19 23 53.17 20 00 35.58 20 09 05.21 20 10 00.03 20 14 32.03 20 25 38.86 20 41 51.16 -79 41 45.9 +20 24 58.7 -30 15 52.8 -32 12 14.1 -78 20 41.7 -52 30 54.0 -42 07 18.9 -53 54 26.5 -34 42 17.0 -30 01 39.8 -38 45 16.3 -34 46 45.5 -24 46 39.8 -34 36 27.2 -30 19 51.8 -26 52 09.0 -31 59 21.5 -37 28 48.5 14 34 19.7 -32 47 00.3 -39 32 54.0 -39 48 42.6 -39 50 24.6 35 13 30.6 -05 09 42.5 28 59 06.6 -61 33 44.1 -75 15 15.6 -57 42 27.3 -18 41 46.2 -67 56 28.5 -28 36 50.5 -09 33 11.9 -04 20 24.7 34 47 24.8 -66 57 03.7 -60 27 27.5 61 14 16.0 -51 38 56.4 -43 25 30.8 -64 52 16.5 -50 10 49.9 -60 16 19.9 -54 25 26.2 -40 36 57.4 -72 54 38.0 -26 13 26.5 -28 01 41.0 36 48 22.5 -56 44 06.3 -32 26 06.8 K2V K5V M5.0V G8Vk G1V K3V A2V M6V K6V M2V M1V M1V K4V M2V M5V M5.5V M9V M1V A3V M1V M8V M0V M4V M0.5V G5V M4.8 F4V G9V K0V K3V A4V A0V M0.5V G0V A5V G5IV G8V M0V F5V A5V A7V G9IV G1V A0V B8V A0V F5V M3V A2V B2IV M4V 8.265±0.015 7.643±0.013 6.615±0.017 6.905±0.019 7.475±0.019 7.870±0.015 3.858±0.264 8.554±0.013 8.217±0.017 7.629±0.025 8.999±0.029 8.431±0.039 6.397±0.011 7.669±0.019 9.641±0.024 9.837±0.021 12.686±0.023 9.981±0.025 1.854±0.274 8.618±0.023 12.995±0.023 8.166±0.029 9.790±0.023 7.401±0.019 6.053±0.013 9.476±0.027 6.975±0.009 7.844±0.019 6.382±0.017 7.729±0.027 5.768±0.029 4.855±0.033 8.377±0.013 6.700±0.017 5.654±0.013 5.288±0.027 8.008±0.007 7.618±0.017 6.159±0.009 4.680±0.246 4.382±0.260 6.856±0.013 7.201±0.015 5.096±0.033 4.173±0.248 3.798±0.248 6.321±0.009 8.651±0.023 4.886±0.306 2.304±0.312 5.807±0.019 7.831±0.055 7.032±0.009 5.970±0.033 6.578±0.043 7.241±0.027 7.353±0.031 3.713±0.244 8.085±0.043 7.558±0.039 6.927±0.037 8.334±0.029 7.727±0.065 5.759±0.023 6.987±0.031 9.030±0.023 9.276±0.020 11.996±0.020 9.381±0.021 1.925±0.194 8.025±0.041 12.388±0.026 7.504±0.039 9.223±0.020 6.734±0.017 5.674±0.035 8.912±0.031 6.787±0.033 7.457±0.027 5.994±0.027 7.135±0.021 5.684±0.043 4.939±0.075 7.779±0.049 6.480±0.033 5.675±0.035 4.907±0.033 7.671±0.015 6.998±0.007 6.022±0.027 4.488±0.041 4.251±0.212 6.486±0.047 6.966±0.017 5.148±0.081 4.195±0.208 3.762±0.234 6.091±0.023 8.014±0.047 4.688±0.242 2.458±0.218 5.201±0.043 7.652±0.021 6.879±0.011 5.716±0.013 6.458±0.019 7.160±0.007 7.194±0.015 3.775±0.282 7.689±0.015 7.297±0.019 6.710±0.021 8.053±0.025 7.491±0.035 5.587±0.015 6.745±0.017 8.765±0.021 9.012±0.023 11.503±0.021 9.151±0.022 1.883±0.192 7.751±0.027 11.945±0.024 7.306±0.013 8.943±0.023 6.552±0.016 5.509±0.017 8.612±0.019 6.715±0.015 7.377±0.015 5.852±0.027 6.948±0.015 5.657±0.013 4.739±0.011 7.547±0.021 6.363±0.013 5.601±0.009 4.702±0.005 7.527±0.017 6.812±0.013 5.913±0.013 4.386±0.009 4.298±0.027 6.366±0.019 6.881±0.023 5.008±0.029 4.195±0.033 3.800±0.258 6.076±0.021 7.733±0.027 4.422±0.009 2.479±0.282 4.944±0.039 7.620±0.024 6.777±0.038 5.592±0.059 6.433±0.042 7.119±0.032 7.133±0.029 3.718±0.100 7.495±0.023 7.101±0.033 6.637±0.038 8.046±0.023 7.635±0.052 5.487±0.062 6.654±0.038 8.796±0.022 8.862±0.061 11.155±0.023 9.008±8.879 2.794±0.083 7.642±0.026 11.556±0.023 7.264±0.029 8.796±0.022 6.371±6.391 5.396±0.073 8.393±0.024 6.673±0.034 7.384±0.027 5.912±0.041 6.934±0.036 5.619±0.058 4.765±0.069 7.449±0.024 6.262±0.043 5.633±5.459 4.589±0.085 7.449±7.464 6.727±0.038 5.882±0.057 4.410±0.081 4.269±0.094 6.257±0.049 6.822±0.035 4.969±0.069 4.223±0.094 4.011±0.118 6.044±0.051 7.609±0.031 4.484±0.092 3.163±0.127 4.680±0.089 7.656±0.020 6.782±0.020 5.374±0.029 6.440±0.021 7.155±0.020 7.170±0.021 3.018±0.076 7.272±0.020 6.947±0.020 6.537±0.021 7.950±0.020 7.545±0.030 5.325±0.032 6.507±0.020 8.436±0.021 8.608±0.053 10.793±0.020 8.810±0.050 1.490±0.083 7.506±0.022 11.009±0.020 7.208±0.020 8.593±0.020 6.301±0.042 5.316±0.027 8.204±0.020 6.685±0.021 7.428±0.018 5.727±0.025 6.889±0.021 5.483±0.022 4.502±0.038 7.443±0.021 6.311±0.023 5.607±0.057 4.234±0.045 7.445±0.026 6.700±0.020 5.841±0.021 3.828±0.049 3.775±0.059 6.285±0.022 6.835±0.021 4.651±0.035 3.850±0.051 3.443±0.071 5.992±0.025 7.446±0.021 3.957±0.047 2.541±0.026 4.067±0.043 Table 2 continued WEIRD Table 2 (continued) 23 Name RA DEC SpT J H K W1 W2 2MASSJ20434114-2433534 HIP 102409 HIP 103311 AB HIP 105388 HIP 105404 AB HIP 107345 HIP 107947 TYC 5899-0026-1 HIP 108195 A HIP 108422 AB HIP 109268 1RXS J221419.3+253411 AB HIP 110526 A HIP 112312 A HD 217343 HIP 114066 HIP 114189 HD 218860 A HIP 115162 HIP 115738 G 190-27 A HIP 116748 A κ And HD 222575 HIP 117452 AB HIP 118121 (J2000.0) (J2000.0) 20 43 41.14 20 45 09.53 20 55 47.67 21 20 49.96 21 20 59.80 21 44 30.12 21 52 09.72 21 52 10.42 21 55 11.39 21 55 11.39 22 08 13.98 22 14 17.66 22 23 29.11 22 44 57.97 23 00 19.82 23 06 04.84 23 07 28.69 23 11 52.05 23 19 39.56 23 26 55.96 23 29 22.58 23 39 39.48 23 40 24.49 23 41 54.29 23 48 55.55 23 57 35.08 -24 33 53.1 M4.1V+M3.7V -31 20 27.2 -17 06 51.0 -53 02 03.1 -52 28 40.1 -60 58 38.9 -62 03 08.5 05 37 35.9 -61 53 11.8 -61 53 11.8 -46 57 39.5 25 34 06.6 32 27 34.1 -33 15 01.7 -26 09 13.5 63 55 34.4 21 08 03.3 -45 08 10.6 42 15 09.8 01 15 20.2 41 27 52.2 -69 11 44.7 44 20 02.1 -35 58 39.8 -28 07 49.0 -64 17 53.6 M1V F8V G7V G9V M1V F6V M3V F3V F3V B6V M4.3V M3V M4IV G5V M1V A5V G5V G4V A0V M4.2V G5V B9V G8V A0V A1V 8.481±0.027 5.436±0.003 6.207±0.009 7.386±0.013 7.184±0.026 8.751±0.019 6.358±0.021 7.740±0.021 5.242±0.033 5.242±0.033 2.021±0.350 10.177±0.016 6.898±0.011 7.786±0.009 7.048±0.007 7.815±0.015 5.383±0.021 7.467±0.019 7.605±0.009 5.317±0.270 8.017±0.021 7.122±0.017 4.624±0.264 8.097±0.017 4.801±0.262 4.910±0.033 7.851±0.034 4.831±0.005 5.945±0.035 7.026±0.035 6.699±0.031 8.087±0.017 6.149±0.027 7.146±0.031 5.227±0.075 5.227±0.075 2.027±0.228 9.624±0.018 6.279±0.007 7.154±0.027 6.448±0.019 7.167±0.037 5.280±0.011 7.109±0.023 7.275±0.007 4.984±0.005 7.406±0.025 6.759±0.017 4.595±0.218 7.771±0.041 4.643±0.075 4.949±0.027 7.641±0.027 4.529±0.013 5.811±0.013 6.908±0.017 6.574±0.024 7.874±0.021 6.027±0.015 6.891±0.027 4.909±0.009 4.909±0.009 2.016±0.244 9.339±0.016 6.054±0.009 6.932±0.025 6.267±0.007 6.977±0.017 5.240±0.011 7.032±0.017 7.224±0.021 4.902±0.011 7.166±0.015 6.676±0.031 4.571±0.354 7.624±0.015 4.532±0.015 4.824±0.015 7.374±0.028 4.499±0.086 5.718±0.056 6.815±0.035 6.536±6.520 7.781±0.024 6.013±0.052 7.209±7.068 4.903±0.074 4.903±0.074 3.659±-9.000 9.197±9.007 5.891±0.052 6.789±0.037 6.111±0.052 6.926±0.033 5.192±0.068 6.968±0.031 7.160±0.032 4.949±0.079 6.976±6.806 6.844±0.066 4.462±3.885 7.553±0.025 4.530±0.078 4.804±0.080 7.336±0.019 3.999±0.048 5.663±0.024 6.872±0.020 6.507±0.041 7.755±0.020 5.954±0.024 7.008±0.030 4.579±0.041 4.579±0.041 9.115±-9.000 8.893±0.024 5.684±0.025 6.595±0.020 6.144±0.022 6.908±0.020 4.997±0.034 7.022±0.020 7.202±0.021 4.670±0.038 6.715±0.031 6.748±0.030 4.410±0.080 7.593±0.021 4.003±0.053 4.517±0.042 24 BARON ET AL. Table 3. Properties of the sample of bona fide members Name µαcosδ µδ Radial Velocities Distance Association HIP 490 HIP 560 HIP 1113 HIP 1134 HIP 1481 HIP 1910 AB HIP 1993 GJ 2006A HIP 2484 B HIP 2578 HIP 2729 HIP 3556 HD 4277 A HIP 4448 A G 132-51 B HD 6569 AB 2MASS J01112542+1526214 CD-12 243 2MUCD 13056 HIP 6485 G 269-153 A HIP 6856 2MASS J01351393-0712517 G271-110 HIP 9141 AB HIP 9685 HIP 9892 AB HIP 9902 HD 13482 A HIP 10602 A HIP 10679 HIP 11152 HIP 11360 HIP 11437 A 1RXSJ022735.8+471021 HIP12394 HIP12413 HIP 12545 AB AF Hor HIP 12635 HIP12925 HD 17332 A HIP 13209 HIP 14551 IS Eri HIP 14807 HIP 14913 A HIP 15247 (mas/yr) 97.53±0.38 97.81±0.42 83.53±0.78 102.79±0.78 89.37±0.48 90.91±2.37 87.76±2.14 117.40±2.80 83.64±0.19 86.66±0.18 88.28±0.92 95.74±1.92 96.81±0.65 95.93±1.57 132.00±5.00 99.29±1.23 180.00±2.00 110.59±0.92 77.40±2.40 92.45±0.92 178.00±20.00 106.09±1.02 96.00±10.00 168.00±5.00 105.08±0.72 75.74±0.45 86.06±0.58 91.11±0.47 125.44±1.45 91.03±0.12 80.15±4.38 92.43±3.05 86.09±1.08 79.78±2.56 119.00±5.00 87.30±0.09 88.20±2.02 79.47±3.05 92.20±1.10 75.73±2.49 75.27±1.45 117.91±0.89 66.81±0.24 66.26±0.50 91.01±1.30 54.86±3.99 81.63±0.55 78.63±0.67 (mas/yr) -76.27±0.44 -47.12±0.21 -47.89±0.75 -66.36±0.36 -59.46±0.50 -47.25±3.04 -57.48±2.37 -29.30±8.10 -54.82±0.18 -50.33±0.17 -53.16±0.91 -58.95±1.87 -74.17±0.53 10.23±1.35 -164.00±5.00 -94.93±0.74 -120.00±5.00 -138.43±0.69 -25.40±9.00 -38.00±0.72 -110.00±20.00 -42.81±1.24 -50.00±10.00 -105.00±5.00 -50.60±0.54 -25.05±0.48 -22.60±0.65 -18.29±0.47 -161.47±0.98 -22.23±0.12 -78.40±4.91 -113.69±2.36 -50.13±0.72 -70.02±1.73 -183.00±5.00 0.09±0.10 -17.82±1.98 -53.89±1.74 -4.20±1.50 -111.45±2.73 -44.78±0.83 -161.81±0.71 -116.52±0.15 -19.09±0.49 -112.21±1.30 -134.25±3.87 -4.57±0.98 -43.82±0.71 (km/s) 1.5±1.2 6.5±3.5 9.3±0.2 -2.2±1.2 6.4±0.4 6.6±0.6 6.4±0.1 8.7±0.2 14.0±5.0 7.7±0.8 -1.0±2.0 -1.6±10.0 -14.8±1.7 1.6±0.5 -10.6±0.3 6.7±1.2 4.0±1.0 8.2±0.4 10.9±3.7 9.2±0.4 19.4±2.7 8.0±0.2 11.7±5.3 12.2±0.4 7.5±1.0 3.4±3.7 10.0±0.5 11.1±0.7 -0.3±0.2 10.2±2.0 6.5±0.7 10.4±2.0 8.8±3.0 7.0±1.1 -6.0±0.7 13.6±0.9 18.0±4.2 10.0±1.0 12.6±0.7 -4.1±0.3 4.3±1.1 3.7±0.3 4.0±4.1 13.8±0.8 14.4±0.7 4.1±0.3 13.5±2.1 9.0±0.7 (pc) 39.38±0.91 39.38±0.58 44.40±1.61 47.14±1.42 41.54±0.89 52.96±7.63 45.80±5.07 32.20±1.05 41.40±0.34 45.55±0.39 43.93±1.91 40.35±4.31 52.52±2.45 40.63±2.70 29.94±1.97 47.34±2.75 21.80±0.79 34.39±1.19 42.10±5.00 49.52±2.03 25.12±1.01 36.02±1.29 37.87±2.29 41.66±0.69 40.89±1.12 47.75±1.04 50.94±1.66 44.16±0.87 36.63±1.59 47.12±0.26 27.33±4.35 28.68±2.33 44.78±2.10 39.95±3.59 36.50±3.07 46.55±0.19 35.68±2.78 42.03±2.65 44.01±0.77 50.42±6.66 54.31±3.06 33.55±0.92 50.78±0.49 54.64±1.49 37.41±1.56 40.16±2.06 42.49±1.11 49.23±1.43 Tucana-Horologium β-Pictoris Tucana-Horologium Columba Tucana-Horologium Tucana-Horologium Tucana-Horologium β-Pictoris Tucana-Horologium Tucana-Horologium Tucana-Horologium Tucana-Horologium AB-Doradus Argus AB-Doradus AB-Doradus β-Pictoris AB-Doradus Tucana-Horologium Tucana-Horologium AB-Doradus Tucana-Horologium β-Pictoris β-Pictoris Tucana-Horologium Tucana-Horologium Tucana-Horologium Tucana-Horologium AB-Doradus Tucana-Horologium β-Pictoris β-Pictoris β-Pictoris β-Pictoris AB-Doradus Tucana-Horologium Columba β-Pictoris Tucana-Horologium AB-Doradus Tucana-Horologium AB-Doradus AB-Doradus Tucana-Horologium AB-Doradus AB-Doradus Tucana-Horologium Tucana-Horologium Table 3 continued WEIRD Table 3 (continued) 25 Name µαcosδ µδ Radial Velocities Distance Association HIP 15353 CD-35 1167 CD-44 1173 V577 Per 2MASS J03350208+2342356 HIP 16853 AB HIP 17248 HIP 17695 HIP 17764 HIP 17782 AB HIP 17797 HIP 18714 AB HD 25457 HD 25953 1RXS J041417.0-090650 HIP 21547 HIP 21632 HIP 21965 HIP 22295 BD+01 2447 CD-56 1032 A HIP 23179 HIP 23362 HIP 23200 HIP 23309 HIP 23418 ABCD GJ 3331 A HIP 24947 HD 35650 AB HIP 25486 V* AB Dor B HIP 26309 HIP 26369 A HIP 26453 HIP 26990 HIP 27321 HIP 28036 HIP28474 AP Col 2MASS J06085283-2753583 CD-35 2722 SCR 0613-2742AB 2MASS J06 HIP 29964 HIP 30030 HIP 30034 A HD 45270 AB AK Pic AB CD-61 1439 A HIP 32104 HIP 32235 HIP 32435 (mas/yr) 56.94±0.30 89.20±2.80 90.90±1.90 68.46±0.96 54.00±10.00 89.74±0.75 96.17±2.49 185.53±3.77 63.46±0.39 61.87±1.98 74.44±0.71 69.46±0.81 149.04±0.42 37.08±1.43 96.00±10.00 44.22±0.34 56.03±0.51 50.25±0.69 46.66±0.49 118.90±4.50 132.90±4.20 46.35±0.63 46.35±0.63 34.60±2.34 36.34±1.42 12.09±9.92 34.20±1.20 38.36±0.29 44.25±0.67 17.55±0.36 33.16±0.39 25.80±0.31 24.05±2.62 24.29±0.44 25.82±0.32 4.65±0.11 20.49±0.44 18.02±0.59 27.33±0.35 8.90±3.50 -6.30±2.80 -14.90±1.00 -8.32±0.86 10.90±0.75 14.36±0.74 -11.29±0.35 -47.84±1.04 -27.92±1.00 7.87±0.66 6.17±0.80 19.66±0.43 (mas/yr) 12.68±0.40 -20.30±2.80 -5.00±1.90 -176.81±0.76 -56.00±10.00 0.29±0.84 -117.69±2.26 -273.48±3.95 24.86±0.49 -70.99±1.67 -9.09±0.87 -7.00±0.85 -253.03±0.43 -94.59±1.34 -138.00±10.00 -64.39±0.27 -11.08±0.72 -11.84±0.78 41.30±0.56 -211.90±4.70 73.90±3.80 -97.80±0.41 -97.80±0.41 -94.27±1.44 70.22±1.27 -74.41±5.71 -33.80±2.10 13.06±0.50 -59.51±1.13 -50.23±0.36 150.83±0.73 -3.04±0.46 13.08±1.82 -4.06±0.74 15.08±0.52 83.10±0.15 9.34±0.44 23.85±0.75 340.92±0.35 10.70±3.50 -56.60±2.80 -2.10±1.00 72.02±1.06 -42.62±0.61 44.66±0.84 64.24±0.30 72.73±0.87 75.34±1.13 -84.32±0.48 61.15±0.87 61.60±0.47 (km/s) 26.0±0.5 13.2±0.3 15.1±0.5 -6.0±0.3 15.5±1.7 14.4±0.9 -3.2±0.6 16.0±1.7 15.5±1.3 -2.2±0.6 15.6±0.4 16.3±0.7 17.6±0.2 15.9±1.3 23.4±0.3 21.0±4.5 18.8±5.0 19.3±2.9 11.5±2.0 26.7±1.5 40.0±19.9 7.7±2.5 7.7±2.5 16.6±1.0 19.4±0.3 18.4±3.0 21.2±0.9 23.9±2.2 31.9±0.3 21.1±1.6 31.0±2.5 22.4±1.2 32.2±0.2 23.5±0.4 22.8±0.6 20.0±0.7 24.1±0.5 23.8±0.4 22.4±0.3 24.0±1.0 31.4±0.4 22.5±0.2 16.2±1.0 19.1±2.4 22.6±0.3 31.2±0.2 32.3±1.0 30.5±0.7 15.0±4.2 20.7±0.1 12.5±0.7 (pc) 54.94±0.90 45.28±0.73 45.24±0.61 34.38±1.20 42.37±2.33 43.34±1.37 35.21±2.70 16.12±0.74 54.05±1.16 51.67±4.32 50.73±2.21 48.49±1.66 18.83±0.11 55.18±2.80 23.80±1.41 29.42±0.29 56.17±2.80 63.57±3.96 61.01±1.89 16.29±0.39 11.17±0.44 52.27±2.15 52.27±2.15 25.87±1.70 26.78±0.81 24.88±1.28 19.19±0.51 48.30±0.95 18.00±0.29 27.04±0.35 14.94±0.11 52.79±1.19 25.63±4.82 56.78±1.99 55.37±1.37 19.44±0.04 54.37±1.30 52.54±1.65 8.38±0.06 31.25±3.51 21.27±1.35 29.40±0.90 38.55±1.33 49.23±1.96 46.06±1.46 23.78±0.15 21.29±0.36 22.35±0.45 43.63±1.27 58.24±2.44 56.02±1.12 AB-Doradus Tucana-Horologium Tucana-Horologium AB-Doradus β-Pictoris Tucana-Horologium Columba AB-Doradus Tucana-Horologium Tucana-Horologium Tucana-Horologium Tucana-Horologium AB-Doradus AB-Doradus AB-Doradus β-Pictoris Tucana-Horologium Tucana-Horologium Tucana-Horologium AB-Doradus AB-Doradus Columba Columba β-Pictoris β-Pictoris β-Pictoris β-Pictoris Tucana-Horologium AB-Doradus β-Pictoris AB-Doradus Columba AB-Doradus Columba Columba β-Pictoris Columba Columba Argus β-Pictoris AB-Doradus β-Pictoris β-Pictoris Columba Carina AB-Doradus AB-Doradus AB-Doradus Columba Carina Tucana-Horologium Table 3 continued 26 BARON ET AL. Table 3 (continued) Name µαcosδ µδ Radial Velocities Distance Association HIP 33737 V* V429 Gem HIP 36349 C HIP 36948 HIP 47135 TWA21 HIP 50191 TWA 22 B TWA 1 TWA 2 A TWA 12 TWA 13 A TWA 4 AC TWA 5 A TWA30 TWA 8 B TWA 26 TWA 9 A HIP 57632 TWA 23 TWA 27 AB TWA 25 TWA 11 C GJ 490 A PX Vir GJ 1167 AB HIP 68994 HIP 74405 HIP 76629 A HD 139751 A HIP 79797 HIP 79881 HIP 81084 HD 152555 HIP 83494 HIP 84586 A HIP 84642 A HD 160934 AB HIP 88399 A HIP 88726 A HIP 92024 A HIP 92680 HIP 94235 AB HIP 95261 HIP 95347 HIP 98495 HIP 99273 2MASSJ20100002-2801410 HIP 99770 HIP 100751 HIP 102141B (mas/yr) 1.56±0.94 -65.80±1.60 -130.00±10.00 -55.71±0.59 -74.85±0.59 -60.70±2.50 -150.09±0.10 -175.80±0.80 -66.19±1.85 -95.50±2.90 -68.30±2.70 -66.40±2.40 -85.40±1.73 -81.60±2.50 -89.60±1.30 -86.00±3.00 -88.00±9.00 -52.44±2.39 -497.68±0.87 -72.70±0.90 -71.60±6.70 -74.00±0.80 -45.10±2.40 -269.00±5.00 -191.13±0.86 -332.00±5.00 -69.88±0.79 -73.87±0.87 -53.98±1.14 -70.13±3.32 -45.99±0.28 -31.19±0.26 -70.05±2.75 -37.25±1.01 -60.92±0.26 -21.83±0.39 -54.62±1.09 -23.30±2.03 4.02±0.60 10.73±1.05 32.40±0.17 17.64±1.13 12.51±0.79 25.57±0.21 30.49±0.35 81.78±0.11 39.17±0.50 40.70±3.00 69.81±0.19 6.90±0.44 270.45±4.63 (mas/yr) 59.94±1.00 -228.10±1.70 -180.00±10.00 74.58±0.62 50.62±0.59 12.80±1.60 49.44±0.11 -21.30±0.80 -13.90±1.47 -23.50±2.80 -12.10±1.50 -12.50±1.80 -33.10±2.12 -29.40±2.40 -25.80±1.30 -22.00±38.00 -34.00±10.00 -22.93±1.66 -114.67±0.44 -29.30±0.90 -22.10±8.50 -27.70±0.80 -20.10±2.30 -149.00±5.00 -218.73±0.68 -210.00±5.00 -29.87±0.60 -73.08±0.92 -106.00±1.27 -159.81±2.39 -84.00±0.35 -100.92±0.18 -177.52±2.29 -114.05±0.73 -5.05±0.34 -136.91±0.42 -91.04±0.84 47.71±2.20 -86.46±0.36 -106.59±0.51 -149.48±0.17 -83.63±0.76 -100.15±0.68 -82.71±0.14 -119.21±0.18 -132.16±0.14 -68.25±0.36 -62.00±1.70 69.14±0.20 -86.02±0.32 -365.60±3.50 (km/s) 17.6±0.1 8.2±0.8 28.1±1.0 22.5±0.1 5.2±0.1 17.5±0.8 7.4±2.7 14.8±2.1 13.4±0.8 10.5±0.5 10.9±1.0 11.6±0.6 9.3±1.0 13.3±2.0 12.3±1.5 8.9±0.2 11.6±2.0 9.5±0.4 -0.2±0.5 8.5±1.2 11.2±2.0 7.5±0.1 9.0±1.0 -2.9±0.6 0.0±0.5 -5.2±2.6 -5.2±1.0 -3.5±0.1 3.6±0.9 -8.9±0.4 -9.0±4.3 -13.0±0.8 -15.0±0.4 -16.5±0.4 -21.5±1.4 3.3±1.6 1.3±0.7 -26.7±0.1 -0.2±0.5 -7.8±0.4 2.0±4.2 -4.2±0.2 8.1±0.6 13.0±4.2 -0.7±2.5 -6.7±0.7 -5.8±2.0 -5.8±0.6 -17.3±2.8 2.0±0.9 -4.0±3.7 (pc) 58.82±3.07 25.77±1.32 14.90±0.71 35.34±1.06 67.98±2.77 54.79±1.47 31.07±0.14 17.54±0.21 53.70±6.17 46.55±2.81 64.14±2.88 55.61±2.22 44.90±4.65 50.07±1.75 23.80±1.13 46.99±2.20 41.98±4.54 46.77±5.42 10.99±0.06 53.90±1.39 52.63±1.10 54.11±3.63 69.01±2.42 18.11±1.01 21.69±0.38 11.49±2.39 64.14±3.33 50.30±2.68 38.53±1.69 40.19±4.34 52.21±1.14 41.28±0.37 30.67±2.32 46.72±2.00 54.97±0.93 31.44±0.49 58.92±4.65 33.12±2.19 48.14±1.29 41.84±1.15 28.54±0.15 51.49±2.59 61.34±2.89 48.21±0.48 55.74±0.68 32.21±0.17 52.21±1.22 47.96±3.05 42.69±0.40 54.82±1.56 10.69±0.41 Carina AB-Doradus AB-Doradus Argus Argus TW Hydrae Argus β-Pictoris TW Hydrae TW Hydrae TW Hydrae TW Hydrae TW Hydrae TW Hydrae TW Hydrae TW Hydrae TW Hydrae TW Hydrae Argus TW Hydrae TW Hydrae TW Hydrae TW Hydrae Tucana-Horologium AB-Doradus Carina Argus Argus β-Pictoris AB-Doradus Argus β-Pictoris AB-Doradus AB-Doradus Tucana-Horologium β-Pictoris Tucana-Horologium AB-Doradus β-Pictoris β-Pictoris β-Pictoris β-Pictoris AB-Doradus β-Pictoris AB-Doradus Argus β-Pictoris β-Pictoris Argus Tucana-Horologium β-Pictoris Table 3 continued WEIRD Table 3 (continued) 27 Name µαcosδ µδ Radial Velocities Distance Association 2MASSJ20434114-2433534 HIP 102409 HIP 103311 AB HIP 105388 HIP 105404 AB HIP 107345 HIP 107947 TYC 5899-0026-1 HIP 108195 A HIP 108422 AB HIP 109268 1RXS J221419.3+253411 AB HIP 110526 A HIP 112312 A HD 217343 HIP 114066 HIP 114189 HD 218860 A HIP 115162 HIP 115738 G 190-27 A HIP 116748 A Kappa And HD 222575 HIP 117452 AB HIP 118121 (mas/yr) 62.00±10.00 279.96±1.26 58.81±0.83 28.77±1.01 25.45±1.69 39.98±2.35 44.05±0.41 105.70±1.50 44.50±0.23 44.50±0.23 126.69±0.14 164.00±5.00 255.30±3.10 184.76±2.64 113.54±2.13 171.46±1.59 107.93±0.60 87.53±1.39 77.52±0.73 86.68±0.31 415.00±7.50 79.00±1.10 80.73±0.14 69.49±1.18 100.80±0.25 79.12±0.47 (mas/yr) -60.00±10.00 -360.61±0.73 -62.83±0.73 -94.19±0.55 -103.88±0.73 -91.66±1.56 -92.02±0.45 -147.40±1.40 -91.07±0.27 -91.07±0.27 -147.47±0.14 -44.00±5.00 -207.80±2.90 -119.76±2.31 -162.04±1.52 -58.55±1.57 -49.63±0.46 -93.36±0.79 -66.90±0.96 -94.29±0.22 -41.00±6.70 -67.10±1.20 -18.70±0.15 -67.53±0.95 -105.34±0.23 -60.80±0.46 (km/s) -6.1±0.3 -4.5±1.3 -9.0±3.0 -0.9±0.7 6.0±2.0 2.3±0.5 1.5±0.6 -15.1±1.5 1.0±3.0 1.0±3.0 10.9±1.7 -19.9±0.3 -20.6±2.1 1.1±1.2 6.3±1.5 -23.7±0.8 -12.6±1.4 11.2±1.3 -19.7±0.2 -4.4±0.6 -14.5±0.5 6.1±0.1 -12.7±0.6 11.1±1.7 8.7±2.0 0.5±0.8 (pc) 35.58±4.93 9.90±0.10 45.66±1.60 42.97±1.80 44.44±2.76 43.64±4.91 45.33±1.35 30.49±5.25 46.46±0.88 46.46±0.88 30.96±0.20 28.73±2.06 15.51±1.56 23.34±1.96 30.54±1.89 24.50±0.96 39.40±1.08 50.76±2.83 50.15±2.86 47.05±0.64 14.79±0.39 46.29±2.78 51.62±0.50 63.69±4.58 42.14±0.39 47.43±1.10 β-Pictoris β-Pictoris β-Pictoris Tucana-Horologium Tucana-Horologium Tucana-Horologium Tucana-Horologium AB-Doradus Tucana-Horologium Tucana-Horologium AB-Doradus Columba AB-Doradus β-Pictoris AB-Doradus AB-Doradus Columba AB-Doradus AB-Doradus AB-Doradus Columba Tucana-Horologium Columba AB-Doradus AB-Doradus Tucana-Horologium 28 BARON ET AL. Table 4. J-band observations for all the target in the sample Name Filter Instrument Obs. Date Nexp Exposition Time FWHM Catalog Conditions HIP490 HIP560 HIP1113 HIP1134 HIP1481 HIP1910 HIP1993 GJ2006A HIP2484 HIP2578 HIP2729 HIP3556 HIP3589 HIP4448 G132-51B HD6569 J_G0802 J_G0802 J_G0802 F2 F2 F2 J WIRCam J_G0802 J_G0802 J_G0802 J_G0802 J_G0802 J_G0802 J_G0802 J_G0802 F2 F2 F2 F2 F2 F2 F2 F2 J WIRCam J_G0802 F2 J WIRCam J_G0802 F2 2MASSJ01112542+1526214 J WIRCam HIP6276 2MUCD13056 HIP6485 G269-153 HIP6856 2MASSJ01351393-0712517 G271-110 HIP9141 HIP9685 HIP9892 HIP9902 HIP10272 HIP10602 HIP10679 HIP11152 HIP11360 HIP11437 1RXSJ022735.8+471021 HIP12394 HIP12413 HIP12545 AFHor HIP12635 HIP12925 HIP13027 HIP13209 HIP14551 IS-Eri HIP14807 HIP14913 HIP15247 J_G0802 J_G0802 J_G0802 F2 F2 F2 J WIRCam J_G0802 J_G0802 F2 F2 J WIRCam F2 F2 F2 F2 F2 F2 J_G0802 J_G0802 J_G0802 J_G0802 J_G0802 J_G0802 J_G0802 WIRCam WIRCam J J_G0802 F2 J J WIRCam WIRCam J_G0802 J_G0802 J_G0802 J_G0802 F2 F2 F2 F2 J J WIRCam WIRCam J_G0802 F2 J J WIRCam WIRCam J_G0802 J_G0802 J_G0802 F2 F2 F2 J WIRCam (s) 612 612 630 960 612 630 630 612 722 630 630 612 960 680 1885 612 884 612 612 225 942 612 612 960 612 612 630 630 612 675 45 960 748 960 1140 630 1224 748 630 960 960 612 960 900 612 612 630 960 2014-07-25 2014-07-25 2016-10-10 2016-08-18 2014-07-25 2016-09-16 2015-09-02 2014-07-25 2014-08-01 2015-11-22 2015-09-02 2014-07-25 2015-11-17 2014-07-26 2015-05-28 2014-07-26 2015-02-16 2014-07-26 2014-08-01 2015-11-27 2015-02-12 2014-08-01 2014-07-25 2016-08-18 2014-07-26 2014-08-01 2016-01-02 2016-09-17 2014-08-03 2015-11-22 2015-11-22 2015-11-17 2014-08-01 2015-11-17 2016-08-18 2015-11-22 2014-07-20 2014-09-22 2016-08-16 2015-11-17 2015-11-17 2014-07-25 2015-11-15 2016-08-18 2014-07-25 2014-08-03 2016-11-05 2015-11-17 9 9 14 16 9 14 14 9 19 14 14 9 16 10 32 9 15 9 9 5 16 9 9 16 9 9 14 14 9 15 15 15 11 17 19 14 18 11 14 16 16 9 16 15 9 9 14 17 Table 4 continued (") 0.72 1.31 0.67 0.70 1.01 0.77 0.89 0.69 1.02 0.80 0.89 0.78 0.21 0.95 0.70 0.78 0.70 0.93 0.90 1.36 0.70 1.06 1.04 0.20 0.73 0.92 0.84 0.83 0.97 0.81 0.24 0.70 1.12 0.70 0.41 0.72 0.69 0.29 0.84 0.70 0.70 0.89 0.70 0.33 0.70 0.99 0.56 0.70 photometric median photometric 2MASS photometric 2MASS patchy clouds 2MASS patchy clouds VISTA photometric 2MASS patchy clouds 2MASS photometric 2MASS patchy clouds 2MASS patchy clouds 2MASS photometric 2MASS photometric 2MASS patchy clouds 2MASS patchy clouds 2MASS patchy clouds 2MASS patchy clouds median photometric 2MASS patchy clouds median photometric 2MASS patchy clouds 2MASS patchy clouds 2MASS patchy clouds VISTA photometric SIMON photometric 2MASS patchy clouds median photometric VISTA patchy clouds VISTA photometric median patchy clouds median patchy clouds VISTA patchy clouds VISTA patchy clouds 2MASS patchy clouds median patchy clouds 2MASS patchy clouds 2MASS 2MASS patchy clouds SIMON patchy clouds patchy clouds median VISTA photometric patchy clouds 2MASS patchy clouds 2MASS photometric 2MASS patchy clouds 2MASS 2MASS patchy clouds photometric 2MASS photometric median photometric median 2MASS patchy clouds WEIRD Table 4 (continued) 29 Name Filter Instrument Obs. Date Nexp Exposition Time FWHM Catalog Conditions HIP15353 CD-351167 CD-441173 V577-Per 2MASSJ03350208+2342356 HIP16853 HIP17248 HIP17695 HIP17764 HIP17782 HIP17797 HIP18714 HIP18859 HIP19183 1RXSJ041417.0-090650 HIP21547 HIP21632 HIP21965 HIP22295 TYC5899-0026-1 CD-561032 HIP23179 HIP23200 HIP23309 HIP23362 HIP23418 GJ3331 HIP24947 HIP25283 HIP25486 HD36705B HIP26309 HIP26369 HIP26453 HIP26990 HIP27321 HIP28036 HIP28474 AP-Col J_G0802 J_G0802 J_G0802 F2 F2 F2 J J WIRCam WIRCam J_G0802 F2 J J WIRCam WIRCam J_G0802 F2 J WIRCam J_G0802 J_G0802 F2 F2 J J J J J J WIRCam WIRCam WIRCam WIRCam WIRCam WIRCam J_G0802 J_G0802 J_G0802 F2 F2 F2 J J WIRCam WIRCam J_G0802 F2 J J J WIRCam WIRCam WIRCam J_G0802 F2 J J WIRCam WIRCam J_G0802 F2 J WIRCam J_G0802 F2 J WIRCam J_G0802 J_G0802 J_G0802 J_G0802 F2 F2 F2 F2 J WIRCam 2MASSJ06085283-2753583 J_G0802 F2 Cd-352722 2MASSJ06131330-2742054 HIP29964 HIP30030 HIP30034 HIP30314 AK-Pic CD-611439 HIP32104 HIP32235 HIP32435 J J WIRCam WIRCam J_G0802 F2 J WIRCam J_G0802 J_G0802 J_G0802 J_G0802 F2 F2 F2 F2 J WIRCam J_G0802 J_G0802 F2 F2 2015-09-02 2015-11-22 2015-09-02 2015-11-17 2017-04-19 2016-09-26 2015-11-17 2015-02-11 2015-09-02 2015-11-17 2015-09-02 2016-02-11 2015-02-11 2015-11-17 2015-02-11 2015-02-11 2016-01-18 2016-01-18 2016-02-15 2014-09-22 2016-02-15 2016-01-18 2015-02-11 2016-01-02 2016-01-18 2015-05-27 2015-05-27 2016-02-16 2015-06-01 2015-05-28 2015-09-13 2016-02-11 2016-02-11 2016-01-18 2014-09-21 2015-09-02 2015-09-02 2015-04-06 2015-06-01 2014-09-22 2016-02-11 2015-06-01 2014-09-22 2015-11-17 2015-03-22 2016-02-14 2016-02-15 2016-02-11 2015-11-17 2015-04-25 2015-04-25 14 14 13 16 16 14 16 24 14 16 14 14 16 16 16 16 17 16 14 10 14 17 21 14 16 16 16 1 16 16 14 8 14 16 9 13 14 10 54 9 13 6 9 17 14 14 14 14 17 14 14 Table 4 continued (s) 630 630 585 960 960 630 960 1355 630 900 630 630 942 960 942 942 963 960 630 680 630 963 1119 630 960 942 942 45 942 942 630 480 630 960 612 585 630 450 2886 612 766 353 612 1020 630 630 630 630 960 630 630 (") 0.96 0.88 1.12 0.70 0.00 0.69 0.70 0.70 0.87 0.70 0.87 0.87 0.70 0.21 0.70 0.70 0.70 0.70 0.75 1.31 0.59 0.70 0.70 0.76 0.70 0.70 0.70 0.65 0.70 0.70 0.84 0.70 0.53 0.70 0.69 0.87 1.00 0.74 0.70 0.87 0.70 0.70 1.01 0.70 0.61 0.57 0.64 0.53 0.40 0.68 0.76 photometric 2MASS patchy clouds 2MASS photometric median patchy clouds 2MASS patchy clouds 2MASS patchy clouds 2MASS patchy clouds 2MASS patchy clouds 2MASS photometric 2MASS patchy clouds 2MASS photometric median photometric 2MASS patchy clouds 2MASS patchy clouds 2MASS patchy clouds VISTA patchy clouds 2MASS patchy clouds 2MASS patchy clouds 2MASS photometric 2MASS patchy clouds 2MASS photometric VISTA patchy clouds 2MASS patchy clouds 2MASS patchy clouds 2MASS patchy clouds 2MASS patchy clouds 2MASS patchy clouds 2MASS patchy clouds 2MASS patchy clouds 2MASS patchy clouds 2MASS patchy clouds 2MASS patchy clouds 2MASS photometric 2MASS patchy clouds 2MASS 2MASS photometric SIMON patchy clouds photometric 2MASS 2MASS patchy clouds patchy clouds 2MASS photometric 2MASS patchy clouds 2MASS 2MASS patchy clouds photometric 2MASS patchy clouds 2MASS patchy clouds 2MASS photometric 2MASS 2MASS photometric photometric 2MASS patchy clouds 2MASS photometric 2MASS 2MASS photometric 30 BARON ET AL. Table 4 (continued) Name Filter Instrument Obs. Date Nexp Exposition Time FWHM Catalog Conditions HIP33737 BD+201790 GJ2060C HIP36948 HIP47135 TWA21 HIP50191 TWA22 HIP51317 TWA1 TWA2 TWA12 TWA13 TWA4 TWA5 TWA30A TWA8B TWA26 TWA9 HIP57632 TWA23 TWA27 TWA25 TWA11C GJ490 PX-Vir GJ1167 HIP68994 HIP74405 HIP76629 HIP76768 HIP79797 HIP79881 HIP81084 HIP82688 HIP83494 HIP84586 HIP84642 HIP86346 HIP88399 HR6750 HIP92024 HIP92680 HIP94235 Eta-TeLA HIP95347 HIP98495 HIP99273 J_G0802 F2 J J WIRCam WIRCam J_G0802 J_G0802 J_G0802 J_G0802 J_G0802 F2 F2 F2 F2 F2 J J J WIRCam WIRCam WIRCam J_G0802 J_G0802 F2 F2 J WIRCam J_G0802 F2 J J J WIRCam WIRCam WIRCam J_G0802 F2 J J WIRCam WIRCam J_G0802 J_G0802 J_G0802 F2 F2 F2 J J J WIRCam WIRCam WIRCam J_G0802 J_G0802 J_G0802 F2 F2 F2 J WIRCam J_G0802 J_G0802 J_G0802 F2 F2 F2 J J WIRCam WIRCam J_G0802 J_G0802 F2 F2 J WIRCam J_G0802 J_G0802 J_G0802 J_G0802 J_G0802 J_G0802 J_G0802 J_G0802 J_G0802 F2 F2 F2 F2 F2 F2 F2 F2 F2 2MASSJ20100002-2801410 HIP99770 HIP100751 J J WIRCam WIRCam J_G0802 F2 (s) 630 2062 942 612 630 630 720 630 884 1119 1355 630 630 960 630 960 1380 1980 1260 960 1500 630 630 630 825 960 942 630 720 630 1649 630 646 612 942 1001 612 630 942 1258 612 630 612 630 612 612 884 612 942 942 630 2015-04-25 2015-06-01 2015-06-01 2014-09-22 2015-04-23 2015-03-03 2016-01-03 2016-01-02 2016-02-15 2015-06-11 2015-06-16 2016-02-10 2016-01-30 2016-01-18 2015-12-29 2016-11-18 2016-01-18 2016-01-18 2016-02-10 2016-01-18 2016-01-18 2016-01-11 2016-02-16 2016-02-15 2016-01-26 2016-01-19 2015-06-01 2015-04-10 2015-03-02 2015-03-03 2015-06-15 2015-04-10 2014-07-25 2014-07-25 2015-07-16 2015-07-16 2014-07-25 2015-04-10 2015-07-16 2014-08-02 2014-07-25 2016-02-24 2014-07-25 2015-09-02 2014-07-25 2014-07-25 2014-08-03 2014-07-25 2015-07-15 2015-07-16 2016-08-15 14 36 16 9 14 14 16 14 15 19 23 14 14 16 14 16 23 34 28 16 26 14 14 14 14 16 16 14 16 14 16 14 19 18 16 16 18 14 16 37 18 14 18 14 9 18 26 9 16 16 14 Table 4 continued (") 0.73 0.70 0.70 1.07 0.72 0.74 0.77 0.77 0.70 0.70 0.70 0.54 0.60 0.70 0.85 0.70 0.70 0.70 0.62 0.70 0.70 0.50 0.79 0.79 0.70 0.70 0.70 0.63 0.75 0.67 0.70 0.71 0.55 0.57 0.70 0.70 0.59 0.63 0.70 0.90 0.60 0.59 0.57 0.82 0.70 0.57 0.98 0.58 0.70 0.70 0.79 2MASS 2MASS 2MASS 2MASS 2MASS 2MASS VISTA 2MASS 2MASS VISTA 2MASS VISTA VISTA 2MASS 2MASS 2MASS 2MASS 2MASS 2MASS 2MASS 2MASS VISTA VISTA VISTA 2MASS VISTA 2MASS 2MASS 2MASS 2MASS 2MASS 2MASS 2MASS VISTA VISTA 2MASS 2MASS 2MASS 2MASS 2MASS VISTA 2MASS 2MASS 2MASS 2MASS VISTA 2MASS 2MASS 2MASS 2MASS VISTA photometric patchy clouds patchy clouds patchy clouds photometric photometric patchy clouds photometric photometric patchy clouds patchy clouds photometric photometric patchy clouds photometric patchy clouds patchy clouds patchy clouds patchy clouds patchy clouds patchy clouds photometric photometric photometric patchy clouds patchy clouds patchy clouds photometric photometric photometric patchy clouds photometric photometric photometric patchy clouds patchy clouds photometric photometric patchy clouds patchy clouds photometric photometric photometric photometric patchy clouds photometric photometric photometric patchy clouds patchy clouds photometric WEIRD Table 4 (continued) 31 Name Filter Instrument Obs. Date Nexp Exposition Time FWHM Catalog Conditions HIP102141 2MASSJ20434114-2433534 HIP102409 HIP103311 HIP105388 HIP105404 HIP107345 HIP107947 HIP108195 HIP108422 HIP109268 1RXSJ221419.3+253411 HIP110526 HIP112312 HIP113579 HIP114066 HR8799 HIP114530 HIP115162 HIP115738 G190-27 HIP116748 HIP116805 HD222575 HIP117452 HIP118121 J J J WIRCam WIRCam WIRCam J_G0802 J_G0802 J_G0802 J_G0802 J_G0802 J_G0802 J_G0802 J_G0802 F2 F2 F2 F2 F2 F2 F2 F2 J J J WIRCam WIRCam WIRCam J_G0802 F2 J WIRCam J_G0802 J_G0802 F2 F2 J J J WIRCam WIRCam WIRCam J_G0802 F2 J WIRCam J_G0802 J_G0802 J_G0802 F2 F2 F2 2014-09-09 2015-07-17 2015-02-12 2014-08-01 2015-04-22 2015-04-23 2015-04-24 2015-04-23 2015-05-13 2015-04-29 2014-07-25 2014-09-12 2014-09-09 2014-09-09 2014-07-20 2016-01-19 2014-07-25 2015-04-29 2015-07-16 2015-07-16 2014-09-09 2014-08-01 2015-07-15 2015-09-02 2016-10-09 2016-06-10 16 15 22 9 16 14 14 14 14 14 9 21 20 20 9 18 9 14 15 16 16 9 19 14 14 14 (s) 942 884 1296 612 720 630 630 630 630 630 612 1237 1119 1119 612 1060 612 630 942 942 942 612 1119 630 630 630 (") 0.70 0.20 0.70 0.70 0.84 0.80 0.68 0.85 0.66 0.74 0.75 0.70 0.70 0.70 0.47 0.70 0.92 0.60 0.70 0.70 0.70 0.96 0.70 0.76 0.54 0.99 2MASS 2MASS 2MASS 2MASS 2MASS VISTA 2MASS 2MASS VISTA median VISTA 2MASS 2MASS 2MASS median 2MASS median median 2MASS VISTA 2MASS 2MASS 2MASS 2MASS median 2MASS patchy clouds patchy clouds patchy clouds photometric photometric photometric photometric photometric patchy clouds photometric patchy clouds patchy clouds patchy clouds patchy clouds photometric patchy clouds photometric patchy clouds patchy clouds patchy clouds patchy clouds patchy clouds patchy clouds photometric patchy clouds photometric 32 BARON ET AL. Table 5. z(cid:48) ab-band observation for all the target in the sample Name Filter Instrument Obs. Date Nexp Exposition Time FWHM Catalog Conditions HIP490 HIP560 HIP1113 HIP1134 HIP1481 HIP1910 HIP1993 GJ2006A HIP2484 HIP2578 HIP2729 HIP3556 HIP3589 HIP4448 G132-51B HD6569 2MASSJ01112542+1526214 HIP6276 2MUCD13056 HIP6485 G269-153 HIP6856 2MASSJ01351393-0712517 G271-110 HIP9141 HIP9685 HIP9892 HIP9902 HIP10272 HIP10602 HIP10679 HIP11152 HIP11360 HIP11437 1RXSJ022735.8+471021 HIP12394 HIP12413 HIP12545 AFHor HIP12635 HIP12925 HIP13027 HIP13209 HIP14551 IS-Eri HIP14807 HIP14913 HIP15247 GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S 2014-09-09 z_G0328 2014-10-10 z_G0328 z_G0328 2015-09-02 z.MP9901 MegaCam 2016-09-07 2014-07-20 z_G0328 2015-08-03 z_G0328 2015-08-24 z_G0328 2016-06-10 z_G0328 z_G0328 2014-09-09 2015-08-03 z_G0328 2015-09-02 z_G0328 z_G0328 2014-10-10 z.MP9901 MegaCam 2015-09-10 z_G0328 2014-08-26 z.MP9801 MegaCam 2014-07-18 z_G0328 2016-06-08 z.MP9801 MegaCam 2014-06-29 2014-09-08 z_G0328 2013-12-25 z_G0328 z_G0328 2015-09-02 z.MP9801 MegaCam 2014-08-24 z_G0328 2014-10-10 z_G0328 2011-09-22 z.MP9901 MegaCam 2016-09-08 2014-10-10 z_G0328 z_G0328 2016-08-04 2015-09-03 z_G0328 2016-08-04 z_G0328 2014-10-07 z_G0328 z_G0328 2016-08-01 z.MP9801 MegaCam 2014-08-24 z.MP9801 MegaCam 2014-08-24 z_G0328 2015-09-28 z.MP9901 MegaCam 2015-09-08 z.MP9901 MegaCam 2016-10-22 2015-08-24 z_G0328 2014-07-20 z_G0328 2015-09-28 z_G0328 z_G0328 2015-09-03 z.MP9901 MegaCam 2015-09-08 z.MP9901 MegaCam 2015-09-09 z_G0328 2014-09-24 z.MP9901 MegaCam 2015-09-09 z.MP9901 MegaCam 2016-11-04 2014-09-08 z_G0328 2014-10-18 z_G0328 z_G0328 2016-08-04 z.MP9901 MegaCam 2015-09-08 GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S (s) 1200 1200 520 1244 1350 585 715 520 1200 520 520 1500 1245 1200 1245 455 1245 1650 602 520 3809 1500 602 1244 1800 520 390 520 1200 520 1244 1867 520 1244 2178 520 1200 455 520 1867 1244 1200 1245 3111 900 1200 520 1245 8 8 8 4 9 9 11 8 8 8 8 10 4 8 4 7 4 11 3 8 8 10 3 4 12 8 6 8 8 8 4 6 8 4 7 8 8 7 8 6 4 8 4 10 6 8 8 4 (") 5.57 0.93 0.76 3.20 0.89 0.98 0.88 0.99 1.04 1.00 0.85 0.90 0.56 0.79 3.30 0.91 3.78 0.65 0.76 0.65 0.59 0.95 0.91 0.67 0.91 0.39 0.75 0.95 0.68 0.84 0.49 3.48 0.86 2.84 0.76 0.91 0.72 0.86 0.77 0.51 3.29 0.79 0.47 0.77 1.08 0.90 0.88 3.53 skymapper panstarrs skymapper skymapper median skymapper skymapper skymapper median median skymapper skymapper panstarrs median SDSS skymapper SDSS panstarrs skymapper skymapper skymapper skymapper SDSS skymapper skymapper median skymapper median SDSS median SDSS SDSS median SDSS median median median SDSS skymapper panstarrs SDSS median panstarrs skymapper skymapper median median panstarrs patchy clouds patchy clouds patchy clouds photometric patchy clouds patchy clouds photometric patchy clouds patchy clouds patchy clouds patchy clouds photometric photometric patchy clouds photometric patchy clouds photometric patchy clouds photometric patchy clouds patchy clouds patchy clouds photometric patchy clouds patchy clouds patchy clouds photometric patchy clouds patchy clouds patchy clouds photometric photometric patchy clouds photometric patchy clouds patchy clouds patchy clouds patchy clouds patchy clouds photometric photometric patchy clouds photometric photometric 0 patchy clouds patchy clouds photometric Table 5 continued WEIRD Table 5 (continued) 33 Name Filter Instrument Obs. Date Nexp Exposition Time FWHM Catalog Conditions GMOS-S GMOS-S HIP15353 CD-351167 CD-441173 V577-Per 2MASSJ03350208+2342356 HIP16853 HIP17248 HIP17695 HIP17764 HIP17782 HIP17797 HIP18714 HIP18859 HIP19183 1RXSJ041417.0-090650 HIP21547 HIP21632 HIP21965 HIP22295 TYC5899-0026-1 CD-561032 HIP23179 HIP23200 HIP23309 HIP23362 HIP23418 GJ3331 HIP24947 HIP25283 HIP25486 HD36705B HIP26309 HIP26369 HIP26453 HIP26990 HIP27321 HIP28036 HIP28474 AP-Col 2MASSJ06085283-2753583 Cd-352722 2MASSJ06131330-2742054 HIP29964 HIP30030 HIP30034 HIP30314 AK-Pic CD-611439 HIP32104 HIP32235 HIP32435 GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S z_G0328 2015-09-28 z_G0328 2015-09-08 z_G0328 2015-09-28 z.MP9901 MegaCam 2015-09-10 2013-01-03 z_G0328 z_G0328 2016-08-03 z.MP9901 MegaCam 2015-09-08 z.MP9801 MegaCam 2014-10-22 z_G0328 2015-09-29 z.MP9901 MegaCam 2015-09-09 z_G0328 2015-09-03 z_G0328 2016-02-11 z.MP9801 MegaCam 2014-08-26 z.MP9901 MegaCam 2015-09-09 z.MP9901 MegaCam 2016-09-08 z.MP9801 MegaCam 2014-08-29 z.MP9901 MegaCam 2016-09-08 z.MP9901 MegaCam 2016-09-08 2016-02-15 z_G0328 z_G0328 2016-02-17 z_G0328 2012-09-20 z.MP9901 MegaCam 2015-09-15 z.MP9801 MegaCam 2014-10-22 z_G0328 2015-09-28 z.MP9901 MegaCam 2017-01-28 z.MP9801 MegaCam 2014-10-26 z.MP9801 MegaCam 2014-10-23 z_G0328 2013-10-10 z.MP9801 MegaCam 2014-10-30 z.MP9801 MegaCam 2014-10-27 z_G0328 2015-08-23 z.MP9901 MegaCam 2016-01-01 z_G0328 2016-02-11 z.MP9901 MegaCam 2016-01-01 2015-09-08 z_G0328 2015-09-28 z_G0328 z_G0328 2015-09-28 z_G0328 2015-03-14 z.MP9801 MegaCam 2014-10-28 z_G0328 2012-10-12 z.MP9801 MegaCam 2014-10-27 z.MP9801 MegaCam 2014-11-17 z_G0328 2014-09-09 z.MP9901 MegaCam 2015-09-17 2015-03-14 z_G0328 2016-02-14 z_G0328 z_G0328 2016-02-14 z_G0328 2016-02-11 z.MP9901 MegaCam 2015-09-15 2015-03-14 z_G0328 z_G0328 2015-03-16 GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S (s) 520 585 520 1245 602 520 1245 1244 520 1245 520 520 1244 1244 1244 2489 1244 1244 520 520 602 1245 1244 520 4761 2800 3111 602 2380 1244 520 2489 585 2489 780 520 520 1350 2380 602 1867 1244 1200 1555 1350 520 585 520 1245 1350 1350 8 9 8 4 3 8 4 4 8 4 8 8 4 4 4 8 4 4 8 8 3 4 4 8 10 9 10 3 5 4 8 8 9 8 12 8 8 9 5 3 6 4 8 5 9 8 9 8 4 9 9 (") 0.72 0.89 0.94 0.46 1.00 1.06 0.48 3.42 0.95 0.49 0.82 0.61 0.53 0.50 0.63 3.59 0.69 0.70 0.79 0.76 1.43 0.72 3.16 0.98 0.84 0.89 3.38 0.56 5.06 0.55 0.73 4.11 0.56 0.73 0.95 0.82 0.85 0.47 3.62 0.91 0.76 0.75 1.34 0.80 0.55 0.62 0.70 0.59 0.79 0.80 0.59 median median skymapper median panstarrs median panstarrs skymapper skymapper panstarrs median median SDSS panstarrs skymapper skymapper skymapper skymapper skymapper panstarrs skymapper panstarrs panstarrs median panstarrs panstarrs skymapper median skymapper skymapper skymapper skymapper skymapper skymapper skymapper median skymapper skymapper skymapper skymapper skymapper skymapper skymapper skymapper median median skymapper skymapper median skymapper skymapper patchy clouds patchy clouds patchy clouds patchy clouds photometric patchy clouds photometric photometric patchy clouds photometric patchy clouds photometric photometric photometric photometric photometric photometric photometric patchy clouds patchy clouds photometric photometric photometric patchy clouds photometric photometric photometric photometric photometric photometric patchy clouds photometric photometric photometric patchy clouds patchy clouds patchy clouds patchy clouds photometric photometric photometric photometric patchy clouds photometric patchy clouds photometric photometric photometric patchy clouds patchy clouds patchy clouds GMOS-S GMOS-S Table 5 continued 34 BARON ET AL. Table 5 (continued) Name Filter Instrument Obs. Date Nexp Exposition Time FWHM Catalog Conditions HIP33737 BD+201790 GJ2060C HIP36948 HIP47135 TWA21 HIP50191 TWA22 HIP51317 TWA1 TWA2 TWA12 TWA13 TWA4 TWA5 TWA30A TWA8B TWA26 TWA9 HIP57632 TWA23 TWA27 TWA25 TWA11C GJ490 PX-Vir GJ1167 HIP68994 HIP74405 HIP76629 HIP76768 HIP79797 HIP79881 HIP81084 HIP82688 HIP83494 HIP84586 HIP84642 HIP86346 HIP88399 HR6750 HIP92024 HIP92680 HIP94235 Eta-TeLA HIP95347 HIP98495 HIP99273 2MASSJ20100002-2801410 HIP99770 HIP100751 GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S z_G0328 2015-03-15 z.MP9801 MegaCam 2014-10-30 z.MP9801 MegaCam 2014-11-18 2016-08-31 z_G0328 2015-03-14 z_G0328 2015-03-03 z_G0328 z_G0328 2016-01-02 z_G0328 2015-11-16 z.MP9901 MegaCam 2016-01-02 z.MP9901 MegaCam 2016-04-12 z.MP9901 MegaCam 2016-01-01 2016-02-10 z_G0328 z_G0328 2016-01-31 z.MP9901 MegaCam 2016-01-01 z_G0328 2015-11-22 z.MP9901 MegaCam 2016-12-24 z.MP9901 MegaCam 2016-01-02 2016-01-02 z.MP9901 z_G0328 2016-02-10 z.MP9801 MegaCam 2015-01-16 z.MP9901 MegaCam 2016-01-03 2015-12-17 z_G0328 2016-02-23 z_G0328 z_G0328 2016-02-15 z.MP9801 MegaCam 2014-07-03 z.MP9801 MegaCam 2014-07-03 z.MP9801 MegaCam 2014-07-03 2015-03-09 z_G0328 z_G0328 2015-03-02 z_G0328 2015-03-04 z.MP9901 MegaCam 2016-05-30 2015-03-14 z_G0328 z_G0328 2014-08-26 z_G0328 2014-09-08 z.MP9901 MegaCam 2016-09-06 z.MP9901 MegaCam 2015-08-20 z_G0328 2014-08-26 z_G0328 2015-03-22 z.MP9901 MegaCam 2016-09-06 2016-02-17 z_G0328 2014-09-02 z_G0328 z_G0328 2016-02-24 2016-03-06 z_G0328 2015-08-24 z_G0328 2014-09-03 z_G0328 2016-03-14 z_G0328 z_G0328 2014-10-09 z_G0328 2016-03-10 z.MP9901 MegaCam 2015-07-10 z.MP9901 MegaCam 2015-06-10 z_G0328 2016-08-05 GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S (s) 1350 1555 1904 520 1350 1350 520 520 1244 2856 2381 520 780 1244 650 2857 1244 1904 520 2489 2856 520 585 520 2489 1244 1244 1350 1050 1200 1244 1350 1200 1350 1244 1556 1200 1350 1245 650 2250 520 520 520 1650 520 1200 585 1867 1245 390 9 5 4 8 9 9 8 8 4 6 5 8 12 4 10 6 4 4 8 8 6 8 9 8 8 4 4 9 7 8 4 9 8 9 4 5 8 9 4 10 15 8 8 8 11 8 8 9 6 4 6 (") 0.64 3.61 4.07 0.62 0.69 0.86 0.83 0.80 3.30 0.93 0.83 0.66 0.73 0.70 0.69 0.69 0.66 3.71 0.71 4.00 0.83 0.84 0.89 0.76 4.40 0.47 3.49 0.46 0.80 0.60 0.55 0.64 0.53 0.76 3.01 0.71 0.83 0.70 0.50 1.09 1.24 0.79 0.72 0.81 1.06 0.72 0.68 0.75 0.79 0.59 0.97 median panstarrs skymapper skymapper skymapper skymapper median skymapper SDSS skymapper skymapper skymapper median skymapper skymapper skymapper skymapper skymapper skymapper panstarrs skymapper skymapper skymapper median SDSS skymapper SDSS median skymapper median skymapper median panstarrs skymapper skymapper median skymapper skymapper median skymapper median median skymapper skymapper median median median skymapper skymapper panstarrs median patchy clouds photometric photometric photometric patchy clouds patchy clouds patchy clouds patchy clouds photometric patchy clouds photometric patchy clouds patchy clouds photometric patchy clouds photometric photometric photometric 0 photometric patchy clouds patchy clouds patchy clouds patchy clouds patchy clouds photometric photometric photometric patchy clouds patchy clouds photometric patchy clouds photometric patchy clouds patchy clouds patchy clouds patchy clouds patchy clouds patchy clouds patchy clouds patchy clouds patchy clouds patchy clouds patchy clouds patchy clouds patchy clouds patchy clouds patchy clouds patchy clouds patchy clouds patchy clouds Table 5 continued WEIRD Table 5 (continued) 35 Name Filter Instrument Obs. Date Nexp Exposition Time FWHM Catalog Conditions HIP102141 2MASSJ20434114-2433534 HIP102409 HIP103311 HIP105388 HIP105404 HIP107345 HIP107947 HIP108195 HIP108422 HIP109268 1RXSJ221419.3+253411 HIP110526 HIP112312 HIP113579 HIP114066 HR8799 HIP114530 HIP115162 HIP115738 G190-27 HIP116748 HIP116805 HD222575 HIP117452 HIP118121 GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S z.MP9801 MegaCam 2014-06-28 z.MP9901 MegaCam 2016-09-06 z.MP9801 MegaCam 2014-09-18 2016-03-25 z_G0328 2015-03-31 z_G0328 2015-04-18 z_G0328 z_G0328 2015-04-19 2015-04-01 z_G0328 2015-04-23 z_G0328 2015-06-04 z_G0328 z_G0328 2014-09-07 z.MP9801 MegaCam 2014-06-22 z.MP9801 MegaCam 2014-06-22 z.MP9801 MegaCam 2014-07-01 z_G0328 2014-07-20 z.MP9801 MegaCam 2014-07-01 2015-08-24 z_G0328 z_G0328 2015-05-13 z.MP9901 MegaCam 2016-09-07 z.MP9901 MegaCam 2016-09-07 z.MP9801 MegaCam 2014-06-23 z_G0328 2016-05-22 z.MP9901 MegaCam 2016-09-07 2016-06-22 z_G0328 z_G0328 2015-08-03 2015-05-13 z_G0328 GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S GMOS-S (s) 1428 1244 2380 520 1350 1950 1800 1350 1800 1950 1200 1245 1245 3809 1200 1245 520 1350 1244 1244 1245 520 1245 520 585 1350 3 4 5 8 9 13 12 9 12 13 8 4 4 8 8 4 8 9 4 4 4 8 4 8 9 9 (") 3.63 0.63 4.11 0.84 0.85 0.80 0.94 0.67 0.74 0.74 0.86 2.96 3.05 0.65 0.65 0.71 0.84 0.67 0.53 0.47 2.73 1.10 0.62 0.80 0.89 0.75 skymapper skymapper skymapper median skymapper skymapper skymapper skymapper median skymapper median median median skymapper panstarrs panstarrs SDSS median panstarrs median panstarrs skymapper SDSS skymapper panstarrs median patchy clouds photometric patchy clouds patchy clouds patchy clouds patchy clouds photometric patchy clouds patchy clouds patchy clouds patchy clouds photometric patchy clouds photometric patchy clouds patchy clouds patchy clouds patchy clouds patchy clouds photometric photometric patchy clouds patchy clouds patchy clouds patchy clouds patchy clouds 36 BARON ET AL. Table 6. 7σ detection limits in the J band Name amin amax HIP490 HIP560 HIP1113 HIP1134 HIP1481 HIP1910 HIP1993 GJ2006A HIP2484 HIP2578 HIP2729 HIP3556 HIP3589 HIP4448 G132-51B HD6569 2MASSJ01112542+1526214 HIP6276 2MUCD13056 HIP6485 G269-153 HIP6856 2MASSJ01351393-0712517 G271-110 HIP9141 HIP9685 HIP9892 HIP9902 HIP10272 HIP10602 HIP10679 HIP11152 HIP11360 HIP11437 1RXSJ022735.8+471021 HIP12394 HIP12413 HIP12545 AFHor HIP12635 HIP12925 HIP13027 HIP13209 HIP14551 IS-Eri HIP14807 HIP14913 HIP15247 HIP15353 (cid:48) 32 46 5 19 27 22 23 26 72 52 28 15 2 22 17 3 8 27 5 19 15 26 13 11 21 30 22 36 33 53 33 2 15 2 11 53 26 19 30 17 15 34 41 25 27 34 52 31 42 (cid:48) 180 180 180 630 180 180 180 180 180 180 180 180 630 180 630 180 630 180 180 180 630 180 180 630 180 180 180 180 180 180 630 630 180 630 630 180 180 180 180 630 630 180 630 630 180 180 180 180 180 amin AU 1241 1808 232 877 1137 1163 1055 823 2996 2378 1242 596 126 907 512 153 183 935 220 945 377 953 505 259 861 1410 1137 1574 1193 2519 894 60 685 96 304 2480 919 779 806 847 815 1142 2087 1377 997 1373 2218 1524 2295 amin AU 7096 7096 8001 29717 7486 9543 8254 5802 7461 8209 7917 7271 33104 7321 18871 8531 13741 6198 7586 8924 15837 6491 6825 15127 7369 8605 9179 7958 6600 8491 17231 18081 8069 25182 17270 8388 6430 7574 4775 31785 34237 6046 32011 34443 6741 7236 7657 8871 9900 Table 6 continued J MJ Mass Limit ' 19.1 17.7 17.7 18.5 17.5 17.9 18.1 18.8 16.9 17.7 18.4 18.2 17.6 17.8 20.2 18.8 20.6 18.1 17.6 18.1 19.7 17.8 18.0 19.1 19.0 17.2 16.9 17.9 17.8 18.4 18.9 19.0 20.0 19.3 18.1 17.7 19.3 18.4 18.9 18.5 18.0 18.1 18.8 17.0 18.3 18.5 18.0 19.6 17.6 22.1 20.7 21.0 21.8 20.6 21.5 21.4 21.3 20.0 21.0 21.6 21.3 21.2 20.9 22.6 22.2 22.3 20.8 20.7 21.6 21.7 20.6 20.9 21.0 22.0 20.5 20.4 21.1 20.6 21.7 21.1 21.3 23.2 22.3 20.3 21.0 22.1 21.6 21.0 22.0 21.7 20.7 22.4 20.6 21.2 21.5 21.1 23.1 21.3 MJup 1.0 1.7 1.7 1.3 1.8 1.6 1.5 1.2 2.2 1.7 1.4 1.4 1.8 1.7 0.7 1.1 0.6 1.5 1.8 1.5 0.9 1.7 1.6 1.0 1.1 2.0 2.2 1.6 1.7 1.4 1.1 1.0 0.8 0.9 4.0 4.7 2.7 3.7 3.1 3.6 4.2 4.1 3.2 5.9 3.9 3.6 4.3 2.4 4.8 WEIRD Table 6 (continued) 37 Name amin amax CD-351167 CD-441173 V577-Per 2MASSJ03350208+2342356 HIP16853 HIP17248 HIP17695 HIP17764 HIP17782 HIP17797 HIP18714 HIP18859 HIP19183 1RXSJ041417.0-090650 HIP21547 HIP21632 HIP21965 HIP22295 TYC5899-0026-1 CD-561032 HIP23179 HIP23200 HIP23309 HIP23362 HIP23418 GJ3331 HIP24947 HIP25283 HIP25486 HD36705B HIP26309 HIP26369 HIP26453 HIP26990 HIP27321 HIP28036 HIP28474 AP-Col 2MASSJ06085283-2753583 Cd-352722 2MASSJ06131330-2742054 HIP29964 HIP30030 HIP30034 HIP30314 AK-Pic CD-611439 HIP32104 HIP32235 HIP32435 HIP33737 (cid:48) 15 30 6 2 34 8 13 33 7 60 17 34 17 2 24 17 23 30 26 29 37 4 28 36 24 38 25 41 31 41 34 33 20 21 60 37 15 4 3 3 6 15 3 17 44 44 25 33 2 22 24 (cid:48) 180 180 630 630 180 630 630 180 630 180 180 630 630 630 630 630 630 180 180 180 630 630 180 630 630 630 180 630 630 180 630 180 630 180 180 180 180 630 180 630 630 180 630 180 180 180 180 630 180 180 180 amin AU 714 1240 206 76 1467 264 208 1761 341 3068 829 633 911 43 697 927 1449 1812 790 328 1913 93 757 2166 605 731 1217 734 836 616 1774 840 1141 1146 1169 1997 785 30 101 57 185 583 162 771 1057 939 551 1453 105 1230 1429 amin AU 8829 7568 21675 26708 7810 22194 10166 9739 32574 9142 8738 11870 34785 15007 18549 35410 40070 10993 5495 2014 32948 16312 4825 38246 15687 12098 8704 11347 17044 2693 33279 4619 35792 9977 3503 9798 9468 5287 5631 13411 18531 6946 31034 8299 4285 3837 4027 27500 10494 10094 10599 Table 6 continued J MJ Mass Limit ' 18.7 17.5 18.1 17.6 18.5 18.1 21.2 17.6 16.9 16.5 18.0 20.6 17.0 20.4 19.8 17.9 16.6 17.3 18.4 20.5 17.4 18.7 19.2 17.3 18.9 19.6 17.6 20.1 19.7 19.4 17.3 20.0 17.1 17.6 19.3 17.2 17.0 22.7 18.1 19.3 18.1 16.8 16.8 18.3 21.6 20.9 19.9 18.2 17.2 17.0 17.3 22.2 20.6 20.8 20.8 21.7 20.8 22.2 21.3 20.5 20.0 21.4 22.0 20.7 22.3 22.1 21.6 20.6 21.2 20.8 20.7 21.0 20.7 21.3 21.2 20.9 21.0 21.0 21.4 21.8 20.3 20.9 22.0 20.9 21.3 20.7 20.9 20.6 22.3 20.5 20.9 20.5 19.7 20.2 21.6 23.5 22.6 21.7 21.4 21.0 20.7 21.2 MJup 3.3 5.0 4.1 4.8 3.5 4.2 1.5 4.8 6.0 6.9 4.2 1.8 5.9 1.9 2.3 4.4 6.6 5.2 3.8 1.8 5.1 3.4 2.8 5.3 3.1 2.4 4.8 2.0 2.4 2.6 1.3 0.6 1.4 1.2 0.7 1.4 1.5 0.0 1.0 0.7 1.0 1.6 1.6 0.9 0.8 0.9 1.4 2.4 3.4 3.7 3.2 38 BARON ET AL. Table 6 (continued) Name BD+201790 GJ2060C HIP36948 HIP47135 TWA21 HIP50191 TWA22 HIP51317 TWA1 TWA2 TWA12 TWA13 TWA4 TWA5 TWA30A TWA8B TWA26 TWA9 HIP57632 TWA23 TWA27 TWA25 TWA11C GJ490 PX-Vir GJ1167 HIP68994 HIP74405 HIP76629 HIP76768 HIP79797 HIP79881 HIP81084 HIP82688 HIP83494 HIP84586 HIP84642 HIP86346 HIP88399 HR6750 HIP92024 HIP92680 HIP94235 Eta-TeLA HIP95347 HIP98495 HIP99273 2MASSJ20100002-2801410 HIP99770 HIP100751 HIP102141 amin amax (cid:48) (cid:48) 21 3 16 24 172 54 20 33 2 32 24 29 4 30 2 2 6 20 102 17 3 26 25 5 22 8 13 19 4 4 4 20 13 3 3 44 24 16 21 42 65 24 30 48 49 64 29 10 2 82 56 630 630 180 180 180 180 180 630 630 630 180 180 630 180 630 630 630 180 630 630 180 180 180 630 630 630 180 180 180 630 180 180 180 630 630 180 180 630 180 180 180 180 180 180 180 180 180 630 630 180 630 amin AU 534 40 554 1603 9399 1672 347 543 129 1466 1547 1622 162 1487 101 113 264 918 1125 922 133 1403 1727 82 475 97 820 960 160 145 216 832 403 140 165 1398 1400 527 1023 1777 1860 1261 1811 2317 2729 2070 1532 460 102 4500 603 amin AU 16245 9393 6369 12249 9873 5599 3161 10274 33851 29344 11557 10021 28303 9023 26473 29619 26461 8428 6933 33978 9483 9750 12435 11418 13672 7245 11557 9063 6943 25334 9409 7439 5527 29453 34651 5666 10618 20878 8675 7539 5144 9278 11054 8688 10043 5805 9409 28427 26913 9878 6741 Table 6 continued J MJ Mass Limit ' 21.0 19.8 16.6 16.5 16.4 17.8 18.5 19.5 17.2 17.2 18.1 17.3 18.4 17.8 17.8 18.3 17.7 18.9 21.7 16.5 18.7 17.8 17.4 20.1 20.3 20.7 15.0 17.6 15.5 17.2 17.9 17.8 19.0 17.3 17.4 18.3 17.0 19.4 16.5 17.2 18.7 17.8 17.1 18.0 17.6 19.1 17.9 19.7 16.2 15.6 20.9 23.1 20.6 19.3 20.7 20.1 20.3 19.8 20.6 20.9 20.5 22.1 21.0 21.6 21.3 20.9 21.6 20.8 22.3 21.9 20.2 22.3 21.4 21.6 21.3 22.0 21.0 19.0 21.1 18.4 20.2 21.5 20.9 21.4 20.7 21.1 20.8 20.9 22.0 19.9 20.3 21.0 21.3 21.0 21.5 21.3 21.7 21.5 23.0 19.3 19.3 21.1 MJup 0.9 1.5 4.2 4.2 4.5 2.8 2.2 1.6 3.4 3.4 2.5 3.3 2.3 2.8 2.8 2.4 2.9 1.9 0.7 4.3 2.0 2.8 3.1 1.3 1.2 1.0 7.2 3.0 6.1 3.4 2.7 2.8 1.9 3.3 3.1 2.4 3.6 1.7 4.2 3.4 2.0 2.8 3.6 2.6 3.0 1.8 2.7 1.5 4.9 5.9 0.9 WEIRD Table 6 (continued) 39 Name amin amax 2MASSJ20434114-2433534 HIP102409 HIP103311 HIP105388 HIP105404 HIP107345 HIP107947 HIP108195 HIP108422 HIP109268 1RXSJ221419.3+253411 HIP110526 HIP112312 HIP113579 HIP114066 HR8799 HIP114530 HIP115162 HIP115738 G190-27 HIP116748 HIP116805 HD222575 HIP117452 HIP118121 (cid:48) 2 37 30 30 31 17 36 30 23 67 3 32 3 22 2 46 24 2 32 3 27 46 26 53 54 (cid:48) 630 630 180 180 180 180 180 180 180 180 630 630 630 180 630 180 180 630 630 630 180 630 180 180 180 amin AU 59 363 1381 1292 1384 723 1640 1414 1337 2063 86 493 77 664 51 1830 1243 120 1482 49 1242 2354 1685 2245 2562 amin AU 17711 6246 8227 7743 8008 7865 8168 8373 10457 5580 18112 9777 14713 5542 15445 7099 9146 31610 29661 9324 8342 32540 11476 7593 8547 J MJ Mass Limit ' 18.7 21.3 18.2 18.0 19.0 18.4 17.7 16.7 17.9 16.3 19.6 21.1 19.1 20.4 18.8 18.7 18.4 16.9 17.3 21.5 17.1 18.3 17.5 18.3 17.5 20.9 21.2 21.5 21.2 22.3 21.6 21.0 20.0 21.7 18.8 21.9 22.1 20.9 22.9 20.7 21.6 22.0 20.4 20.7 22.3 20.4 21.9 21.5 21.4 20.8 MJup 2.0 0.9 2.4 2.6 1.9 2.3 2.9 4.0 2.7 4.6 1.6 0.9 1.8 1.1 2.0 2.1 2.2 2.2 1.9 0.4 5.7 0.9 3.0 2.4 3.1 40 BARON ET AL. Table 7. 7σ detection limits in the [4.5] band Name amin amax HIP490 HIP560 HIP1113 HIP1134 HIP1481 HIP1910 HIP1993 GJ2006A HIP2484 HIP2578 HIP2729 HIP3556 HIP3589 HIP4448 G132-51B HD6569 2MASSJ01112542+1526214 HIP6276 2MUCD13056 HIP6485 G269-153 HIP6856 2MASSJ01351393-0712517 G271-110 HIP9141 HIP9685 HIP9892 HIP9902 HIP10272 HIP10602 HIP10679 HIP11152 HIP11360 HIP11437 1RXSJ022735.8+471021 HIP12394 HIP12413 HIP12545 AFHor HIP12635 HIP12925 HIP13027 HIP13209 HIP14551 IS-Eri HIP14807 HIP14913 HIP15247 (cid:48) 28 25 28 24 25 26 22 20 66 28 30 25 29 25 25 29 19 35 11 30 26 28 14 16 29 24 29 31 36 17 36 17 28 29 14 31 42 22 31 36 28 37 52 24 23 22 64 28 (cid:48) 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 amin AU 1087 993 1226 1132 1047 1398 989 657 2733 1257 1318 1017 1513 1024 754 1364 419 1197 455 1486 663 994 545 374 1178 1146 1467 1378 1319 792 984 482 1236 1151 395 1453 1499 908 827 1815 1499 1248 2621 1311 853 867 2703 1359 amin AU 18858 18858 21261 22574 19892 25360 21933 15417 19826 21813 21037 19322 25147 19456 14335 22670 10438 16471 20157 23715 12030 17248 18136 11491 19583 22865 24391 21148 17538 22564 13089 13735 21442 19129 13119 22291 17088 20126 12688 24145 26008 16067 24317 26164 17912 19229 20348 23575 Table 7 continued [4.5] M[4.5] Mass Limit ' 15.7 15.4 15.6 14.8 15.4 15.0 15.5 16.2 14.3 15.1 15.6 15.7 14.4 15.7 16.1 15.3 17.0 15.7 15.7 15.3 16.5 15.9 15.6 16.7 15.7 15.2 15.2 15.3 15.7 13.9 16.1 16.2 13.5 15.5 16.1 14.7 15.2 15.5 16.5 15.0 14.8 15.7 14.0 14.6 15.7 15.4 14.9 15.1 18.7 18.4 18.8 18.2 18.5 18.7 18.8 18.8 17.4 18.4 18.8 18.8 18.0 18.7 18.5 18.7 18.6 18.4 18.8 18.8 18.5 18.6 18.5 18.6 18.7 18.6 18.8 18.6 18.5 17.3 18.3 18.5 16.7 18.6 18.2 18.0 18.0 18.6 18.6 18.5 18.5 18.3 17.5 18.3 18.5 18.4 18.0 18.6 MJup 0.3 0.4 0.3 0.7 0.4 0.6 0.4 < 0.5 0.9 0.5 0.3 0.3 0.9 0.3 0.1 0.5 < 0.5 0.3 0.3 0.5 < 0.5 0.2 0.3 < 0.5 0.3 0.5 0.5 0.4 0.3 1.2 0.1 < 0.5 1.7 0.4 1.1 2.3 1.7 1.5 0.9 1.9 2.1 1.3 3.5 2.4 1.4 1.6 2.0 1.8 WEIRD Table 7 (continued) 41 Name amin amax HIP15353 CD-351167 CD-441173 V577-Per 2MASSJ03350208+2342356 HIP16853 HIP17248 HIP17695 HIP17764 HIP17782 HIP17797 HIP18714 HIP18859 HIP19183 1RXSJ041417.0-090650 HIP21547 HIP21632 HIP21965 HIP22295 TYC5899-0026-1 CD-561032 HIP23179 HIP23200 HIP23309 HIP23362 HIP23418 GJ3331 HIP24947 HIP25283 HIP25486 HD36705B HIP26309 HIP26369 HIP26453 HIP26990 HIP27321 HIP28036 HIP28474 AP-Col 2MASSJ06085283-2753583 Cd-352722 2MASSJ06131330-2742054 HIP29964 HIP30030 HIP30034 HIP30314 AK-Pic CD-611439 HIP32104 HIP32235 HIP32435 (cid:48) 35 23 24 42 10 32 18 29 31 22 44 36 49 23 14 32 28 29 25 19 47 41 25 29 32 31 50 28 23 26 38 28 34 23 22 60 29 22 32 8 28 29 26 26 18 30 35 23 31 22 30 (cid:48) 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 amin AU 1912 1117 1008 1444 407 1404 634 465 1686 1116 2253 1746 927 1258 343 954 1551 1831 1538 586 523 2133 652 771 1966 777 967 1333 410 714 574 1457 861 1295 1196 1166 1566 1135 272 263 587 847 1018 1300 829 713 741 510 1361 1258 1681 amin AU 26308 23461 20110 16465 20288 20754 16859 7723 25881 24744 24292 23220 9017 26424 11400 14091 26899 30439 29213 14602 5352 25029 12391 12823 29053 11916 9190 23130 8619 12948 7157 25280 12274 27189 26512 9308 26036 25160 4016 14963 10187 14077 18458 23575 22054 11386 10196 10702 20890 27886 26824 Table 7 continued [4.5] M[4.5] Mass Limit ' 15.2 15.4 15.7 15.4 15.3 15.4 14.8 17.5 15.0 14.1 14.5 15.3 16.4 14.7 16.8 15.4 15.0 14.6 14.8 15.1 18.1 13.9 16.5 16.6 14.4 16.5 16.9 15.1 17.3 16.1 16.6 14.9 16.5 14.8 14.8 15.5 15.0 15.1 18.9 16.4 16.9 16.2 15.5 14.6 14.2 16.6 16.4 16.9 14.7 14.6 14.6 18.9 18.8 18.8 18.1 18.5 18.6 17.6 18.6 18.7 17.6 18.0 18.7 17.8 18.4 18.7 17.7 18.7 18.6 18.7 17.5 18.3 17.5 18.6 18.8 18.3 18.5 18.4 18.5 18.6 18.2 17.4 18.5 18.5 18.6 18.5 17.0 18.7 18.7 18.6 18.9 18.6 18.5 18.4 18.1 17.5 18.5 18.0 18.6 17.9 18.5 18.4 MJup 1.8 1.6 1.4 1.6 1.6 1.6 2.1 0.5 1.9 3.3 2.6 1.7 0.9 2.3 0.8 1.6 1.9 2.4 2.1 1.9 0.3 3.6 0.9 0.8 2.8 0.9 0.7 1.9 0.6 1.1 0.8 0.4 < 0.5 0.4 0.4 0.1 0.3 0.3 < 0.5 < 0.5 < 0.5 < 0.5 0.1 0.5 0.7 0.3 0.5 0.2 1.4 1.5 1.5 42 BARON ET AL. Table 7 (continued) Name amin amax HIP33737 BD+201790 GJ2060C HIP36948 HIP47135 TWA21 HIP50191 TWA22 HIP51317 TWA1 TWA2 TWA12 TWA13 TWA4 TWA5 TWA30A TWA8B TWA26 TWA9 HIP57632 TWA23 TWA27 TWA25 TWA11C GJ490 PX-Vir GJ1167 HIP68994 HIP74405 HIP76629 HIP76768 HIP79797 HIP79881 HIP81084 HIP82688 HIP83494 HIP84586 HIP84642 HIP86346 HIP88399 HR6750 HIP92024 HIP92680 HIP94235 Eta-TeLA HIP95347 HIP98495 HIP99273 2MASSJ20100002-2801410 HIP99770 HIP100751 (cid:48) 19 25 16 22 23 12 37 12 35 19 30 19 30 23 23 0 26 10 34 62 20 10 25 14 34 37 7 7 18 12 28 23 26 28 23 29 35 14 37 29 30 38 24 19 30 36 37 25 14 22 36 (cid:48) 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 0 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 amin AU 1129 649 232 764 1550 658 1156 211 567 1031 1397 1232 1669 1024 1142 0 1241 403 1572 686 1100 505 1364 994 609 807 83 462 905 462 1109 1191 1090 847 1065 1583 1094 849 1232 1387 1255 1096 1236 1178 1446 2007 1198 1316 649 922 1974 amin AU 28165 12340 7136 16925 32549 26236 14879 8400 7804 25714 22291 30712 26630 21500 23976 0 22500 20101 22395 5267 25811 25200 25909 33043 8674 10386 5503 30712 24085 18451 19244 25003 19769 14687 22374 26322 15057 28214 15860 23052 20033 13668 24655 29374 23086 26689 15425 25003 21594 20444 26250 [4.5] M[4.5] Mass Limit ' 14.7 16.2 16.5 14.6 14.0 12.8 14.9 14.8 17.2 15.1 15.3 14.5 14.8 15.4 15.1 -3.1 15.2 15.8 15.2 15.2 15.0 14.8 15.1 14.3 17.4 16.8 18.5 8.97 14.3 10.8 15.4 13.5 14.2 16.0 14.9 15.0 15.0 13.5 16.1 13.8 13.1 15.2 14.4 14.3 14.7 13.9 15.3 14.8 15.2 11.8 12.8 18.5 18.3 17.3 17.3 18.2 16.4 17.3 16.0 18.2 18.8 18.6 18.5 18.6 18.6 18.6 0.00 18.5 18.9 18.5 15.4 18.7 18.4 18.7 18.5 18.7 18.5 18.8 13.0 17.8 13.8 18.4 17.1 17.3 18.4 18.2 18.7 17.5 17.3 18.7 17.2 16.2 17.4 17.9 18.3 18.2 17.6 17.8 18.4 18.4 14.9 16.5 MJup 1.4 0.5 0.4 1.5 2.0 4.2 1.2 1.3 0.1 1.0 0.9 1.6 1.3 0.9 1.0 319 1.0 0.7 1.0 1.0 1.1 1.3 1.0 1.8 < 0.5 0.3 < 0.5 41. 1.7 11. 0.9 2.7 1.8 0.6 1.2 1.1 1.1 2.9 0.6 2.4 3.5 1.0 1.7 1.7 1.4 2.2 0.9 1.3 1.0 7.4 4.2 Table 7 continued WEIRD Table 7 (continued) 43 Name amin amax HIP102141 2MASSJ20434114-2433534 HIP102409 HIP103311 HIP105388 HIP105404 HIP107345 HIP107947 HIP108195 HIP108422 HIP109268 1RXSJ221419.3+253411 HIP110526 HIP112312 HIP113579 HIP114066 HR8799 HIP114530 HIP115162 HIP115738 G190-27 HIP116748 HIP116805 HD222575 HIP117452 HIP118121 (cid:48) 38 19 38 25 25 24 20 25 31 25 60 18 34 31 30 14 38 25 20 29 32 46 30 20 34 30 (cid:48) 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 479 amin AU 411 540 381 1151 1083 1067 890 1142 1450 1463 1858 517 521 728 923 353 1513 1279 1023 1355 479 2111 1549 1299 1416 1423 amin AU 5121 13454 4745 21863 20576 21280 20899 21704 22249 27789 14828 13759 7427 11176 14728 11732 18865 24305 24012 22532 7083 22167 24719 30497 20177 22713 [4.5] M[4.5] Mass Limit ' 17.2 16.1 17.4 15.1 15.5 15.5 15.4 15.4 15.1 14.7 13.2 16.5 17.5 16.5 15.6 14.9 15.1 15.2 14.9 15.1 17.6 15.3 14.3 14.7 14.7 14.9 17.3 18.3 17.4 18.4 18.6 18.7 18.6 18.6 18.4 18.5 15.7 18.7 18.4 18.3 18.0 16.9 18.1 18.7 18.4 18.4 18.5 18.6 17.9 18.8 17.9 18.3 MJup 0.1 0.6 < 0.5 1.1 0.9 0.8 0.9 0.9 1.1 1.4 3.2 0.4 < 0.5 0.4 0.8 1.2 1.0 1.0 0.7 0.6 < 0.5 1.7 0.6 1.4 1.4 1.2 44 BARON ET AL. Table 8. Properties of the candidates without MIPS data or detection RA DEC Host Star z(cid:48) ab J(cid:48) [3.6] (J2000.0) (J2000.0) (mag) (mag) (mag) (mag) 02:46:00.708 02:50:00.567 04:48:00.751 05:01:00.177 05:01:00.270 05:20:00.536 06:00:00.277 06:00:00.613 06:00:00.859 06:46:00.315 07:00:00.902 09:36:00.173 09:36:00.649 11:39:00.275 23:05:00.214 23:11:00.882 23:40:00.795 HIP12925 HIP13209 HIP22295 HIP23362 HIP23362 HIP24947 HIP28474 HIP28474 HIP28474 HIP32435 HIP33737 HIP47135 HIP47135 TWA26 >23.83 >21.48 05:35:00.82 >22.73 >20.93 27:16:00.52 >24.75 >21.71 -80:46:00.34 >24.04 >20.94 -20:03:00.08 >24.04 >20.94 -20:01:00.43 >22.15 >20.00 -39:45:00.03 >24.82 >20.92 -44:53:00.56 >24.82 >20.92 -44:52:00.49 >24.82 >20.92 -44:53:00.07 >25.06 >20.50 -83:59:00.83 >24.87 >20.67 -79:40:00.69 >24.45 >20.51 -78:19:00.80 >24.45 >20.51 -78:19:00.76 >24.43 >20.93 -32:00:00.12 HIP114066 >22.46 >20.94 63:58:00.78 -45:08:00.82 HIP114530 >23.99 >21.29 44:18:00.89 HIP116805 >24.28 >20.89 16.68±0.08 15.59±0.07 18.42±0.08 18.62±0.08 17.77±0.08 18.02±0.08 18.36±0.05 17.78±0.05 17.80±0.25 19.38±0.08 18.00±0.08 17.61±0.25 19.02±0.25 17.50±0.08 18.58±0.07 16.33±0.05 17.06±0.25 [4.5] ' 16.76±0.08 15.62±0.07 18.11±0.08 17.83±0.08 17.32±0.08 17.71±0.08 18.07±0.05 17.83±0.05 17.50±0.25 18.65±0.08 18.04±0.08 17.54±0.25 18.64±0.25 17.38±0.08 17.62±0.07 16.29±0.05 16.96±0.25 Separation Distance pc 67.5130±0.019 69.7032±0.021 107.158±0.188 57.8078±0.017 80.2475±0.022 97.8883±0.035 75.5571±0.029 80.8067±0.024 83.5713±0.032 65.1345±0.164 70.2442±0.063 63.8135±0.035 73.1604±0.039 78.5368±0.024 192.299±0.072 78.7520±0.030 87.3615±0.024 54.3±3.0 50.7±0.4 61.0±1.8 60.6±2.1 60.6±2.1 48.3±0.9 52.5±1.6 52.5±1.6 52.5±1.6 56.0±1.1 58.8±3.0 67.9±2.7 67.9±2.7 41.9±4.5 24.5±0.9 50.7±2.8 51.6±0.5 Table 9. Properties of the candidates identified through color cuts Host star RA DEC (J2000.0) (J2000.0) HIP 14913 A 48.085635 94.595099 HIP 29964 244.61165 HIP 79881 HD 152555 253.51768 -44.426938 -72.054682 -28.608996 -4.3371192 Mz(cid:48) ab (mag) 18.26±0.13 18.35±1.70 18.35±1.60 18.83±0.07 MJ (mag) 15.98±0.06 16.88±0.02 16.36±0.09 16.46±0.12 M[3.6] (mag) M[4.5] (mag) 13.82±0.10 14.46±0.07 14.05±0.06 14.08±0.12 13.31±0.09 13.46±0.05 13.31±0.05 13.65±0.11 Separation µαcosδ µδ Rejected at (AU) 5559±101 1670±116 6365±442 4895±89 (mas/yr) -46.71±26.00 27.52±17.48 3.73±9.76 -10.10±13.25 (mas/yr) (sigma) 16.75±30.2700 12.81±22.5900 -6.13±9.53000 -12.86±17.6900 3 3 8 5 WEIRD 45 Table 10. Parametrization of the J-band images fraction of pixel as a logistic function Name a0 a1 a2 HIP490 HIP560 HIP1113 HIP1134 HIP1481 HIP1910 HIP1993 GJ2006A HIP2484 HIP2578 HIP2729 HIP3556 HIP3589 HIP4448 G132-51B HD6569 2MASSJ01112542+1526214 HIP6276 2MUCD13056 HIP6485 G269-153 HIP6856 2MASSJ01351393-0712517 G271-110 HIP9141 HIP9685 HIP9892 HIP9902 HIP10272 HIP10602 HIP10679 HIP11152 HIP11360 HIP11437 1RXSJ022735.8+471021 HIP12394 HIP12413 HIP12545 AFHor HIP12635 HIP12925 HIP13027 HIP13209 HIP14551 IS-Eri HIP14807 HIP14913 8.56 13.15 25.65 -0.03 9.54 27.53 16.40 15.08 23.01 11.85 20.45 14.25 12.73 14.46 15.34 25.93 1.58 9.43 24.26 17.04 1.57 10.55 15.31 1.03 11.55 11.99 20.89 20.12 -7.50 -7.50 0.20 16.98 11.00 14.92 1.54 13.79 7.66 28.66 33.15 14.59 0.06 12.36 -0.39 6.97 8.47 27.60 6.98 0.83 1.09 5.59 0.16 0.86 3.61 2.14 2.11 1.86 0.94 2.48 1.83 5.60 1.52 7.27 7.47 0.41 1.01 5.56 1.65 0.36 1.08 2.05 0.40 1.21 0.99 2.58 2.25 0.12 0.12 0.11 7.29 0.97 6.69 0.48 1.05 0.72 3.92 4.56 6.49 0.22 1.04 0.06 2.76 0.89 7.14 0.61 0.0331 0.0777 0.0308 0.0131 0.0308 0.0190 0.0159 0.0723 0.1607 0.0705 0.0422 0.0220 0.0085 0.0380 0.0293 0.0006 -0.0013 0.0352 0.0069 0.0153 0.0016 0.0148 0.0097 0.0069 0.0233 0.0556 0.0311 0.0444 0.0014 0.0014 -0.0124 0.0091 0.0427 0.0055 0.0028 0.1264 0.0215 0.0079 0.0716 0.0189 0.0131 0.0392 -0.0202 0.0645 0.0246 0.1400 0.0980 Table 10 continued 46 BARON ET AL. Table 10 (continued) Name a0 a1 a2 HIP15247 HIP15353 CD-351167 CD-441173 V577-Per 2MASSJ03350208+2342356 HIP16853 HIP17248 HIP17695 HIP17764 HIP17782 HIP17797 HIP18714 HIP18859 HIP19183 1RXSJ041417.0-090650 HIP21547 HIP21632 HIP21965 HIP22295 TYC5899-0026-1 CD-561032 HIP23179 HIP23200 HIP23309 HIP23362 HIP23418 GJ3331 HIP24947 HIP25283 HIP25486 HD36705B HIP26309 HIP26369 HIP26453 HIP26990 HIP27321 HIP28036 HIP28474 AP-Col 2MASSJ06085283-2753583 Cd-352722 2MASSJ06131330-2742054 HIP29964 HIP30030 HIP30034 HIP30314 AK-Pic CD-611439 HIP32104 HIP32235 14.17 10.55 16.33 22.96 11.54 14.97 33.27 15.42 0.81 13.24 13.72 10.60 0.24 0.53 0.11 11.22 13.27 16.60 2.09 10.35 -2.14 2.10 -7.50 -7.50 20.14 5.12 19.32 -0.36 21.43 19.84 5.58 5.12 5.14 15.41 -1.29 8.56 8.61 10.41 16.54 15.45 21.45 19.88 18.96 15.90 13.02 12.87 9.40 10.71 9.11 -0.59 31.16 1.65 0.98 2.15 3.20 5.82 6.73 7.23 6.96 0.25 1.34 4.14 0.75 0.01 0.18 0.16 4.99 5.31 4.36 0.72 1.24 0.01 0.29 0.12 0.12 2.41 1.73 7.49 0.11 3.15 6.68 1.97 0.43 1.89 2.26 0.06 0.88 0.61 1.05 2.13 5.37 5.92 7.25 8.37 1.68 4.82 1.63 2.74 0.89 1.14 0.10 12.96 0.0688 0.0476 0.0170 0.0031 0.0171 0.0031 0.1034 0.0172 0.0006 0.0156 0.0124 0.1383 -1.7864 0.0426 -0.0017 0.0019 0.0318 0.0107 0.0418 0.0383 0.1188 0.0559 0.0014 0.0014 0.0403 0.0720 0.0358 -0.0089 0.0365 0.0210 0.0576 0.0418 0.0568 0.1350 -0.0164 0.0219 0.1049 0.0289 0.0264 0.0041 0.0084 0.0028 0.0169 0.0304 0.0052 0.0255 0.1025 0.1051 0.0365 0.0042 0.0080 Table 10 continued WEIRD Table 10 (continued) Name a0 a1 a2 47 HIP32435 HIP33737 BD+201790 GJ2060C HIP36948 HIP47135 TWA21 HIP50191 TWA22 HIP51317 TWA1 TWA2 TWA12 TWA13 TWA4 TWA5 TWA30A TWA8B TWA26 TWA9 HIP57632 TWA23 TWA27 TWA25 TWA11C GJ490 PX-Vir GJ1167 HIP68994 HIP74405 HIP76629 HIP76768 HIP79797 HIP79881 HIP81084 HIP82688 HIP83494 HIP84586 HIP84642 HIP86346 HIP88399 HR6750 HIP92024 HIP92680 HIP94235 Eta-TeLA HIP95347 HIP98495 HIP99273 2MASSJ20100002-2801410 HIP99770 16.84 15.96 11.76 14.93 21.77 0.98 29.96 16.72 20.99 0.37 12.85 0.73 22.69 9.03 21.23 18.50 19.91 16.13 11.52 4.98 -2.14 0.99 34.27 38.04 11.94 14.66 1.18 4.43 19.10 17.95 21.62 18.39 9.18 18.20 24.60 13.27 18.18 9.89 30.29 4.39 15.56 27.01 14.68 14.96 15.54 -0.47 13.18 11.52 8.02 12.99 12.90 2.00 2.26 5.77 6.76 2.48 0.20 8.66 1.21 3.56 0.18 5.43 0.29 3.60 1.11 7.43 2.22 7.73 7.30 3.07 0.78 0.01 0.35 10.88 13.60 0.66 4.69 0.22 1.07 9.87 2.73 11.09 6.07 0.94 7.20 5.29 4.93 5.94 0.97 4.59 1.59 1.75 12.57 1.22 1.90 1.97 0.11 1.13 0.85 0.83 3.07 5.41 0.0758 0.0229 0.0355 0.0199 0.0199 -0.0111 0.0521 0.0821 0.0257 0.0158 0.0120 0.0035 0.0143 0.0076 0.0105 0.0133 0.0034 0.0047 0.0018 0.0195 0.0034 -0.0049 0.0066 0.0373 0.4632 0.0089 0.0118 0.0193 0.1318 0.0097 0.0776 0.0090 0.0414 0.0672 0.0026 0.0157 0.0091 0.0484 0.0371 0.0410 0.0290 0.0306 0.1004 0.0229 0.0372 -0.0471 0.1059 0.0821 0.0442 0.0051 0.0476 Table 10 continued 48 BARON ET AL. Table 10 (continued) Name a0 a1 a2 HIP100751 HIP102141 2MASSJ20434114-2433534 HIP102409 HIP103311 HIP105388 HIP105404 HIP107345 HIP107947 HIP108195 HIP108422 HIP109268 1RXSJ221419.3+253411 HIP110526 HIP112312 HIP113579 HIP114066 HR8799 HIP114530 HIP115162 HIP115738 G190-27 HIP116748 HIP116805 HD222575 HIP117452 HIP118121 9.37 -2.14 16.28 -2.14 8.16 13.80 16.31 23.04 9.72 8.81 27.56 18.00 26.00 -7.50 -7.50 0.72 15.78 5.75 11.54 28.82 4.28 15.71 8.53 0.63 14.18 8.29 12.73 4.43 0.01 6.92 0.15 0.79 1.65 1.96 3.32 0.98 0.76 3.51 0.90 8.60 0.12 0.12 0.31 7.65 0.49 1.37 10.39 1.53 6.28 0.77 0.33 1.81 0.71 1.05 0.1676 0.0085 0.0096 0.0031 0.0429 0.0211 0.0240 0.0179 0.0259 0.0592 0.0335 0.3030 0.0146 0.0014 0.0014 -0.0191 0.0101 0.0423 0.0553 0.0034 0.0745 0.0027 0.0177 0.0582 0.0231 0.0953 0.0941 WEIRD 49 Table 11. Parametrization of SPIT ZER's fraction of pixel as a logistic function Name a0 a1 a2 HIP490 HIP560 HIP1113 HIP1134 HIP1481 HIP1910 HIP1993 GJ2006A HIP2484 HIP2578 HIP2729 HIP3556 HIP3589 HIP4448 G132-51B HD6569 2MASSJ01112542+1526214 HIP6276 2MUCD13056 HIP6485 G269-153 HIP6856 2MASSJ01351393-0712517 G271-110 HIP9141 HIP9685 HIP9892 HIP9902 HIP10272 HIP10602 HIP10679 HIP11152 HIP11360 HIP11437 1RXSJ022735.8+471021 HIP12394 HIP12413 HIP12545 AFHor HIP12635 HIP12925 HIP13027 HIP13209 HIP14551 IS-Eri HIP14807 HIP14913 1.91 -0.88 5.30 0.25 1.36 5.74 3.00 6.53 -5.58 1.34 8.04 2.13 8.53 4.30 -1.26 -2.87 3.17 -0.24 -7.50 6.61 -1.17 4.19 2.78 -0.60 7.90 1.04 4.00 6.45 1.22 2.64 4.56 -3.43 -1.32 -3.68 17.71 -0.51 1.37 2.43 -3.26 -3.68 2.81 2.64 0.77 2.27 3.77 -2.81 2.04 0.45 0.14 0.98 0.30 0.34 1.00 0.56 1.46 0.01 0.31 1.43 0.42 1.68 0.79 0.11 0.08 0.68 0.10 0.12 1.24 0.12 0.76 0.61 0.18 1.46 0.33 0.81 1.34 0.22 0.57 0.94 0.11 0.16 0.04 5.62 0.13 0.28 0.57 0.06 0.04 0.55 0.57 0.26 0.40 0.77 0.08 0.35 0.0003 -0.0013 0.0030 -0.0014 -0.0005 0.0034 0.0040 0.0006 -0.0007 -0.0007 0.0015 0.0031 0.0047 0.0003 -0.0003 0.0001 0.0019 -0.0018 0.0014 0.0016 -0.0016 0.0019 0.0033 -0.0014 0.0034 0.0007 0.0028 0.0001 7.1525 0.0009 0.0002 0.0012 -0.0008 1.1444 0.0011 -0.0027 0.0017 0.0021 4.1483 4.0411 0.0021 0.0009 0.0009 0.0004 0.0019 0.0003 -0.0003 Table 11 continued 50 BARON ET AL. Table 11 (continued) Name a0 a1 a2 HIP15247 HIP15353 CD-351167 CD-441173 V577-Per 2MASSJ03350208+2342356 HIP16853 HIP17248 HIP17695 HIP17764 HIP17782 HIP17797 HIP18714 HIP18859 HIP19183 1RXSJ041417.0-090650 HIP21547 HIP21632 HIP21965 HIP22295 TYC5899-0026-1 CD-561032 HIP23179 HIP23200 HIP23309 HIP23362 HIP23418 GJ3331 HIP24947 HIP25283 HIP25486 HD36705B HIP26309 HIP26369 HIP26453 HIP26990 HIP27321 HIP28036 HIP28474 AP-Col 2MASSJ06085283-2753583 Cd-352722 2MASSJ06131330-2742054 HIP29964 HIP30030 HIP30034 HIP30314 AK-Pic CD-611439 HIP32104 HIP32235 2.25 -0.19 5.27 2.64 -3.00 -1.31 2.92 3.00 -3.20 3.28 3.99 5.37 -1.68 -1.42 5.55 1.85 2.70 3.46 2.10 4.21 1.83 6.79 0.35 0.65 3.26 -0.57 2.33 5.33 3.95 2.41 2.26 -1.65 2.17 4.46 2.23 2.56 0.42 0.53 3.24 -3.39 5.37 3.37 4.07 3.33 -7.50 1.80 -0.52 -0.25 2.93 1.64 4.99 0.53 0.18 0.91 0.57 0.05 0.27 0.57 0.95 0.09 0.79 0.79 1.03 0.06 0.12 1.15 0.43 0.93 0.67 0.56 0.74 0.36 1.30 0.14 0.27 0.58 0.12 0.43 1.20 0.74 0.55 0.64 0.07 0.40 0.99 0.45 0.54 0.18 0.18 0.87 0.10 2.41 0.59 0.81 0.68 0.12 0.41 0.11 0.13 0.61 0.43 0.88 0.0021 -0.0010 0.0027 0.0009 -0.0007 0.0001 0.0030 0.0007 -0.0004 0.0023 0.0046 -2.4497 -0.0019 -0.0015 7.8436 0.0057 0.0001 0.0006 0.0015 0.0040 0.0020 0.0049 -0.0007 -0.0013 0.0008 -0.0023 0.0041 0.0007 0.0008 0.0007 0.0010 -0.0015 -0.0001 0.0004 0.0011 0.0018 -0.0002 -0.0003 -4.1604 -0.0001 0.0034 0.0053 0.0034 0.0048 0.0014 0.0019 -0.0015 -0.0015 0.0026 0.0022 0.0049 Table 11 continued WEIRD Table 11 (continued) Name a0 a1 a2 51 HIP32435 HIP33737 BD+201790 GJ2060C HIP36948 HIP47135 TWA21 HIP50191 TWA22 HIP51317 TWA1 TWA2 TWA12 TWA13 TWA4 TWA5 TWA30A TWA8B TWA26 TWA9 HIP57632 TWA23 TWA27 TWA25 TWA11C GJ490 PX-Vir GJ1167 HIP68994 HIP74405 HIP76629 HIP76768 HIP79797 HIP79881 HIP81084 HIP82688 HIP83494 HIP84586 HIP84642 HIP86346 HIP88399 HR6750 HIP92024 HIP92680 HIP94235 Eta-TeLA HIP95347 HIP98495 HIP99273 2MASSJ20100002-2801410 HIP99770 3.89 4.50 -2.35 -3.05 4.53 4.87 3.01 -0.48 -2.14 -4.14 5.41 3.31 5.48 6.36 2.28 5.40 6.54 4.63 8.07 -1.42 0.72 4.58 -2.14 8.21 10.54 2.64 -4.00 3.50 -1.87 7.52 2.64 4.16 3.80 -1.12 -2.00 2.49 3.17 -1.56 2.82 6.27 -3.41 2.90 2.11 3.86 3.63 1.02 -3.26 -0.22 2.34 3.39 2.97 0.76 1.45 0.13 0.09 0.89 1.19 0.75 0.17 0.00 0.08 0.92 0.56 0.99 1.32 0.44 0.93 3.19 0.99 3.90 0.08 0.24 1.00 0.41 1.49 5.29 0.57 0.10 1.26 0.30 1.55 0.57 0.85 0.98 0.16 0.11 0.52 0.64 0.11 0.76 1.35 0.07 0.53 0.49 1.10 0.73 0.31 0.09 0.13 0.53 0.83 1.00 0.0029 0.0026 -0.0003 0.0020 0.0047 0.0007 0.0109 -0.0014 0.0035 -0.0002 0.0041 0.0035 0.0057 0.0003 0.0003 0.0024 0.0040 0.0011 0.0029 -0.0015 0.0532 0.0002 0.0027 0.0016 0.0008 0.0009 -0.0001 0.0046 0.0075 0.0027 0.0009 0.0083 0.0030 -0.0010 -0.0006 0.0028 0.0020 -0.0003 0.0032 0.0022 0.0013 0.0039 0.0023 0.0019 0.0024 0.0004 3.1232 -0.0025 0.0028 0.0010 0.0064 Table 11 continued 52 BARON ET AL. Table 11 (continued) Name a0 a1 a2 HIP100751 HIP102141 2MASSJ20434114-2433534 HIP102409 HIP103311 HIP105388 HIP105404 HIP107345 HIP107947 HIP108195 HIP108422 HIP109268 1RXSJ221419.3+253411 HIP110526 HIP112312 HIP113579 HIP114066 HR8799 HIP114530 HIP115162 HIP115738 G190-27 HIP116748 HIP116805 HD222575 HIP117452 HIP118121 2.51 -2.08 12.05 -2.93 0.33 4.74 4.20 2.65 2.35 1.91 4.43 -0.61 -3.08 -3.02 -0.62 -0.20 3.89 4.45 7.00 3.09 -0.51 -2.66 -1.80 -0.01 5.47 -0.48 -0.29 0.45 0.10 6.37 0.12 0.19 0.92 0.73 0.59 0.44 0.36 0.94 0.15 0.10 0.09 0.17 0.21 0.90 0.94 1.48 0.96 0.13 0.11 0.11 0.14 1.09 0.19 0.15 0.0028 -0.0009 0.0032 -0.0004 -0.0015 0.0041 0.0017 0.0023 0.0013 0.0003 0.0035 -0.0021 0.0018 -0.0004 -0.0013 -0.0013 0.0043 0.0001 0.0005 0.0004 -0.0023 -0.0001 -0.0008 -0.0019 -0.0002 -0.0010 -0.0024 WEIRD REFERENCES 53 Ahn, C. P., Alexandroff, R., Allende Prieto, C., et al. 2012, The Chauvin, G., Lagrange, A.-M., Dumas, C., et al. 2004, Astronomy Astrophysical Journal Supplement Series, 203, 21 and Astrophysics, 425, L29 Albert, L. 2006, Ph.D. Thesis, 13 Albert, L., Artigau, É., Delorme, P., et al. 2011, The Astronomical Chauvin, G., Lagrange, A.-M., Zuckerman, B., et al. 2005, Astronomy and Astrophysics, 438, L29 Journal, 141, 203 Chauvin, G., Vigan, A., Bonnefoy, M., et al. 2015, Astronomy and Alibert, Y., Pont, F., Baraffe, I., et al. 2009, Astronomy and Astrophysics, 573, A127 Astrophysics, 506, 391 Baraffe, I., Chabrier, G., Barman, T. S., Allard, F., & Hauschildt, P. H. 2003, Astronomy & Astrophysics, 402, 701 Baruteau, C., & Masset, F. 2013, Lecture Notes in Physics, Berlin Springer Verlag, 861, 201 Bate, M. R. 2012, Monthly Notices of the Royal Astronomical Society, 419, 3115 Bate, M. R., Bonnell, I. A., & Bromm, V. 2002, Monthly Notices of the Royal Astronomical Society, 332, L65 Beichman, C., Gelino, C. R., Kirkpatrick, J. D., et al. 2014, The Astrophysical Journal, 783, 68 Bell, C. P. M., Mamajek, E. E., & Naylor, T. 2016, Young Stars &amp; Planets Near the Sun, 314, 41 Berger, J. O., Bernardo, J. M., & Sun, D. 2009, The Annals of Statistics, 37, 905 Bertin, E. 2010a, Astrophysics Source Code Library, ascl:1010.063 -- . 2010b, Astrophysics Source Code Library, ascl:1010.068 Bertin, E., & Arnouts, S. 1996, Astronomy and Astrophysics Supplement Series, 117, 393 Biller, B. A., Liu, M. C., Wahhaj, Z., et al. 2010, The Astrophysical Journal Letters, 720, L82 -- . 2013, The Astrophysical Journal, 777, 160 Binks, A. S., Jeffries, R. D., & Maxted, P. F. L. 2015, Monthly Notices of the Royal Astronomical Society, 452, 173 Blunt, S., Nielsen, E. L., Rosa, D., et al. 2017, The Astronomical Journal, 153, 229 Boss, A. P. 2011, The Astrophysical Journal, 731, 74 Boulade, O., Vigroux, L. G., Charlot, X., et al. 1998, in , 614 Bowler, B. P. 2016, Publications of the Astronomical Society of the Pacific, 128, 102001 Bowler, B. P., Liu, M. C., Shkolnik, E. L., & Tamura, M. 2015, The Astrophysical Journal Supplement Series, 216, 7 Brandeker, A., Jayawardhana, R., Khavari, P., Haisch, Jr., K. E., & Mardones, D. 2006, The Astrophysical Journal, 652, 1572 Brandt, T. D., McElwain, M. W., Turner, E. L., et al. 2014, The Astrophysical Journal, 794, 159 Choquet, É., Pueyo, L., Soummer, R., et al. 2015, Techniques and Instrumentation for Detection of Exoplanets VII, 9605, 96051P Cutri, R. M., Skrutskie, M. F., van Dyk, S., et al. 2003, VizieR Online Data Catalog, 2246, 0 Daddi, E., Dickinson, M., Morrison, G., et al. 2007, The Astrophysical Journal, 670, 156 de la Reza, R., Torres, C. A. O., Quast, G., Castilho, B. V., & Vieira, G. L. 1989, The Astrophysical Journal Letters, 343, L61 Delorme, P., Delfosse, X., Albert, L., et al. 2008, Astronomy and Astrophysics, 482, 961, 00101 Delorme, P., Albert, L., Forveille, T., et al. 2010, Astronomy and Astrophysics, 518, A39 Delorme, P., Gagné, J., Girard, J. H., et al. 2013, Astronomy and Astrophysics, 553, L5 Dupuy, T. J., & Liu, M. C. 2012, The Astrophysical Journal Supplement Series, 201, 19 Durkan, S., Janson, M., & Carson, J. C. 2016, The Astrophysical Journal, 824, 58 Eikenberry, S., Bandyopadhyay, R., Bennett, J. G., et al. 2012, Ground-based and Airborne Instrumentation for Astronomy IV, 8446, 84460I Faherty, J. K., Riedel, A. R., Cruz, K. L., et al. 2016, The Astrophysical Journal Supplement Series, 225, 10 Fazio, G. G., Hora, J. L., Allen, L. E., et al. 2004, The Astrophysical Journal Supplement Series, 154, 10 Fortney, J. J., Marley, M. S., Saumon, D., & Lodders, K. 2008, The Astrophysical Journal, 683, 1104 Gagné, J., Allers, K. N., Theissen, C. A., et al. 2018a, The Astrophysical Journal Letters, 854, L27 Gagné, J., Burgasser, A. J., Faherty, J. K., et al. 2015a, The Astrophysical Journal Letters, 808, L20 Gagné, J., Lafrenière, D., Doyon, R., Malo, L., & Artigau, É. 2014, The Astrophysical Journal, 783, 121 Gagné, J., Faherty, J. K., Cruz, K. L., et al. 2015b, The Astrophysical Journal Supplement Series, 219, 33 Gagné, J., Mamajek, E. E., Malo, L., et al. 2018b, The Astrophysical Journal, 856, 23 Bryan, M. L., Bowler, B. P., Knutson, H. A., et al. 2016, The Gaia Collaboration, Brown, A. G. A., Vallenari, A., et al. 2018, Astrophysical Journal, 827, 100 ArXiv e-prints, 1804, arXiv:1804.09365 Chambers, K., & Team, P.-S. 2018, American Astronomical Gaia Collaboration, Prusti, T., de Bruijne, J. H. J., et al. 2016, Society Meeting Abstracts Astronomy and Astrophysics, 595, A1 Chatterjee, S., Ford, E. B., Matsumura, S., & Rasio, F. A. 2008, Galicher, R., Marois, C., Macintosh, B., et al. 2016, Astronomy The Astrophysical Journal, 686, 580 and Astrophysics, 594, A63 54 BARON ET AL. Gauza, B., Béjar, V. J. S., Pérez-Garrido, A., et al. 2015, The Looper, D. L., Bochanski, J. J., Burgasser, A. J., et al. 2010, The Astrophysical Journal, 804, 96 Astronomical Journal, 140, 1486 Gimeno, G., Roth, K., Chiboucas, K., et al. 2016, in Ground-based Macintosh, B., Graham, J. R., Barman, T., et al. 2015, Science, and Airborne Instrumentation for Astronomy VI, Vol. 9908 (International Society for Optics and Photonics), 99082S Goldman, B., Marsat, S., Henning, T., Clemens, C., & Greiner, J. 2010, Monthly Notices of the Royal Astronomical Society, 405, 1140 Hawley, S. L., Covey, K. R., Knapp, G. R., et al. 2002, The Astronomical Journal, 123, 3409 Heinze, A. N., Hinz, P. M., Kenworthy, M., et al. 2010, The Astrophysical Journal, 714, 1570 Hook, I. M., Jørgensen, I., Allington-Smith, J. R., et al. 2004, Publications of the Astronomical Society of the Pacific, 116, 425, 00290 Inaba, S., Wetherill, G. W., & Ikoma, M. 2003, Icarus, 166, 46 Janson, M., Bonavita, M., Klahr, H., et al. 2011, The Astrophysical Journal, 736, 89 Janson, M., Quanz, S. P., Carson, J. C., et al. 2015, Astronomy and Astrophysics, 574, A120 Janson, M., Brandt, T. D., Moro-Martín, A., et al. 2013, The Astrophysical Journal, 773, 73 Kastner, J. H., Zuckerman, B., Weintraub, D. A., & Forveille, T. 1997, Science, 277, 67 Kipping, D. M. 2013, Monthly Notices of the Royal Astronomical Society, 434, L51 Kirkpatrick, A., Pope, A., Alexander, D. M., et al. 2012a, The Astrophysical Journal, 759, 139 Kirkpatrick, J. D., Gelino, C. R., Cushing, M. C., et al. 2012b, The Astrophysical Journal, 753, 156 Kiss, L. L., Moór, A., Szalai, T., et al. 2011, Monthly Notices of the Royal Astronomical Society, 411, 117 Kraus, A. L., Shkolnik, E. L., Allers, K. N., & Liu, M. C. 2014, The Astronomical Journal, 147, 146, 00000 Lafrenière, D., Marois, C., Doyon, R., Nadeau, D., & Artigau, É. 2007a, The Astrophysical Journal, 660, 770 Lafrenière, D., Doyon, R., Marois, C., et al. 2007b, The Astrophysical Journal, 670, 1367 Lagrange, A.-M., Gratadour, D., Chauvin, G., et al. 2009, Astronomy and Astrophysics, 493, L21 Lannier, J., Delorme, P., Lagrange, A. M., et al. 2016, Astronomy and Astrophysics, 596, A83 350, 64 Makarov, V. V., & Urban, S. 2000, Monthly Notices of the Royal Astronomical Society, 317, 289 Malo, L., Doyon, R., Feiden, G. A., et al. 2014, The Astrophysical Journal, 792, 37 Malo, L., Doyon, R., Lafrenière, D., et al. 2013, The Astrophysical Journal, 762, 88 Marley, M. S., Fortney, J. J., Hubickyj, O., Bodenheimer, P., & Lissauer, J. J. 2007, The Astrophysical Journal, 655, 541 Marois, C., Macintosh, B., Barman, T., et al. 2008, Science, 322, 1348 Marois, C., Zuckerman, B., Konopacky, Q. M., Macintosh, B., & Barman, T. 2010, Nature, 468, 1080 Mayor, M., & Queloz, D. 1995, Nature, 378, 355 McMahon, R. G., Banerji, M., Gonzalez, E., et al. 2013, The Messenger, 154, 35 Millar-Blanchaer, M. A., Graham, J. R., Pueyo, L., et al. 2015, The Astrophysical Journal, 811, 18 Montet, B. T., Bowler, B. P., Shkolnik, E. L., et al. 2015, The Astrophysical Journal Letters, 813, L11 Moór, A., Szabó, G. M., Kiss, L. L., et al. 2013, Monthly Notices of the Royal Astronomical Society, 435, 1376 Mordasini, C., Alibert, Y., Klahr, H., & Henning, T. 2012, Astronomy and Astrophysics, 547, A111 Morzinski, K. M., Males, J. R., Skemer, A. J., et al. 2015, The Astrophysical Journal, 815, 108 Naud, M.-E., Artigau, É., Doyon, R., et al. 2017, The Astronomical Journal, 154, 129 Naud, M.-E., Artigau, É., Malo, L., et al. 2014, The Astrophysical Journal, 787, 5, 00000 Nayakshin, S. 2017, Monthly Notices of the Royal Astronomical Society, 470, 2387 Nielsen, E. L., Close, L. M., Biller, B. A., Masciadri, E., & Lenzen, R. 2008, The Astrophysical Journal, 674, 466 Nielsen, E. L., Liu, M. C., Wahhaj, Z., et al. 2013, The Astrophysical Journal, 776, 4 Padoan, P., & Nordlund, A. 2004, The Astrophysical Journal, 617, 559 Lépine, S., & Simon, M. 2009, The Astronomical Journal, 137, Pollack, J. B., Hubickyj, O., Bodenheimer, P., et al. 1996, Icarus, 3632 Lindegren, L., Hernandez, J., Bombrun, A., et al. 2018, arXiv:1804.09366 [astro-ph], arXiv: 1804.09366 Liu, M. C., Magnier, E. A., Deacon, N. R., et al. 2013, The Astrophysical Journal Letters, 777, L20 124, 62 Puget, P., Stadler, E., Doyon, R., et al. 2004, Ground-based Instrumentation for Astronomy, 5492, 978 Rajan, A., Rameau, J., De Rosa, R. J., et al. 2017, The Astronomical Journal, 154, 10, arXiv: 1705.03887 Lodieu, N., Béjar, V. J. S., & Rebolo, R. 2013, Astronomy and Rameau, J., Chauvin, G., Lagrange, A.-M., et al. 2013, Astronomy Astrophysics, 550, L2 and Astrophysics, 553, A60 WEIRD 55 Riedel, A. R., Finch, C. T., Henry, T. J., et al. 2014, The Torres, C. A. O., da Silva, L., Quast, G. R., de la Reza, R., & Astronomical Journal, 147, 85 Rodriguez, D. R., Bessell, M. S., Zuckerman, B., & Kastner, J. H. 2011, The Astrophysical Journal, 727, 62 Rodriguez, D. R., Zuckerman, B., Kastner, J. H., et al. 2013, The Astrophysical Journal, 774, 101 Schlieder, J. E., Lépine, S., & Simon, M. 2010, The Astronomical Journal, 140, 119 -- . 2012a, The Astronomical Journal, 143, 80 -- . 2012b, The Astronomical Journal, 144, 109 Schneider, G., Song, I., Zuckerman, B., et al. 2004, American Astronomical Society Meeting Abstracts, 11.14 Shkolnik, E., Liu, M. C., & Reid, I. N. 2009, The Astrophysical Journal, 699, 649 Shkolnik, E. L., Allers, K. N., Kraus, A. L., Liu, M. C., & Flagg, L. 2017, The Astronomical Journal, 154, 69 Shkolnik, E. L., Anglada-Escudé, G., Liu, M. C., et al. 2012, The Astrophysical Journal, 758, 56 Shkolnik, E. L., Liu, M. C., Reid, I. N., Dupuy, T., & Weinberger, A. J. 2011, The Astrophysical Journal, 727, 6 Jilinski, E. 2000, The Astronomical Journal, 120, 1410 Torres, C. A. O., Quast, G. R., Melo, C. H. F., & Sterzik, M. F. 2008, in Handbook of Star Forming Regions, Volume II, 757 Veras, D., Crepp, J. R., & Ford, E. B. 2009, The Astrophysical Journal, 696, 1600 Vigan, A., Patience, J., Marois, C., et al. 2012, Astronomy and Astrophysics, 544, A9 Vigan, A., Bonavita, M., Biller, B., et al. 2017, Astronomy and Astrophysics, 603, A3 Wahhaj, Z., Liu, M. C., Nielsen, E. L., et al. 2013, The Astrophysical Journal, 773, 179 Wang, J., Fischer, D. A., Xie, J.-W., & Ciardi, D. R. 2015, The Astrophysical Journal, 813, 130 Wertz, O., Absil, O., González, G., et al. 2017, Astronomy and Astrophysics, 598, A83 Wolf, C., Onken, C. A., Luvaul, L. C., et al. 2018, Publications of the Astronomical Society of Australia, 35, e010 Zuckerman, B. 2001, Young Stars Near Earth: Progress and Smart, R. L., Marocco, F., Caballero, J. A., et al. 2017, Monthly Prospects, 244, 122 Notices of the Royal Astronomical Society, 469, 401 Soderblom, D. R. 2010, Annual Review of Astronomy and Astrophysics, 48, 581 Zuckerman, B., Rhee, J. H., Song, I., & Bessell, M. S. 2011, The Astrophysical Journal, 732, 61 Zuckerman, B., & Song, I. 2004, Annual Review of Astronomy Song, I., Zuckerman, B., & Bessell, M. S. 2003, The Astrophysical and Astrophysics, 42, 685 Journal, 599, 342 Soummer, R., Pueyo, L., & Larkin, J. 2012, The Astrophysical Journal Letters, 755, L28 Soummer, R., Choquet, E., Pueyo, L., et al. 2016, American Astronomical Society Meeting Abstracts #227, 227, 137.03 Stamatellos, D., Hubber, D. A., & Whitworth, A. P. 2007, Monthly Notices of the Royal Astronomical Society, 382, L30 Tonry, J. L., Stubbs, C. W., Lykke, K. R., et al. 2012, The Astrophysical Journal, 750, 99 Zuckerman, B., Song, I., & Bessell, M. S. 2004, The Astrophysical Journal Letters, 613, L65 Zuckerman, B., Song, I., Bessell, M. S., & Webb, R. A. 2001a, The Astrophysical Journal Letters, 562, L87 Zuckerman, B., Song, I., & Webb, R. A. 2001b, The Astrophysical Journal, 559, 388 Zuckerman, B., & Webb, R. A. 2000, The Astrophysical Journal, 535, 959
1204.4390
1
1204
2012-04-19T15:54:44
Formation and long-term evolution of 3D vortices in protoplanetary discs
[ "astro-ph.EP", "astro-ph.SR" ]
In the context of planet formation, anticyclonic vortices have recently received lots of attention for the role they can play in planetesimals formation. Radial migration of intermediate size solids toward the central star may prevent their growth to larger solid grains. On the other hand, vortices can trap the dust and accelerate this growth, counteracting fast radial transport. Multiple effects have been shown to affect this scenario, such as vortex migration or decay. The aim of this paper is to study the formation of vortices by the Rossby wave instability and their long term evolution in a full three dimensional protoplanetary disc. We use a robust numerical scheme combined with adaptive mesh refinement in cylindrical coordinates, allowing to affordably compute long term 3D evolutions. We consider a full disc stratified both radially and vertically that is prone to formation of vortices by the Rossby wave instability. We show that the 3D Rossby vortices grow and survive over hundreds of years without migration. The localized overdensity which initiated the instability and vortex formation survives the growth of the Rossby wave instability for very long times. When the vortices are no longer sustained by the Rossby wave instability, their shape changes toward more elliptical vortices. This allows them to survive shear-driven destruction, but they may be prone to elliptical instability and slow decay. When the conditions for growing Rossby wave-related instabilities are maintained in the disc, large-scale vortices can survive over very long timescales and may be able to concentrate solids.
astro-ph.EP
astro-ph
Astronomy & Astrophysics manuscript no. arXiv January 25, 2020 c(cid:13) ESO 2020 Formation and long-term evolution of 3D vortices in protoplanetary discs H. Meheut1, R. Keppens2, F. Casse3 and W. Benz1 1 Physikalisches Institut & Center for Space and Habitability, Universitat Bern, 3012 Bern, Switzerland e-mail: [email protected] 2 Centre for Plasma Astrophysics, Department of Mathematics, KU Leuven, Celestijnenlaan 200B, 3001 Heverlee, Belgium 3 AstroParticule et Cosmologie (APC), Universite Paris Diderot, 10 rue A. Domon et L. Duquet, 75205 Paris Cedex 13, France 2 1 0 2 r p A 9 1 . ] P E h p - o r t s a [ 1 v 0 9 3 4 . 4 0 2 1 : v i X r a Preprint online version: January 25, 2020 ABSTRACT Context. In the context of planet formation, anticyclonic vortices have recently received lots of attention for the role they can play in planetesimals formation. Radial migration of intermediate size solids toward the central star may prevent their growth to larger solid grains. On the other hand, vortices can trap the dust and accelerate this growth, counteracting fast radial transport. Multiple effects have been shown to affect this scenario, such as vortex migration or decay. Aims. The aim of this paper is to study the formation of vortices by the Rossby wave instability and their long term evolution in a full three dimensional protoplanetary disc. Methods. We use a robust numerical scheme combined with adaptive mesh refinement in cylindrical coordinates, allowing to afford- ably compute long term 3D evolutions. We consider a full disc stratified both radially and vertically that is prone to formation of vortices by the Rossby wave instability. Results. We show that the 3D Rossby vortices grow and survive over hundreds of years without migration. The localized overdensity which initiated the instability and vortex formation survives the growth of the Rossby wave instability for very long times. When the vortices are no longer sustained by the Rossby wave instability, their shape changes toward more elliptical vortices. This allows them to survive shear-driven destruction, but they may be prone to elliptical instability and slow decay. Conclusions. When the conditions for growing Rossby wave-related instabilities are maintained in the disc, large-scale vortices can survive over very long timescales and may be able to concentrate solids. Key words. Planets and satellites: formation - Protoplanetary discs - Hydrodynamics - Instabilities - Accretion discs 1. Introduction The origin of the kilometer size planetesimals is still an issue of planet formation theory. The growth of micron size solids toward centimetre or meter blocks is predicted by coagulation models (Dominik 2009), but collisions usually lead to destruc- tion of such solids (Benz 2000). These collisions arise because the pressure supported gas moves at sub-Keplerian speed and the solid bodies experience a head wind from the gas due to the drag forces. The consequence is a loss of angular momentum and a radial drift of solids spiraling toward the central star. This can occur on a timescale as short as a hundred years for meter size blocks at one astronomical unit (Weidenschilling 1977). Such timescales are far too short to explain subsequent growth into larger particles that are unaffected by the head wind. To overcome these difficulties, anticyclonic vortices, where the velocity streamlines rotate in the opposite direction to the Keplerian flow (Marcus 1990), may be a good environment for planetesimal growth. They induce a drag force towards the centre of the structures, and are therefore proposed as possible nurseries for intermediate size blocks, concentrating the solids and accelerating the planetesimal formation processes (Barge & Sommeria 1995; Tanga et al. 1996; Bracco et al. 1999; Godon & Livio 1999, 2000; Johansen et al. 2004; Heng & Kenyon 2010). As mentioned by Armitage (2011), vortices can also be important for mechanisms that require an increase in the dust-to-gas ratio, such as the streaming instability (Johansen et al. 2009). Furthermore, Klahr & Bodenheimer (2006) argued that anticyclonic vortices are regions with lower turbulence. Fragmentation is less likely to occur if lower velocity fluctua- tions prevail. One of the main difficulties with this scenario is the question of the formation of vortices. Different instabilities have been proposed for the vortex generation, such as the baroclinic instability (Klahr & Bodenheimer 2003; Klahr 2004) in which interest has been recently revived (Lesur & Papaloizou 2010; Lyra & Klahr 2011), or potentially the magneto-rotational instability (Fromang & Nelson 2005). In this paper we will study the Rossby wave instability (RWI), that may form vortices with unusual elongated shape in regions of particular interest for the long term survival of those structures. It has been recently pointed out that the survival of the vortices on a sufficient timescale may be an issue. The elliptical instability may destroy three dimensional (3D) elliptical vortices (Lesur & Papaloizou 2009). Moreover, Paardekooper et al. (2010) have shown that vortices can be subject to radial migration toward the disc centre in a short timescale. In this paper, we investigate the generation, the 3D struc- ture and the long term evolution of vortices formed by the RWI 1 Meheut et al.: 3D vortices in protoplanetary discs the entire system. The pressure is then p = S ργ, with the adia- batic index γ = 5/3 and the constant S related to entropy. The s = γp/ρ = S γργ−1 and the tempera- sound speed is given by c2 ture by T ∼ p/ρ = S ργ−1. The temperature is normalised to the temperature at 1 AU. 2.3. Numerics The numerical methods used for these simulations are inspired by those of Meheut et al. (2010). We use the Message Passing Interface-Adaptive Mesh Refinement Versatile Advection Code (MPI-AMRVAC) developed by Keppens and Meliani (Keppens et al. 2012). The use of a code that allows the mesh to adapt dur- ing the simulation has multiple advantages. One aim is to reach higher resolution in the vortex regions, as well as to have a better resolution in the upper region of the disc, where the gas density decreases abruptly forming a 'corona'. The use of AMR offers flexibility to enforce a lower resolution in this physically irrel- evant, but computationally challenging, region. Since low den- sity material can easily be accelerated, in turn enforcing smaller timesteps to maintain numerical stability in explicit time step- ping schemes, handling the corona at low resolution is computa- tionally advantageous. The current AMR approach can therefore model the unstable disc on long timescales with similar comput- ing resources as the short timescale run done earlier. Also, by better numerically representing the disc-corona interface, com- bined with a sharper limiter function, we have a more robust numerical treatment of the governing equations. The numerical scheme is the same for all refinement levels, namely the Total Variation Diminishing Lax-Friedrich scheme (see T´oth & Odstrcil (1996)) with a third order accurate Koren limiter (Koren 1993) on the primitive variables. We use a cylin- drical grid with r  [1, 6] AU, z  [0, 0.5] AU and the full az- imuthal direction ϕ  [0, 2π]. The simulations consider only the upper half of the disc, as the disc mid-plane appeared to be a symmetry plane for the instability in our earlier full vertical sim- ulations (Meheut et al. 2010). If not specified, the length is given in AU and the code unit time corresponds to 1/(2π) yr. The res- olution of the base level is (64, 32, 32). Up to three levels of re- finement are allowed, corresponding to an effective resolution of (256, 128, 128). In the region of the density bump where the in- stability is expected to grow, the resolution is fixed to the higher level during the whole simulation. The initial grid is presented in Fig. 2, and then evolves to follow the growth of the instabil- ity. The refinement criterion being the density variation, the grid follows the growth of the spiral density waves propagating radi- ally. At the end of the simulation ∼ 70% of the grid volume is at the highest resolution level. In the beginning, it is only about ∼ 26%, as visually clear from Fig. 2. Since refinement happens in all directions, each grid block handled at the base resolution level represents a gain factor of 64 in computational cost com- pared to a uniform high resolution run. This gain is even more dramatic when memory issues are considered as well. During the entire run, the corona region is maintained at this lowest resolu- tion, dramatically impacting stability, memory and wall-clock time. The boundary conditions are transparent at inner and outer radius, we consider a mid-plane symmetry and transparent boundary conditions are also applied at the upper boundary, whereas the azimuthal direction is periodic. Fig. 1. Mid-plane density of the gas in g cm−3. The density bump is placed at 3 AU. within a protoplanetary disc. This follows the work presented in Meheut et al. (2010), revisited with a new code that allows to handle long term simulations. The numerical and physical setup is described in the next section. The results are presented in sec- tion 3, followed by a detailed discussion including comparisons with previous works. 2. Methods and setup 2.1. Rossby wave instability The RWI (Lovelace et al. 1999; Li et al. 2000, 2001) has been studied and discussed in various situations of differentially ro- tating discs, from galactic discs (Lovelace & Hohlfeld 1978; Sellwood & Kahn 1991) to microquasar and protoplanetary discs (Papaloizou & Pringle 1985; Lovelace et al. 1999). It can be seen as the form that the Kelvin-Helmholtz instability takes in differ- entially rotating disks, and has a similar instability criterion: It requires an extremum in a vorticity related quantity, defined in a non-magnetised thin disc as (Li et al. 2000): L = (pΣ−γ)2/γ , (1) ΣΩ κ2 (pΣ−γ)2/γ = Σ 2(∇ × v)z where Σ is the surface density, p the pressure, v is the velocity of the fluid, γ the adiabatic index, Ω the rotation frequency and κ2 = 4Ω2+2rΩΩ(cid:48) the squared epicyclic frequency (so that κ2/2Ω is the vorticity). Here the prime denotes a radial derivative. For the isentropic discs we consider here, this criterium is reduced to an extremum of ΣΩ κ2 . This quantity is directly related to vortensity defined as the ratio of vorticity to surface density. 2.2. Equations We work in cylindrical coordinates (r, z, ϕ) with the 3D Euler equations ∂tρ + ∇ · (vρ) = 0 , (2) ∂t(ρv) + ∇ · (vρv) + ∇p = −ρ∇ΦG , (3) where ρ is the mass density of the fluid, v its velocity, and p its pressure. ΦG = − GM∗√ r2+z2 is the gravity potential of the central object with G the gravitational constant and M∗ the mass of the star. We consider a barotropic flow, i.e. the entropy is constant in 2 Meheut et al.: 3D vortices in protoplanetary discs Fig. 2. Initial gas density in the disc in g cm−3 (up) and AMR grid (down). The three levels are plotted in white, blue and dark blue from lower to higher resolution. The highest resolution is reached in the overdensity region and at the interface between the disc and corona. 2.4. Initial conditions The initial conditions for the simulations are chosen to represent a protoplanetary disc in radial and vertical equilibrium. We con- sider that an overdensity is formed at some radius (rbump) of the disc. This bump could have been formed due to multiple effects. gas is then sub-Keplerian. There is no vertical velocity initially, whereas a very low radial velocity perturbation is added to the equilibrium as a seed for the instability. These are random per- turbations, meaning that all the spatial frequencies in the three directions are present in the seed. 1. The presence of a dead zone in the protoplanetary disc, cor- responding to a region of lower ionisation and resistivity, leads to a lower accretion rate in this region (Gammie 1996). This can induce the formation of an overdensity at the edges of this region (Varni`ere & Tagger 2006; Lyra et al. 2008; Kretke et al. 2009), which can in turn trigger the RWI. The initial density bump used in this paper, is a Gaussian fit to the one obtained in Varni`ere & Tagger (2006), when the in- stability started to grow. 2. The ice sublimation front ('snow line') can also be responsi- ble for the formation of an overdensity (Kretke & Lin 2007). One can note that an extremum of entropy should be located in this region, which may also trigger the RWI (Lovelace et al. 1999). 3. It has also been shown that the RWI grows at the edge of planet gaps (Koller et al. 2003; Lin & Papaloizou 2011). The radial positions of these regions depend on the physical characteristics of the disc (such as accretion rate and tempera- ture) or the position of the planets. We have chosen to place the bump at a distance of 3 AU which is a plausible region for these different effects. The initial mid-plane density is given by (cid:18) (cid:17)2(cid:19) (cid:16) r − rB√ 2σ , 1 + χ exp ρ(r, 0, ϕ) = ρ0rα (4) with ρ0 = 3.10−9g cm−3 the density at r = 1 AU, and α = −1.5 the power law index of the underlying density. This value for the alpha parameter gives a surface density varying approximately as Σ ∼ r−1/2 in the absence of the bump. Parameters rB = 3 AU, χ = 1, σ = 0.1 AU are the radius, amplitude and width of the Gaussian bump. The vertical and radial force equilibria give the initial verti- cal structure in density and azimuthal velocity, respectively. The pressure is calculated with constant entropy, with S = 10−3. The mid-plane and vertical density profile chosen as initial condition are shown on Fig. 1 and Fig. 2. The azimuthal velocity is close to Keplerian where the de- viation from Keplerian rotation is due to the pressure gradient. In most of the disc, except in the inner part of the bump, the 3. Results The evolution of the pressure bump in a protoplanetary disc is followed over more than 600 years (tmax = 4000). This timescale corresponds to ∼ 123 rotations of the overdensity and ∼ 43 ro- tations at the outer edge of the grid. Such a configuration allows the development of the Rossby wave instability and we present here the characteristics of the 3D flow under this instability and its non-linear evolution. 3.1. Growth of the RWI The RWI is characterised by vortex waves in the region of the pressure bump and spiral density waves propagating on each side of the bump. These features can be seen on Fig. 3 where the vortex waves can be identified by the vortensity extrema or the velocity streamlines characteristic for such waves. On the upper plot, the inner and outer Rossby waves are visible. Spiral density waves are emitted from each vortex inward and outward (e.g. at t = 300). The growth of the instability is quantified in Fig. 4 and 5. The linear phase of the instability corresponds to exponential growth of the perturbations until t ∼ 300 when the density perturbation reaches ∼ 20% of the initial density. The straight line corresponds to a linear fit of the amplitude growth of the density perturbation in logarithmic scale. This fit gives a growth rate of γ = 0.033Ω0 in code units which corresponds to 0.17Ω(rB) or 0.21 yr−1. We have also plotted the amplitude evo- lution of selected azimuthal modes, out of the Fourier transform of the density ρ(r, z, ϕ, t) = ρm(r, z, t) exp(−imϕ). (5) (cid:88) m The most unstable mode during the exponential growth has az- imuthal mode number m = 5, but lower azimuthal mode num- bers are then excited by the m = 5 mode leading to their growth, with m = 1 and m = 2 shown on the figure. The mode number corresponds, on Fig. 3, to the number of anticyclonic vortices, which are those counter-rotating the Keplerian rotation. They correspond to high pressure and negative vortensity regions. A high number of anticyclonic vortices come out from the initial random perturbations, rapidly decreasing to 5. 3 Meheut et al.: 3D vortices in protoplanetary discs Fig. 3. Density (left), perturbation of density and velocity streamlines (center), perturbed vortensity defined as ζ − (cid:104)ζ(cid:105)ϕ with ζ = (∇ × v)z/ρ (right) at t = 100, 300, 800 from top to bottom. 3.2. Saturation of the RWI After the exponential growth, the instability reaches saturation due to non-linearities inducing mode mixing. Fig. 6 shows the amplitude of the modes at saturation. The fifth mode clearly dominates for quite some time, and modes with high azimuthal mode number are less prominent, but very low and intermediate azimuthal mode numbers are non-linearly excited by the prevail- ing mode. However one can see on Fig. 3 that later on (t = 800) m = 2 dominates, and then m = 1 until the end of the sim- ulation (see Fig. 6). On the other hand, the higher mode num- bers (5 < m < 10) are not dominant after saturation. So the general sketch of the instability is first the linear growth of the fifth mode, then saturation is reached and the larger wavelength modes (lower mode number) dominate with a transfer towards larger scales. This evolution diminishes the number of vortices through merging as described in Godon & Livio (1999). This is similar to the 2D simulations of Baty et al. (2003) where pla- nar shear layers were studied in high resolution Cartesian set- tings. They found that the pairing/merging of vortices in Kelvin- Helmholtz unstable shear layers, studied from hydro to magne- tized cases, is controlled by growth of subharmonic modes and phase differences between them. In the present 3D cylindrical configurations, their results represent local box evolutions, per- formed in the frame rotating at the local Keplerian speed. This pairing/merging trend to large scale structure is then predomi- nantly a 2D process known from planar hydrodynamical evo- lutions. As the RWI is closely related to the Kelvin-Helmholtz instability, the merging of vortices seen in our 3D, thin disc sim- ulations is consistent with the Baty et al. (2003) findings. 3.3. Decay of the vortices The simulation has been run over more than 600 years to inves- tigate the long term evolution of the vortices. After the growing phase, the vortices survive during hundreds of years but then start to decrease and have disappeared at the end of the simula- tion. After a few hundreds of inner rotations, the dominant az- imuthal mode does not change anymore, and has a global m = 1 value. This corresponds to the largest scale possible. One can see on Fig. 7 that during the decaying phase (after about 100- 150 years), vertical mode structure appears in the m = 5 mode. This structure is present inside the vortex (r = 3) but not in the outer region (r = 3.5). One can notice that the spiral density waves disappear af- ter the growing phase when the vortices survive. This implies that Rossby wave instability is not active anymore due to the weakening of the bump by the instability itself and no angular 4 Meheut et al.: 3D vortices in protoplanetary discs Fig. 5. Time evolution of enstrophy in a logarithmic scale. The time is given in code units (lower axis) and in years (upper axis). Fig. 4. Time evolution of the amplitudes of the density pertur- bations in a logarithmic scale. The time is given in code units (lower axis) and in years (upper axis). We also show the am- plitude of some of the modes with low mode number. The upper figure is a zoom on the exponential growth where we have added a linear fit for the dominant mode whose growth rate is 0.033 in code units. momentum is transferred radially. This is quantified in Fig. 8, where the radial accretion rate is plotted: the bump stops to de- crease due to radial accretion. When the Rossby vortices are not anymore sustained by the RWI, they should be very efficiently destroyed by differential rotation (Tagger 2001), but the shape of the streamlines changes to become more similar to closed el- liptical streamlines as presented in Fig. 9. The vortex streamlines are not directly linked to the shearing sub-Keplerian streamlines, and can survive over more than one rotation time. By the end of the simulation, the vortices have been completely destroyed. This will be discussed in the following section. Fig. 6. Comparison of the amplitude of the first 10 azimuthal modes between t = 0 and t = 400 (left) and over the whole simulation (right). The instability is dominated by the mode m = 5 but both lower and higher azimuthal modes are excited during the growth phase. The decaying phase shows a shift to lower azimuthal mode number. 4. Discussion 4.1. Comparison with previous 3D simulations From a numerical point of view, the main difference with the pre- vious simulation of Meheut et al. (2010) is the use of an AMR grid. This allowed to increase the resolution at the interface be- tween the disk and the corona above it and to avoid numerical difficulties due to dynamics in this low density region. Our cur- rent simulation uses an improved limiter and benefits from grid adaptivity, allowing to robustly compute and follow the growth of the instability over longer times. The limit in time is now re- lated to computational resources. We tested simulations of the initial growth of the RWI with and without AMR, getting near- identical results. 5 Meheut et al.: 3D vortices in protoplanetary discs Fig. 7. Vertical structure in density of the dominant fifth az- imuthal mode. The amplitude of the mode is normalised to its value in z = 0, so its structure at different times and positions can be compared. The accretion rate is defined as −(cid:82) (cid:82) Fig. 8. Accretion as a function of radius during the exponential growth (solid line) and at the end of the simulation (dashed line). dϕdzrρvr, and the ampli- tude is given in code units which correspond to ∼ 10−4M(cid:12)yr−1. Fig. 9. Vortex streamline at t = 300 and t = 1000: the shape of the vortex evolves during the simulation. The radial extent of the vortices is constant and fixed by the radial extent of the density bump, whereas their azimuthal extent increases when the m mode number decreases. On the right figure, the streamlines become elliptic and an aspect ratio can be estimated (∼ 6 here, but each vortex has a different aspect ratio as can be seen on Fig. 3). 4.2. Vortex migration We don't see any migration of the vortices in our simulation over more than 600 years. This result is expected, since the growth mechanism of these vortices is localised in the bump region. Vortex migration is oriented in the direction of positive density gradient, and as a consequence the presence of a density max- imum at the location of the vortices locks their radial position. One can see on Fig. 11 that the overdensity is diminished by the Rossby wave instability but it survives the growth of the RWI. The vortices are then indefinitely blocked in this region. The simulations presented in this paper show that the most unstable mode of the RWI in the current disc equilibrium has an azimuthal mode number m = 5. In Meheut et al. (2010) we stressed the global m = 1 mode. There, only one mode was present in the seed perturbation, whereas we now choose to use random perturbations that include all modes. For confirmation, we have run the simulation of Meheut et al. (2010) with ran- dom initial perturbations, and the instability was dominated by the m = 5 mode. The difference with our earlier work is then not due to disparity in initial equilibrium or numerical approach. This result is also coherent with linear calculations of Meheut et al. (2012). The unusual 3D structure of the Rossby vortices with an non- negligible vertical velocity is confirmed by these higher resolu- tion simulations, as can be seen on Fig. 10. We have selected some streamlines which pass next to the center of the vortices, they show the 'eye' of the vortices and the larger basis. The di- rection of the flow in the eye of the vortices differs between the cyclones (downward flow) and anticyclones (upward flow). In the outskirts of the vortices, the flows have opposite directions. 4.3. Elliptical instability The elliptical instability is a purely 3D instability that is due to resonances between the vortex turnover frequency and an iner- tial wave frequency (Kerswell 2002). Lesur & Papaloizou (2009) have shown that 3D elliptical vortices are vulnerable to this in- stability, and can destroy them. In our simulation, the elliptical instability does not prevent the emergence of vortices by the RWI and their survival over hundreds of years. This was ex- pected as closed velocity streamlines are needed for the turnover frequency to be properly defined and the elliptical instability to grow. Some streamlines of the (non-steady) Rossby vortices can be seen on Fig. 9 (see also Fig. 10). After saturation of the RWI, the streamlines tend to close, on the one hand this allows them to survive against shearing, but on the other hand, they become prone to the elliptical instability. This may be the reason for the decay of the vortices. The growth of this instability can ex- plain the appearance of smaller scale vertical perturbations (as the ones plotted in Fig. 7) whereas the azimuthal scale is fixed. Indeed contrary to the RWI, the elliptical instability is a local in- stability without any interaction between the vortices, it can not 6 Meheut et al.: 3D vortices in protoplanetary discs Fig. 11. Mean azimuthal density in the mid-plane of the bump region, initially and at t = 1000. The bump has survived the growth of the RWI. buoyant, e.g. subject to a subcritical baroclinic instability, and Lesur & Papaloizou (2010) have shown with higher resolution simulations that this secondary instability does not fully destroy the vortices (see also Lyra & Klahr 2011). 4.4. Outlooks Fig. 10. Some streamlines which pass next to the centre of the anticyclone (upper figure) and cyclone (lower figure) at t = 300 are plotted in 3D. The direction of the flow is given by the colors: the flow moves from the blue part to the red part. The cyclone is a downward flow, contrary to the upward anticyclonic flow. The vortices have a coherent structure over the whole height of the disc. The radial extent of this figure corresponds approximately to half of the one of Fig.9: only the inner part of the vortices are shown here. change their number but affects only their structure. The detailed study of the decay of the vortices by the elliptical instability, in- cluding elongated vortices or short wavelengths, needs a very high resolution as has been done in Lesur & Papaloizou (2009). They studied the effects of the aspect ratio, the ratio of semi- major to semi-minor axis of elliptical streamlines, and showed that elliptical vortices are always unstable. However, the ellip- tical instability becomes weak for large aspect ratios (i.e. elon- gated vortices, as is the case in Fig. 9). Therefore, the low az- imuthal mode number vortices are expected to survive longer, as observed in our simulations. As pointed out by Godon & Livio (1999), we must caution that the inherently higher numer- ical dissipation of our finite volume method compared to a spec- tral method, also induces vortex decay. The elliptical instability is also present in the vortices formed when the disk is radially We have shown that Rossby vortices can emerge in a 3D pro- toplanetary disc with an overdensity, and that they can survive for hundreds of years without migrating. When they are not sus- tained anymore by the RWI, the structure of the vortices will evolve to previously studied vortices in isolation (Godon & Livio 1999; Lesur & Papaloizou 2009). They are then expected to be destroyed and may be prone to the elliptical instability and vis- cous decay. However, our simulations show that the Rossby vor- tices can survive for long time scales, if they keep being sus- tained by the RWI and their unusual structure is preserved. That would be the case if the mechanism that forms the overdensity and launch the RWI, is permanent, such as present in a disc with both active and dead zone regions. A point to explore is the study of this instability when the ac- tive zone is also considered. On the one hand we would expect viscosity to decrease the growth rate of the instability and even- tually to destroy the vortices. On the other hand, if a dead zone is included, a density bump arises that will in turn sustain the instability. A full MHD simulation could allow to study jointly the bump formation process and its decay due to the RWI, in combination with magneto-rotational driven accretion. The results presented here are especially interesting in the context of planetesimal formation. Future work should also study the joint evolution of the gas and dust particles to ques- tion the influence of the dust particles on the growth of the RWI and their concentration in Rossby vortices. Previous studies have proposed a bi-dimensional approach (Chang & Oishi 2010; Johansen et al. 2004; Wolf & Klahr 2005) but a full 3D study, as the one presented here including solid particles, will be nec- essary to handle both vertical stratification and vortex concen- tration. With this intent, a module of MPI-AMRVAC has been developed and was recently used for circumstellar wind model- ing (van Marle et al. 2011). 7 Meheut et al.: 3D vortices in protoplanetary discs On an analytical point of view, the only work we are aware of, that deals with the RWI in 3D is the one by Umurhan (2010). The vertical structure of the modes should be studied in more detail with dedicated tools as the one presented in Zhang & Lai (2006). This will be addressed in Meheut et al. (2012). Acknowledgements. This work was partially supported by the Tournesol pro- gram of the PHC (Partenariat Hubert Curien) and by the Swiss National Science Foundation. References Armitage, P. J. 2011, ARA&A, 49, 195 Barge, P. & Sommeria, J. 1995, A&A, 295, L1 Baty, H., Keppens, R., & Comte, P. 2003, Physics of Plasmas, 10, 4661 Benz, W. 2000, Space Sci. Rev., 92, 279 Bracco, A., Chavanis, P. H., Provenzale, A., & Spiegel, E. A. 1999, Physics of Fluids, 11, 2280 Chang, P. & Oishi, J. S. 2010, ApJ, 721, 1593 Dominik, C. 2009, in Astronomical Society of the Pacific Conference Series, Vol. 414, Cosmic Dust - Near and Far, ed. T. Henning, E. Grun, & J. Steinacker, 494 -- + Fromang, S. & Nelson, R. P. 2005, MNRAS, 364, L81 Gammie, C. F. 1996, ApJ, 457, 355 Godon, P. & Livio, M. 1999, ApJ, 523, 350 Godon, P. & Livio, M. 2000, ApJ, 537, 396 Heng, K. & Kenyon, S. J. 2010, MNRAS, 408, 1476 Johansen, A., Andersen, A. C., & Brandenburg, A. 2004, A&A, 417, 361 Johansen, A., Youdin, A., & Mac Low, M.-M. 2009, ApJ, 704, L75 Keppens, R., Meliani, Z., van Marle, A., et al. 2012, Journal of Computational Physics, 231, 718 , special Issue: Computational Plasma Physics Kerswell, R. R. 2002, Annual Review of Fluid Mechanics, 34, 83 Klahr, H. 2004, ApJ, 606, 1070 Klahr, H. & Bodenheimer, P. 2006, ApJ, 639, 432 Klahr, H. H. & Bodenheimer, P. 2003, ApJ, 582, 869 Koller, J., Li, H., & Lin, D. N. C. 2003, ApJ, 596, L91 Koren, B. 1993, A robust upwind discretization method for advection, diffusion and source terms, Vol. 45, Notes on numerical fluid mechanics, ed. C. B. Vreugdenhil & B. Koren, 117 Kretke, K. A. & Lin, D. N. C. 2007, ApJ, 664, L55 Kretke, K. A., Lin, D. N. C., Garaud, P., & Turner, N. J. 2009, ApJ, 690, 407 Lesur, G. & Papaloizou, J. C. B. 2009, A&A, 498, 1 Lesur, G. & Papaloizou, J. C. B. 2010, A&A, 513, A60+ Li, H., Colgate, S. A., Wendroff, B., & Liska, R. 2001, ApJ, 551, 874 Li, H., Finn, J. M., Lovelace, R. V. E., & Colgate, S. A. 2000, ApJ, 533, 1023 Lin, M.-K. & Papaloizou, J. C. B. 2011, MNRAS, 415, 1426 Lovelace, R. V. E. & Hohlfeld, R. G. 1978, ApJ, 221, 51 Lovelace, R. V. E., Li, H., Colgate, S. A., & Nelson, A. F. 1999, ApJ, 513, 805 Lyra, W., Johansen, A., Klahr, H., & Piskunov, N. 2008, A&A, 491, L41 Lyra, W. & Klahr, H. 2011, A&A, 527, A138+ Marcus, P. S. 1990, Journal of Fluid Mechanics, 215, 393 Meheut, H., Casse, F., Varniere, P., & Tagger, M. 2010, A&A, 516, A31+ Meheut, H., Yu, C., & Lai, D. 2012, MNRAS, 2748 Paardekooper, S., Lesur, G., & Papaloizou, J. C. B. 2010, ApJ Papaloizou, J. C. B. & Pringle, J. E. 1985, MNRAS, 213, 799 Sellwood, J. A. & Kahn, F. D. 1991, MNRAS, 250, 278 Tagger, M. 2001, A&A, 380, 750 Tanga, P., Babiano, A., Dubrulle, B., & Provenzale, A. 1996, Icarus, 121, 158 T´oth, G. & Odstrcil, D. 1996, J. Comput. Phys., 128, 82 Umurhan, O. M. 2010, A&A, 521, A25+ van Marle, A. J., Meliani, Z., Keppens, R., & Decin, L. 2011, ApJ, 734, L26+ Varni`ere, P. & Tagger, M. 2006, A&A, 446, L13 Weidenschilling, S. J. 1977, MNRAS, 180, 57 Wolf, S. & Klahr, H. 2005, in ESA Special Publication, Vol. 577, ESA Special Publication, ed. A. Wilson, 473 -- 474 Zhang, H. & Lai, D. 2006, MNRAS, 368, 917 8
1112.0359
1
1112
2011-12-01T23:42:05
The Frequency of Hot Jupiters in the Galaxy: Results from the SuperLupus Survey
[ "astro-ph.EP" ]
We present the results of the SuperLupus Survey for transiting hot Jupiter planets, which monitored a single Galactic disk field spanning 0.66 sq. deg for 108 nights over three years. Ten candidates were detected: one is a transiting planet, two remain candidates, and seven have been subsequently identified as false positives. We construct a new image quality metric, S_j, based on the behaviour of 26,859 light curves, which allows us to discard poor images in an objective and quantitative manner. Furthermore, in some cases we are able to identify statistical false positives by analysing temporal correlations between S_j and transit signatures. We use Monte Carlo simulations to measure our detection efficiency by injecting artificial transits onto real light curves and applying identical selection criteria as used in our survey. We find at 90% confidence level that 0.10 (+0.27/-0.08)% of dwarf stars host a hot Jupiter with a period of 1-10 days. Our results are consistent with other transit surveys, but appear consistently lower than the hot Jupiter frequencies reported from radial velocity surveys, a difference we attribute, at least in part, to the difference in stellar populations probed. In light of our determination of the frequency of hot Jupiters in Galactic field stars, previous null results for transiting planets in open cluster and globular cluster surveys no longer appear anomalously low.
astro-ph.EP
astro-ph
The Frequency of Hot Jupiters in the Galaxy: Results from the SuperLupus Survey Daniel D. R. Bayliss1 [email protected] Penny D. Sackett1 ABSTRACT We present the results of the SuperLupus Survey for transiting hot Jupiter planets, which monitored a single Galactic disk field spanning 0.66 deg2 for 108 nights over three years. Ten candidates were detected: one is a transiting planet, two remain candidates, and seven have been subsequently identified as false pos- itives. We construct a new image quality metric, Sj, based on the behaviour of 26,859 light curves, which allows us to discard poor images in an objective and quantitative manner. Furthermore, in some cases we are able to identify statis- tical false positives by analysing temporal correlations between Sj and transit signatures. We use Monte Carlo simulations to measure our detection efficiency by injecting artificial transits onto real light curves and applying identical se- lection criteria as used in our survey. We find at 90% confidence level that 0.10+0.27 −0.08% of dwarf stars host a hot Jupiter with a period of 1-10 days. Our re- sults are consistent with other transit surveys, but appear consistently lower than the hot Jupiter frequencies reported from radial velocity surveys, a difference we attribute, at least in part, to the difference in stellar populations probed. In light of our determination of the frequency of hot Jupiters in Galactic field stars, previous null results for transiting planets in open cluster and globular cluster surveys no longer appear anomalously low. Subject headings: Planets and satellites: detection - Techniques: photometric 1Research School of Astronomy and Astrophysics, The Australian National University, Mt Stromlo Ob- servatory, Cotter Rd, Weston Creek, ACT 2611, Australia -- 2 -- 1. Introduction The discovery of short-period, giant extrasolar planets (Mayor & Queloz 1995; Marcy & Butler 1996) provided the exciting potential for large numbers of planets to be discov- ered by the transit method, as these "hot Jupiters" have a ∼10% geometric probability of transiting, and do so every few days. However early predictions greatly overestimated the actual discovery rate (Horne 2003). This discrepancy resulted from simplistic assumptions and a misunderstanding of the effects that systematic noise would play in lowering detec- tion efficiency (Pont et al. 2006). It also resulted from an over-estimation of the frequency of hot Jupiters. More accurate, and lower, predictions were provided in Beatty & Gaudi (2008), where it was noted that objective and quantifiable detection criteria were required for more robust inferences of planet frequencies. By adopting such objective and quantifi- able detection criteria, we propose that the frequency of hot Jupiters in the field has been overestimated by a factor of three (Bayliss & Sackett 2011). The SuperLupus Survey was established to detect hot Jupiters in a field positioned just above the Galactic plane (b = 11◦), and also to determine the fraction of stars that host hot Jupiters in this typical Galactic field. The survey design and data analysis was constructed so as to fulfill both of these objectives. In Section 2 of this paper, we describe the SuperLupus observations and data reduction. The photometry is described in Section 3. The criteria for candidate selection are detailed in Section 4, and the 10 identified candidates are described and analyzed in Section 5. In Section 6, we set out the details of the Monte Carlo simulations performed to calculate the detection efficiency and the effective number of stars probed for planets in the SuperLupus Survey. In Section 7, we apply this efficiency to the actual results of the survey to determine the fraction of stars in the field that host a hot Jupiter. Finally, in Section 8 we summarize and discuss the implications of our results. 2. Observations and Data Reduction The SuperLupus Survey is an extension of the Lupus Survey (Bayliss et al. 2009), in which the duration of the original survey was approximately doubled to provide sensitivity to transiting planets with periods as long as 10 days. Many details of the data reduction procedure are set out elsewhere (Bayliss et al. 2009); here we summarize the procedure and highlight aspects that differed between the Lupus and SuperLupus projects. Both surveys monitored a single field using the Wide-Field Imager on the ANU 40- -- 3 -- Inch Telescope at Siding Spring Observatory. The field, in the constellation of Lupus, is 11◦ above the Galactic plane centered at R.A.=15h30m36.3s, Decl.=−42◦53′53.0′′ (J2000). In total, 5158 images of 300 s exposure were taken in a custom (V +R) filter. Of these, 2801 images were from the original Lupus Survey (2005 and 2006) and an additional 2357 new images were taken in 2008. The average full-width half-maximum of the point spread function (PSF) over all 5158 images is 2.02′′. Initial data reduction (bias, dark, and flatfield correction) was carried out on the mosaic frames using standard IRAF1 tasks in the package MSCRED. Sky flats were obtained at twilight whenever conditions permitted, and master flatfields were produced by median com- bining the ∼30 sky flats most proximate in time to each night. Bad pixels and columns were masked and the mosaic frames were split into individual CCD frames. A world coordinate system solution was calculated for each image in order to identify stars for photometry. 3. Photometry and Detrending Photometry for the original Lupus Survey was performed using Difference Imaging Anal- ysis (DIA: Alard & Lupton 1998; Wozniak 2000). For the SuperLupus Survey, we instead used Source Extractor (Bertin & Arnouts 1996) aperture photometry. This approach was motivated by the fact that many of the survey images have PSFs that are asymmetric due to tracking or focusing issues. Such PSFs can cause potential problems for DIA photometry, but are not of concern for aperture photometry as long as the photometric aperture is large enough. Initial tests revealed that aperture photometry could achieve the same (and even better) photometric precision for the majority of the brighter stars (V < 20) in the field. In fact we note that the final SuperLupus aperture photometry resulted in approximately 10% more stars with a root-mean-squared (RMS) variability less than 0.025 mag compared with the original Lupus Survey using DIA. We found that a photometric aperture radius of 10 pixels (3.75′′) resulted in the highest precision lightcurves for the majority of these stars. A reference catalog of 50,907 target stars (V < 20) was produced from a single, high quality, reference image. Aperture photometry was performed on these stars for each survey image, with apertures re-centered on the point of peak flux for each star in each image. 1IRAF is distributed by the National Optical Astronomy Observatories, which are operated by the As- sociation of Universities for Research in Astronomy, Inc., under cooperative agreement with the National Science Foundation. -- 4 -- Background subtraction was performed using a background map produced for each image by median filtering and bi-cubic-spline interpolation over a 8×8 pixel grid. Systematic trends in the resulting lightcurves were removed using the Sys-Rem algorithm (Tamuz et al. 2005) with 12 iterations. To further refine this dataset, we removed stars with lightcurves that displayed RMS variability greater than 0.05 mag. We also discarded poor quality images as described below. In total this left us with 26,859 stars over 3585 images. 3.1. Image Quality Metric To identify poor quality images, we define an image quality metric (Sj) as: j = X S 2 i r2 ij σ2 ij , (1) where for the ith star on the jth image, rij is the average subtracted residual magnitude and σij is the photometric uncertainty. This metric is computed in the Sys-Rem algorithm, where linear systematic effects are removed in order to minimise Sj (Tamuz et al. 2005). In essence, Sj is a measure of how all the stars in image j vary from the mean of all images in the survey. We retained only those images with Sj < 0.02 for searching for planetary transits. The advantage of using this metric is that it uses information from all stars over the chip in a single quality factor that is directly related to the precision of the resulting stellar photometry. This image quality metric also proved useful to determine if transit candidates were statistical false positives by looking for correlations between "transit" events and Sj (see Section 5). 4. Transit Search and Detection Criteria In order to accurately determine the efficiency of the SuperLupus Survey, candidates must be identified in an automated and systematic manner that can be applied identically to synthetic lightcurves in the Monte Carlo simulation (see Section 6). This was implemented by way of a set of six selection criteria that were automatically applied to each processed lightcurve. Initially all 26,859 SuperLupus lightcurves were searched for transit events using the Box-fitting Least-Squares (BLS) algorithm (Kov´acs et al. 2002). Trial periods in the range -- 5 -- 1.01 < P < 10 d were tested with 55,000 equally-spaced frequency steps and 200 phase bins per trial frequency. Candidates were then identified based on their BLS Signal Detection Efficiency (SDE), as defined in Kov´acs et al. (2002). After testing synthetic lightcurves with the same temporal and noise characteristics as our data, a candidate threshold value of SDE = 7.0 was adopted. We defined the effective signal-to-noise of the detected transit event, α, for all candi- dates, as: α = √nobs , δ σ (2) where δ is the BLS-determined transit depth, σ is the RMS variability of the entire (Sys-Rem corrected) lightcurve, and nobs is the actual number of data points within the BLS-determined transit event, i.e. the number of "in-transit" data points. A threshold of α = 10 was adopted, identical to the threshold adopted in two other transit surveys of faint stars (Burke et al. 2006; Hartman et al. 2009). At this stage we also rejected candidates for which more than 90% of in-transit data originated from an event on a single night. As with other ground-based surveys, many of our candidates displayed integer-day pe- riods resulting from nightly systematic effects. Therefore candidates with 1.01 < P < 1.02 and 1.98 < P < 2.02 were rejected. We also required that candidates had a magnitude of V < 18.8. Finally candidates were cross-matched with the catalog of known variable stars in the field (Weldrake & Bayliss 2008) and eclipsing binary systems were removed. The selection criteria are summarized in Table 1. 5. Initial Candidates A total of 10 lightcurves fulfilled the six selection criteria set out in Section 4. Their characteristics are described in Table 2. The candidate SL-07 was immediately recognised as Lupus-TR-3, a star with a transiting planet that had been discovered during the original Lupus Survey (Weldrake et al. 2008; Bayliss et al. 2009). To investigate the likely nature of the other candidates, we used the ηp diagnostic method proposed by Tingley & Sackett (2005). The results are set out in Table 3. Candidates SL-02 and SL-04 show very high diagnostic numbers, indicating that their transit parameters make them unlikely to be real transiting planets. In the case of candidate SL-04, a very clear transit was observed on 6 June 2005, with a transit time of approximately 6 hr, in contrast to the BLS determined duration of 8.08 hr. Use of the shorter duration reduces -- 6 -- the ηp diagnostic value to 1.9, which, while still high, could be explained by a large radius planet, since the ηp diagnostic is normalized to the assumption RP = 1RJ . To help identify statistical false positives, we checked if the detected transit event was correlated with the image quality, using the calculated value of the metric Sj, as defined in Eq. (1). Three candidates, SL-08, SL-09 and SL-10, were found to have transit events that were strongly correlated with Sj, indicating that the detected transit signature is almost certainly a systematic effect that the Sys-Rem algorithm had not fully removed. An example of such a correlation is shown in Figure 1. We inspected the images of the candidates for signs that the transit signature might be caused by proximity to the edge of a CCD, a bad column, or a bright star. We also inspected the image for evidence of significant flux contributions from neighboring stars within the photometric aperture. SL-01 was found to be blended with a very bright, saturated neighboring star at a distance of 27′′. We therefore ruled this out as being a genuine candidate. We found that the candidate SL-03 had three neighbors close to the photometric aperture, and that SL-05 had a neighbor well within the photometric aperture. Photometry for these candidates was re-derived using a smaller photometric aperture to avoid the contaminating neighbors. This analysis revealed that the transit event seen in SL-05 was in fact due to a very deep eclipse (40%) occurring in a faint neighbour situated just 2.6′′ from SL-05. It also revealed that while the SL-03 target star was responsible for the observed transit event, close neighbors and the large aperture of the original photometry had diluted the size of the dip from the true depth of 0.100 mag to a shallower 0.042 mag. The true depth is characteristic of an eclipsing binary and does not pass our selection criterion. In summary, we conclude that of the 10 candidates identified in the SuperLupus Survey, only SL-07 can be confirmed as a genuine transiting planet. We rule out seven candidates (SL-01, SL-02, SL-03, SL-05, SL-08, SL-09, SL-10) as being false positives, while two candi- dates (SL-04 and SL-06) remain. These two remaining candidates are V=17.8 and V=18.6 respectively, making follow-up extremely difficult. Consequently, we have not observed these candidates further, although formally we cannot exclude them. In Section 7 we calculate hot Jupiter frequencies assuming only SL-07 is a transiting planet, but also provide figures for the case where one of these remaining candidates is a transiting planet. -- 7 -- g a M ) R + V ( δ j S -0.20 -0.10 0.00 0.10 -0.01 0.00 0.01 0.02 0.03 0.04 -0.10 -0.05 0.05 0.10 0.00 Phase Fig. 1. -- Phase-wrapped lightcurve for candidate SL-08 (top panel) and the corresponding image quality, Sj, for images with Sj > 0.005 (bottom panel) wrapped at the same period and phase. The correlation between the transit event and image quality indicates that SL-08 is a statistical false positive. -- 8 -- 6. Monte Carlo Simulations to Determine Detection Efficiency To determine the efficiency of the SuperLupus Survey to detecting transiting planets, we carried out a Monte Carlo simulation of the SuperLupus Survey. Synthetic transit signatures were inserted into actual SuperLupus lightcurves and then recovery attempted using the same methodology and selection criteria used for our survey. 6.1. Modeling the Stellar Population Generating a realistic set of synthetic transit lightcurves against which to test a detection algorithm requires knowledge of the actual distribution of transit depths, durations, and periods of transiting planets. Ideally, this would be achieved by determining the radius of each monitored star in the field. This is possible when monitoring an equidistant population such as an open cluster or a globular cluster (e.g. Hartman et al. 2009; Weldrake et al. 2005). Since the distances to field stars cannot be determined from photometry alone, the exact radius of each star in the SuperLupus field is unknown. We therefore used the Beason¸con model of the Galaxy (Robin et al. 2003) to provide a statistical distribution of stellar radii and masses for the stars monitored in our field. The population was synthesized over a 0.66 deg2 field centered at b = 11◦, l = 331◦. Interstellar extinction was set at 0.6 mag kpc−1, based on the Schlegel dust maps of the region (Schlegel et al. 1998). Stars were selected in the magnitude range 14 < V < 18.8 to match the SuperLupus observational and selection constraints. The Beason¸con model returned a population of stars that closely matched both the total star count, and the distribution of apparent magnitudes (see Figure 2), giving us confidence this model accurately simulates our field. We note that the Beason¸con model indicated that 24% of the survey stars have log g < 4.0. In our Monte Carlo these stars are classified as giants and given zero probability for transit detection. We assign a radius to each star in the simulation based on the stellar mass from the Beasan¸con model and the simple relationship given by Cox (2000): log R R⊙ = 0.917 log M M⊙ − 0.020 . (3) 6.2. Lightcurve Generation Lightcurves for the Monte Carlo were created by first randomly selecting a star from the Beasan¸con model population. The temporal sampling of the star is set to the actual t n u o c r a t s 4000 3500 3000 2500 2000 1500 1000 500 0 14 -- 9 -- 15 16 17 18 V Mag Fig. 2. -- Comparison of the apparent V magnitude distribution between the Beason¸con model (dashed line) and our observed field (solid line). timestamps of the SuperLupus Survey. We then need to attach an appropriate photometric uncertainty to each point on the synthetic curve. This was done by using the V magnitude of the chosen star to randomly select an RMS uncertainty from the actual distributions in RMS observed in the SuperLupus Survey for stars of that V magnitude, as described below. The RMS scatter for each real lightcurve in the SuperLupus Survey, after the application of Sys-Rem, shows a wide range in RMS scatter at a given V magnitude. Therefore we calculated the actual distribution of RMS scatter for each V magnitude in the survey, as displayed in Figure 3. As an example, the distribution for V = 17.72 is shown in Figure 4. It can be thought of as a vertical slice through the density plot in Figure 3. It is evident that the RMS scatter distributions are non-Gaussian, with long tails to- wards high RMS scatter. This further highlights the importance of basing the RMS scatter of synthetic lightcurves on the actual distributions of photometric precision rather than ap- proximating them using a single-valued function of magnitude as has been done in some previous studies (e.g. Weldrake et al. 2005; Gould et al. 2006). A value for the photometric precision is selected randomly, using the appropriate RMS scatter distribution as a probability weighting, for each Beason¸con star. Once the value of -- 10 -- 0.05 0.04 0.03 0.02 0.01 n o i s i c e r P c i r t e m o t o h P S M R 0 13 14 15 16 V Mag 17 18 1.8 1.6 1.4 1.2 1 0.8 0.6 0.4 0.2 0 l o g n u m b e r d e n s i t y Fig. 3. -- The number density of stars with a given RMS photometric precision plotted against V mag. The number density is given on a logarithmic scale. The bin width is 0.04 mag for V magnitude, and 0.3 mmag for RMS photometric precision. -- 11 -- s r a t S f o r e b m u N 30 25 20 15 10 5 0 0.01 0.015 0.02 0.025 0.03 RMS Photometric Precision Fig. 4. -- RMS photometric precision distribution profile for stars in the SuperLupus Survey between V = 17.70 and V = 17.74. -- 12 -- the photometric precision is determined, a real SuperLupus lightcurve is selected with the same photometric precision, and used as the base lightcurve into which we inject a transit. 6.3. Injecting Transits To simulate the lightcurve from a transiting hot Jupiter, we first determined the param- eters for each simulated system. The stellar mass and radius were taken from the Beason¸con model and Eq. 3. The planetary mass was fixed at 1 MJ , but we note that this does not significantly alter the transit parameters when MP << M⋆. Four different planetary radii were simulated: RP = 0.8, 1.0, 1.2, and 1.4 RJ . The inclination of the orbit, i, was determined by choosing a random cos i between zero and cos imin, the limiting inclination that results in a transit (Sackett 1999): cos imin = R⋆ + RP a , (4) where a is the planet's semi-major orbital axis. Although there is some evidence for a "pile-up" of planets at ∼3 day periods (Cumming et al. 2008), the intrinsic distribution of planets with periods between 1-10 days is not well understood. Therefore it is necessary to make an assumption in assigning the period distribution in the Monte Carlo simulation. Hartman et al. (2009) use a uniform logarithmic distribution. We test this distribution, as well as a uniform linear distribution in period. With the parameters M⋆, MP , R⋆, RP , cos i, and P assigned to each star, a transit depth and duration was calculated and a box-shaped transit inserted into each synthetic lightcurve. For each parameter set, we simulated 10 randomly-selected phases. We adopted a simple box-shape model as limb darkening effects have been shown to be negligible in this signal-to-noise regime (Gould et al. 2006). The transits were inserted into lightcurves that had already been detrended using Sys-Rem, whereas a transit signal in our survey would need to pass through the SyS-Rem detrending. We therefore carefully tested and optimised the number of Sys-Rem iterations we used so that we would not degrade a real astrophysical signal. Only those trends that are present in many light curves with the same temporal timestamps are removed. -- 13 -- 6.4. Detection Efficiency and Effective Number of Stars Probed The BLS algorithm (Kov´acs et al. 2002) was applied to each synthetic lightcurve, using the same parameters used for the actual SuperLupus dataset (see Section 4). In total, 800,000 lightcurves were generated for the Monte Carlo and searched for transit signatures. Transits were deemed to be detected if the selection criteria used in the actual SuperLupus search (see Section 4) were satisfied. The detection efficiency, ε, is defined as the fraction of lightcurves in which the inserted transit signal is detected in the Monte Carlo simulation: ε = 1 Nsim Nsim X i=1 δi, , (5) where δi is a delta function representing whether the transit of star i is detected (δi = 1) or not detected (δi = 0), and Nsim is the number of simulated stars in the Monte Carlo simulation. As discussed in Section 6.1, 24% of our survey stars are giants (log(g) < 4.0), and were automatically assigned a δi = 0 regardless of the other system parameters. We also note that only transiting systems were simulated, so ε does not include the geometric probability of transit. To determine the frequency of hot Jupiters in the field of the SuperLupus Survey we need to determine how many dwarf stars we have effectively probed for transiting planets, Npr. This number depends on the geometric probability of transit, the number of stars monitored to a given precision (for SuperLupus this is 20,465), as well as the detection efficiency. We calculated the geometric probability of transit, f geo , for each system generated in our Monte Carlo simulation. Npr is therefore given as: i Npr = Nsur Nsim Nsim X i=1 δif geo i . (6) Both the detection efficiency and the effective number of dwarf stars probed are set out in Table 4 for a variety of period ranges and assumed period distributions of hot Jupiters. Also tabulated are the mean transit depth, transit duration, and mean transit probability for each set of Monte Carlo assumptions about the size and period distribution of hot Jupiters. The results are presented for all periods simulated (1-10 days), as well as for period ranges of 1-3 days, 3-5 days, and 5-10 days in order to allow for direct comparison with the results of Gould et al. (2006) and Hartman et al. (2009). In Figure 5, Npr is plotted as a function of planet radii for each of the four planetary radii simulated. As expected, more stars are probed in all period ranges for large radii -- 14 -- planets, and more stars are probed for shorter period planets. The difference between the uniform and uniform logarithmic period distributions is most marked for the 1-3 d period planets. ) r p N ( d e b o r P s r a t S f o r e b m u N 2500 2000 1500 1000 500 0 0.8 1 1.2 1.4 Planet Radius (RJ) Fig. 5. -- The effective number of dwarf stars probed (Npr) for transiting hot Jupiters in the SuperLupus Survey as a function of planet radius. The solid lines represent the assumption of a uniform period distribution, while the dashed lines represent the assumption of a log uniform distribution. The square symbols are for 1-3 d periods, the circles for 3-5 d periods, and the triangles for 5-10 d periods. -- 15 -- 7. Frequency of Planets in the Survey The frequency of dwarf stars that host hot Jupiter planets, f , can be calculated simply using the Monte Carlo results and the SuperLupus Survey results: f = n Npr , (7) where n is the number of planets detected in the survey. With only one (or possibly two) planets detected, Poisson statistics are used to determine confidence intervals, or upper limits in the cases of no planet detections. The probability of detecting n planets when λ planets are expected is: P (n, λ) = eλ λn n! (8) . Since the underlying distribution of planets is unknown, a Bayesian approach is used, with a flat prior for the uniform distribution, and a log prior for the logarithmic distribution. The upper and lower 90% confidence intervals (σhigh and σlow, respectively) are calculated by solving: = 0.9 . (9) σlow R σhigh R ∞ eλλn n! dλ eλλn n! dλ 0 The values of σhigh and σlow were determined numerically from Eq. 9, with a symmetric confidence interval such that 5% of the probability distribution was below σlow and 5% above σhigh. The limits σhigh and σlow were then used to calculate the upper and low limits for 90% confidence intervals. Where no planets were detected in a period range, a one-sided 95% confidence upper limit is used so that it can be directly compared with the two-sided 90% limits. In that case, n = 0, and Eq. 9 simplifies to: This method of calculating confidence limits is similar to that used by Gould et al. (2006) and Hartman et al. (2009). 1 − e−σhigh = 0.95 . (10) The follow-up work presented in Weldrake et al. (2008) reveals that one of the SuperLu- pus candidates, SL-07, is indeed a hot Jupiter planet. Seven other candidates can be ruled out as false positives, while SL-04 and SL-06 remain as candidates, although follow-up would be difficult due to their faintness. We therefore present our frequency results for the case where Lupus-TR-3b is the only transiting planet detected (n = 1) and also for the case where either SL-04 or SL-06 is also a transiting planet (n = 2). The results are set out in Table 5, again for both uniform and uniform logrithmic period distributions, and are tabulated in -- 16 -- the same period ranges as used in Table 4. Frequencies for each of the four simulated planet radii are calculated, along with the mean over these radii, which is RP = 1.1. 7.1. Comparison with Other Surveys There are four other transit studies of non-cluster fields that report statistics for hot Jupiter frequency. The Kepler Mission (Borucki et al. 1997) yields very high precision, near continiuos photometry for around 150,000 target stars in a selected galactic field. Howard et al. (2011) calculate the the frequency of planets within 0.25AU of solar-type hosts stars selected from the Kepler target stars. For the period range of 1.2<P<10 d, and radii 0.7< RJ <1.4, very similiar to those considered in this work, the frequency of hot Jupiters is given as 0.37% (Figure 4 in Howard et al. 2011). This frequency is based only on "solar type" Kepler target stars, defined as those with temperatures of Tef f = 4100 − 6100 K and gravities of 4.0 < log g < 4.9. While this gravity cut is essentially the same as is used in this work, the temperature cut means that a more limited sample of stars are being probed. Coupled with the fact that these stars are already drawn from a set of targets selected using "detectability metrics" that suggested they were most likely to give a detectable transit signal for terrestrial- size planets (Batalha et al. 2010), the frequency of 0.37% is higher than one might expect for ground based survey of random Galactic dwarfs in the field. Indeed the Kepler result is at the high end of the frequency of hot Jupiters determined from our SuperLupus survey, just consistent within our uncertainty. The MMT transit survey of the open cluster M37 (Hartman et al. 2009) included a significant number of non-cluster field stars, and these were analyzed separately to determine the hot Jupiter frequency for Galactic dwarf stars. Assuming a Jupiter-sized planet, and a logarithmic period distribution, that study found that the fraction of field stars hosting planets with periods of 1-3 d is <0.8%, and those with planets having periods of 3-5 d is <2.7%. These statistics are upper limits only, as there were no confirmed transiting planets. Analysis of the OGLE-III survey for transiting planets puts the frequency of hot Jupiters at 0.14% for 1-3 d period planets, and 0.31% for periods of 3-5 d (Gould et al. 2006). These statistics draw on survey fields in both the Galactic plane and the Galactic bulge, and were derived based on five transiting planets that had been confirmed from the OGLE-III survey. They rely on an estimation of the efficiency of the "by-eye" selection that was used by the OGLE team to select transiting planet candidates. The Hubble Space Telescope (HST) was used to undertake a survey for transiting planets -- 17 -- in the crowded bulge region of the Galaxy (Sahu et al. 2006). Sixteen candidates were dis- covered from a sample of 180,000 stars monitored, however only one of these (SWEEPS-11) was bright enough to be confirmed in the usual manner using radial velocity measurements. Assuming that all the candidates are indeed planets, it was reported that 0.42% of bulge stars more massive than 0.44 M⊙ have giant planets with periods up to 4.2 days (Sahu et al. 2006). Our result for 1-3 d period planets, f < 0.13%, is consistent with the limits from the deep MMT survey of Hartman et al. (2009). It is lower than the OGLE results, but consistent within the lower half of their 90% confidence interval. For 3-5 d period planets, our result of f = 0.15+0.41 −0.11 is consistent with the upper limit from the deep MMT survey. It is lower than the OGLE results, but still within the 90% confidence intervals. For 5-10 d periods we are able to place limits on the frequency of hot Jupiters of f < 0.93%, assuming a logarithmic distribution in period and also that SL-04 and SL-06 do not host genuine hot Jupiters. If one of these candidates is a transiting hot Jupiter this frequency in this period range would be 0.44+1.19 −0.34%, closer to the Kepler figure of 0.37% but still lower than the RV survey results. The comparisons are summarised in Table 6. The frequency of hot Jupiters can also be determined from radial velocity surveys. Cumming et al. (2008) report a frequency of 0.65± 0.4% from the Keck Planet Search, while Naef et al. (2005) report a frequency of 0.7 ± 0.5% from the ELODIE planet search. These comparisons are also set out in Table 6. Both of these radial velocity results are above our upper 90% confidence interval (0.56%), although the confidence intervals of the two survey results overlap. There are, however, major differences between transit surveys and these radial velocity surveys that could lead to discrepancies in the frequency derived for hot Jupiters. Radial ve- locity surveys are sensitive to planetary mass, while transit surveys are sensitive to planetary radius. Also transit and radial velocity surveys usually monitor different stellar populations. Radial velocity surveys typically monitor carefully selected bright, sun-like stars. More re- cently, bright M dwarfs have also been monitored (Forveille et al. 2009). Transit searches monitor every star in the survey field that has the requisite signal-to-noise to yield high precision photometry. In a deep survey such as the SuperLupus Survey, this will translate to a stellar population with a sub-solar mean mass (0.9M⊙ in the case of the SuperLupus Survey). If hot Jupiters of a given period are less frequent around lower mass stars, then the frequency of hot Jupiters from deep transit surveys will be lower than radial velocity surveys. -- 18 -- 8. Discussion The frequency of hot Jupiters in the Galaxy is an important quantity that will ultimately provide a constraint on models of planet formation and migration. It has been suggested that the frequency of hot Jupiters in globular cluster environments is lower than that of field stars, and that this may be due to crowding or low metallicity affecting planet formation, migration, or survival (Gilliland et al. 2000; Weldrake et al. 2005). Our results indicate, however, that there is little statistical disagreement between hot Jupiter frequencies in cluster and non-cluster environments, even if one of the remaining candidates turns out to be a genuine planet. This is consistent with the work presented in (Saders & Gaudi 2011) which concludes that there is no evidence to support that open clusters have a lower frequency of hot Jupiters. Initial estimates for planet yields from transit surveys turned out to be far in excess of the actual discovery rate (Horne 2003). One of the many factors that led to this over- estimation was the adoption of the hot Jupiter frequency derived from early radial velocity surveys, which as we have shown is higher than is found from transit surveys. The Kepler result of 0.37% (Howard et al. 2011) obviously provides a robust statistic for hot Jupiter frequencies due to the high detection efficiency of that survey. However the result should be adopted cautiously when calculating expected yields from typical transit surveys, as it is based on a sample of "solar type" stars drawn from Kepler target stars, rather than the ensemble field stars monitored by most blind transit surveys. The authors thank David Weldrake and Brandon Tingley, who initiated the original Lupus Survey with PDS. We also thank Grant Kennedy and Tom Evans for assisting in gathering data for the SuperLupus project. Facilities: SSO:1m (WFI) -- 19 -- REFERENCES Alard, C., & Lupton, R. H. 1998, ApJ, 503, 325 Batalha, N. M., et al. 2010, ApJ, 713, L109 Bayliss, D. D. R., & Sackett, P. D. 2011, in EPJ 2011 Vol. 11, Detection and Dynamics of Transiting Exoplanets, ed. F. Bouchy, R. Diaz, & C. Moutou, 01008 Bayliss, D. D. R., Weldrake, D. T. F., Sackett, P. D., Tingley, B. W., & Lewis, K. M. 2009, AJ, 137, 4368 Beatty , T. G. & Gaudi, B. S. 2008,ApJ686, 1302 Bertin, E., & Arnouts, S. 1996, A&AS, 117, 393 Borucki, W. J., Koch, D. G., Dunham, E. W., & Jenkins, J. M. 1997, in Astronomical Society of the Pacific Conference Series, Vol. 119, Planets Beyond the Solar System and the Next Generation of Space Missions, ed. D. Soderblom, 153 Borucki, W. J., et al. 2003, in SPIE Conf. Ser., Vol. 4854, ed. J. C. Blades & O. H. W. Siegmund, 129 -- 140 Burke, C. J., Gaudi, B. S., DePoy, D. L., & Pogge, R. W. 2006, AJ, 132, 210 Cox, A. N., ed. 2000, Allen's Astrophysical Quantities (4th ed.; New York: AIP Press) Cumming, A., Butler, R. P., Marcy, G. W., Vogt, S. S., Wright, J. T., & Fischer, D. A. 2008, PASP, 120, 531 Forveille, T., et al. 2009, A&A, 493, 645 Gilliland, R. L., et al. 2000, ApJ, 545, L47 Gould, A., Dorsher, S., Gaudi, B. S., & Udalski, A. 2006, Acta Astronomica, 56, 1 Hartman, J. D., et al., ApJ, 695, 336 Horne, K. 2003, in ASP Conf. Ser., Vol. 294, Scientific Frontiers in Research on Extrasolar Planets, ed. D. Deming & S. Seager, 361 -- 370 Howard, A. W., et al. 2011, arXiv1103.2541 Kov´acs, G., Zucker, S., & Mazeh, T. 2002, A&A, 391, 369 -- 20 -- Marcy, G. W., & Butler, R. P. 1996, ApJ, 464, L147 Mayor, M., & Queloz, D. 1995, Nature, 378, 355 Naef, D., Mayor, M., Beuzit, J.-L., Perrier, C., Queloz, D., Sivan, J.-P., & Udry, S. 2005, in ESA Special Publication, Vol. 560, 13th Cambridge Workshop on Cool Stars, Stellar Systems and the Sun, ed. F. Favata, G. A. J. Hussain, & B. Battrick, 833 Pont, F., Zucker, S., & Queloz, D. 2006, MNRAS, 373, 231 Robin, A. C., Reyl´e, C., Derri`ere, S., & Picaud, S. 2003, A&A, 409, 523 Sackett, P. D. 1999, in NATO ASIC Proc. 532: Planets Outside the Solar System: Theory and Observations, ed. J.-M. Mariotti & D. Alloin, 189 van Saders, J. L. & Gaudi, B. S. 2011 arXiv:1009.3013v1 Sahu, K. C., et al. 2006, Nature, 443, 534 Schlegel, D. J., Finkbeiner, D. P., & Davis, M. 1998, ApJ, 500, 525 Tamuz, O., Mazeh, T., & Zucker, S. 2005, MNRAS, 356, 1466 Tingley, B., & Sackett, P. D. 2005, ApJ, 627, 1011 Weldrake, D. T. F., & Bayliss, D. D. R. 2008, AJ, 135, 649 Weldrake, D. T. F., Bayliss, D. D. R., Sackett, P. D., Tingley, B. W., Gillon, M., & Setiawan, J. 2008, ApJ, 675, L37 Weldrake, D. T. F., Sackett, P. D., Bridges, T. J., & Freeman, K. C. 2005, ApJ, 620, 1043 Wozniak, P. R. 2000, Acta Astronomica, 50, 421 This preprint was prepared with the AAS LATEX macros v5.2. -- 21 -- Table 1. Candidate selection criteria Criteria BLS Signal Detection Efficiency Effective S/N of transit BLS-Period Magnitude Single-event fraction Not a variable star Threshold SDE > 7.0 α > 10 P 6≤ 1.02 d P 6= 2.00 ± 0.02 d V < 18.8 f rac. < 0.9 Weldrake & Bayliss (2008) catalog Table 2. Properties of Initial SuperLupus Candidates ID R.A. (J2000) Decl. (J2000) Mag Period (V ) (days) Depth Duration (V +R) (hr) TC (HJD) SDE α ηp SL-01 SL-02 SL-03 SL-04 SL-05 SL-06 SL-07 SL-08 SL-09 SL-10 15:30:47.11 15:31:15.28 15:30:43.66 15:31:37.13 15:30:24.24 15:30:08.77 15:30:18.71 15:28:56.97 15:29:37.86 15:29:56.55 -42:44:29.5 -42:47:28.4 -43:01:43.8 -43:10:16.9 -43:15:10.7 -43:16:04.5 -42:58:41.3 -42:55:18.2 -43:07:36.6 -42:33:30.2 16.8 16.7 18.2 17.8 18.4 18.6 17.6 18.2 18.4 17.9 8.59642 7.44930 5.38421 7.72108 1.88490 6.80834 3.91403 7.08871 5.88226 5.12897 0.006 0.005 0.042 0.036 0.026 0.039 0.012 0.031 0.035 0.020 6.03 10.32 3.66 8.08 2.15 1.92 3.24 1.96 1.80 1.97 2454616.0322 2453894.0654 2453880.1516 2453883.2487 2453530.1826 2454615.9351 2453530.9068 2454615.9576 2454615.9640 2454528.1707 7.53 8.00 8.60 8.30 7.70 7.80 9.05 9.13 7.53 7.70 11.4 13.1 21.0 32.4 16.2 10.2 13.4 11.4 12.0 10.8 1.2 2.1 1.4 1.9 1.0 0.7 1.0 0.6 0.6 0.6 -- 2 2 -- -- 23 -- Table 3. Candidate Test Results ID Discriminant Conclusion Saturated close neighbor High ηp Additional Photometric Analysis SL-01 SL-02 SL-03 SL-04 Transit depth (0.042mag) suggests EcB SL-05 SL-06 SL-07 SL-08 SL-09 SL-10 Transit correlated to Sj Transit correlated to Sj Transit correlated to Sj Additional Photometric Analysis Low α (10.2) Lupus-TR-3 Statistical False Positive EcB Blended EcB Candidate EcB Candidate Planet Statistical False Positive Statistical False Positive Statistical False Positive -- 24 -- Table 4. Monte Carlo results Mean Mean Planet Mean Stars Radius Depth Duration probability Efficiency Probed (Npr) (RJ ) Detection of transit (mag) (hr) (ε) 0.8 1.0 1.2 1.4 0.8 1.0 1.2 1.4 0.8 1.0 1.2 1.4 0.8 1.0 1.2 1.4 0.8 1.0 1.2 1.4 0.8 1.0 1.2 1.4 0.8 1.0 1.2 1.4 0.8 0.010 0.016 0.023 0.031 0.010 0.016 0.023 0.031 0.010 0.016 0.022 0.031 0.010 0.016 0.023 0.031 0.010 0.016 0.023 0.031 0.010 0.016 0.023 0.031 0.010 0.015 0.022 0.031 0.010 2.44 2.49 2.53 2.60 1.80 1.81 1.84 1.89 2.26 2.30 2.35 2.40 2.78 2.84 2.89 2.96 2.15 2.20 2.23 2.28 1.72 1.77 1.79 1.83 2.24 2.25 2.34 2.36 2.76 0.091 0.093 0.095 0.097 0.156 0.161 0.164 0.167 0.095 0.097 0.099 0.101 0.063 0.064 0.066 0.067 0.120 0.122 0.125 0.128 0.168 0.171 0.174 0.178 0.096 0.098 0.101 0.102 0.065 0.12 0.27 0.43 0.55 0.28 0.51 0.73 0.85 0.15 0.34 0.54 0.71 0.05 0.13 0.26 0.36 0.18 0.37 0.54 0.68 0.28 0.54 0.73 0.85 0.13 0.32 0.53 0.70 0.05 231 488 762 967 697 1311 1890 2219 228 527 851 1127 46 141 275 392 416 841 1216 1517 759 1466 2007 2391 193 502 841 1124 50 Period Range, Distributiona 1-10 d, Uni 1-3 d, Uni 3-5 d, Uni 5-10 d, Uni 1-10 d, Log 1-3 d, Log 3-5 d, Log 5-10 d, Log -- 25 -- Table 4 -- Continued Period Range, Distributiona Planet Mean Stars Radius Depth Duration probability Efficiency Probed (Npr) (RJ ) Detection of transit (mag) (hr) Mean Mean (ε) 1.0 1.2 1.4 0.016 0.023 0.031 2.79 2.83 2.92 0.066 0.068 0.069 0.14 0.26 0.38 152 281 431 aUni=Uniform distribution, Log=Logarithmic distribution Table 5. Frequency of hot Jupiters Period Range & Planets Detectedb Planet Frequency (f ) Distribution a (n) 0.8 RJ 1.0 RJ 1.2 RJ 1.4 RJ mean RP 1-10 d, Uni 1-3 d, Uni 3-5 d, Uni 5-10 d, Uni 1-10 d, Log 1-3 d, Log 3-5 d, Log 5-10 d, Log 1 [2] 0 1 0 [1] 1 [2] 0 1 0 [1] 0.43+1.62 −0.28 % <0.43% 0.44+1.65 −0.28 % <6.53% 0.24+0.65 −0.18 % <0.28% 0.52+1.40 −0.40 % <4.27% <0.62% [0.20+0.77 −0.13 %] <0.39% <0.31% 0.16+0.62 [0.32+0.70 −0.10 % −0.19 %] <0.23% <0.57% <0.16% <0.14% <0.20% <0.35% <0.27% 0.15+0.54 −0.10 % <2.13% [0.71+2.66 −0.46 %] <1.09% <0.77% <1.41% [0.47+1.75 −0.30 %] <0.25% (0.12+0.32 −0.09 %) <0.18% <0.14% 0.10+0.27 (0.20+0.32 −0.08 % −0.14 %) <0.15% <0.43% <0.11% <0.09% <0.13% <0.25% <0.19% 0.15+0.41 −0.11 % <1.41% [0.66+1.78 −0.50 %] <0.76% <0.50% <0.93% [0.44+1.19 −0.34 %] aUni=Uniform distribution, Log= Logarithmic distribution bStatistics in square brackets are for case where either SL-04 or SL-06 is a planet -- 2 6 -- Table 6. Comparison of Planet Frequencies Survey & Reference Method & Population Assumptions Kepler (Howard et al. 2011) Transit, "solar-type" dwarf stars 1.2 < P < 10 day 0.7 < RP < 1.4RJ n = 22 Keck Planet Search (Cumming et al. 2008) Radial velocity, nearby dwarf stars P < 5 day Mp > 0.5 MJ , ELODIE Planet Search Radial velocity, (Naef et al. 2005) nearby dwarf stars P < 5 day Mp > 0.5 MJ , SWEEPS (Sahu et al. 2006) Transit, bulge stars, M⋆ > 0.5 M⊙ P < 4.2 day n = 16 Planet Frequency 0.37 % 0.65 ± 0.4% 0.7 ± 0.5% 0.4+0.4 −0.2 % -- 2 7 -- OGLE-III (Gould et al. 2006) Transit, bulge and Galactic disk dwarfs stars n=5, 1.0 < RP < 1.25 RJ log P distribution, 1 ≤ P ≤ 3 day 3 ≤ P ≤ 5 day 0.14+0.15 0.31+0.42 −0.08 % −0.18 % Table 6 -- Continued Survey & Reference Method & Population Assumptions Planet Frequency Deep MMT (Hartman et al. 2009) Galactic disk Transit, dwarf stars n=0, R = 1.0 RJ log P distribution 0.4 ≤ P ≤ 1 day 1 ≤ P ≤ 3 day 3 ≤ P ≤ 5 day SuperLupus (this work) n=1, R = 1.1 RJ Transit, Galactic disk log P distribution dwarf stars n=1 (Lupus-TR-3b) 1 ≤ P ≤ 3 day 3 ≤ P ≤ 5 day 5 ≤ P ≤ 10 day 1 ≤ P ≤ 10 day -- 2 8 -- < 0.3% < 0.8% < 2.7% <0.15% 0.15+0.41 −0.11 % <0.93 % 0.10+0.27 −0.08 % -- 29 --
1706.04990
2
1706
2017-06-19T03:45:42
Eccentric Companions to Kepler-448b and Kepler-693b: Clues to the Formation of Warm Jupiters
[ "astro-ph.EP", "astro-ph.SR" ]
I report the discovery of non-transiting close companions to two transiting warm Jupiters (WJs), Kepler-448/KOI-12b (orbital period $P=17.9\,\mathrm{days}$, radius $R_{\rm p}=1.23^{+0.06}_{-0.05}\,R_{\rm Jup}$) and Kepler-693/KOI-824b ($P=15.4\,\mathrm{days}$, $R_{\rm p}=0.91\pm0.05\,R_{\rm Jup}$), via dynamical modeling of their transit timing and duration variations (TTVs and TDVs). The companions have masses of $22^{+7}_{-5}\,M_{\rm Jup}$ (Kepler-448c) and $150^{+60}_{-40}\,M_{\rm Jup}$ (Kepler-693c), and both are on eccentric orbits ($e=0.65^{+0.13}_{-0.09}$ for Kepler-448c and $e=0.47^{+0.11}_{-0.06}$ for Kepler-693c) with periastron distances of $1.5\,\mathrm{au}$. Moderate eccentricities are detected for the inner orbits as well ($e=0.34^{+0.08}_{-0.07}$ for Kepler-448b and $e=0.2^{+0.2}_{-0.1}$ for Kepler-693b). In the Kepler-693 system, a large mutual inclination between the inner and outer orbits ($53^{+7}_{-9}\,\mathrm{deg}$ or $134^{+11}_{-10}\,\mathrm{deg}$) is also revealed by the TDVs. This is likely to induce a secular oscillation of the inner WJ's eccentricity that brings its periastron close enough to the host star for tidal star-planet interactions to be significant. In the Kepler-448 system, the mutual inclination is weakly constrained and such an eccentricity oscillation is possible for a fraction of the solutions. Thus these WJs may be undergoing tidal migration to become hot Jupiters (HJs), although the migration via this process from beyond the snow line is disfavored by the close-in and massive nature of the companions. This may indicate that WJs can be formed in situ and could even evolve into HJs via high-eccentricity migration inside the snow line.
astro-ph.EP
astro-ph
Accepted for publication in AJ on June 18 Preprint typeset using LATEX style AASTeX6 v. 1.0 ECCENTRIC COMPANIONS TO KEPLER-448b AND KEPLER-693b: CLUES TO THE FORMATION OF WARM JUPITERS Kento Masuda1,2,3 1Department of Astrophysical Sciences, Princeton University, Princeton, NJ 08544, USA 2Department of Physics, The University of Tokyo, Tokyo 113-0033, Japan 3NASA Sagan Fellow ABSTRACT I report the discovery of non-transiting close companions to two transiting warm Jupiters (WJs), Kepler-448/KOI-12b (orbital period P = 17.9 days, radius Rp = 1.23+0.06−0.05 RJup) and Kepler-693/KOI- 824b (P = 15.4 days, Rp = 0.91 ± 0.05 RJup), via dynamical modeling of their transit timing and duration variations (TTVs and TDVs). The companions have masses of 22+7−5 MJup (Kepler-448c) and 150+60−40 MJup (Kepler-693c), and both are on eccentric orbits (e = 0.65+0.13−0.09 for Kepler-448c and e = 0.47+0.11−0.06 for Kepler-693c) with periastron distances of 1.5 au. Moderate eccentricities are detected for the inner orbits as well (e = 0.34+0.08−0.07 for Kepler-448b and e = 0.2+0.2−0.1 for Kepler-693b). In the Kepler-693 system, a large mutual inclination between the inner and outer orbits (53+7−9 deg or 134+11−10 deg) is also revealed by the TDVs. This is likely to induce a secular oscillation of the inner WJ's eccentricity that brings its periastron close enough to the host star for tidal star -- planet interactions to be significant. In the Kepler-448 system, the mutual inclination is weakly constrained and such an eccentricity oscillation is possible for a fraction of the solutions. Thus these WJs may be undergoing tidal migration to become hot Jupiters (HJs), although the migration via this process from beyond the snow line is disfavored by the close-in and massive nature of the companions. This may indicate that WJs can be formed in situ and could even evolve into HJs via high-eccentricity migration inside the snow line. Keywords: planets and satellites: formation -- planets and satellites: individual (KOI-12, Kepler-448, individual (KOI-824, Kepler-693, KIC 5164255) KIC 5812701) -- planets and satellites: -- techniques: photometric 1. INTRODUCTION Warm Jupiters (WJs), giant planets in moderately close-in orbits (7 days < P < 100 days), pose a simi- lar conundrum to that of hot Jupiters (HJs). A dozen WJs have been found to reside in circular orbits in multi- transiting systems (Huang et al. 2016), in which the or- bital planes of the planets are likely well aligned. Such an architecture points to the formation via disk migra- tion (Goldreich & Tremaine 1980; Lin et al. 1996) as originally proposed for HJs, or in situ formation inside the snow line (Boley et al. 2016; Batygin et al. 2016). Alignments of the planetary orbits with their host stars' equators, as confirmed for some of them (e.g., Sanchis- Ojeda et al. 2012; Hirano et al. 2012), may also support the absence of past dynamical eccentricity/inclination excitation via planet -- planet scattering (Rasio & Ford 1996) or secular chaos (Wu & Lithwick 2011). [email protected] On the other hand, roughly half of WJs with mea- sured masses from Doppler surveys have significant ec- centricities that seem too large to result from disk mi- gration or subsequent planet -- planet scattering (Petro- vich et al. 2014), but yet too small for their orbits being tidally circularized (Socrates et al. 2012b; Dawson et al. 2015). A possible explanation is that those moderately eccentric WJs are experiencing "slow Kozai -- Lidov mi- gration" (Petrovich 2015): their eccentricities are still undergoing large oscillations driven by the secular per- turbation from a close companion (Dong et al. 2014), without being suppressed by other short-range forces (Wu & Murray 2003), and their orbits shrink only at the high-eccentricity phase. This scenario may indeed reproduce the observed eccentricity distribution of WJs with outer companions (Petrovich & Tremaine 2016). Observationally, long-period, massive companions to WJs are nearly as common as those of HJs (Knutson et al. 2014) and their orbital properties might be con- sistent with what is expected from this scenario (Bryan 2 Masuda et al. 2016). Indeed, the apsidal misalignments of some of those eccentric WJs with outer companions provide statistical evidence for the oscillating eccentricity due to a large mutual orbital inclination (Dawson & Chiang 2014). However, there has been no direct measurement of such a large orbital misalignment between WJs and their outer companions as to induce a large eccentric- ity oscillation, partly because they are mostly the sys- tems detected with the radial velocity (RV) technique that yields no information on the orbital direction. No- table exceptions include the Kepler-419 system (Dawson et al. 2014) and the doubly-transiting giant-planet sys- tem Kepler-108 (Mills & Fabrycky 2017), where the mu- tual inclinations were constrained via dynamical model- ing of transit timing and duration variations (TTVs and TDVs), though their mutual inclinations are likely too small to drive secular eccentricity oscillations. Transiting WJs without transiting companions pro- vide a unique opportunity to search for close and mutually-inclined companions as direct evidence for the slow Kozai migration, because the full 3D architecture of the system can be dynamically constrained with a similar technique as used in the above systems. In ad- dition, WJs on eccentric and intermediate orbits, un- like HJs, may still be interacting with the companion strongly, and their eccentricity can also help the TTV inversion for non-transiting objects by producing spe- cific non-sinusoidal features. The TTV search for the outer companions on such "hierarchical" orbits is also complementary to the TTV studies so far, which have mainly focused on nearly sinusoidal signals typical for planets near mean-motion resonances (Hadden & Lith- wick 2016; Jontof-Hutter et al. 2016). In this paper, I perform a search for non-transiting companions around transiting WJs using transit varia- tions (Section 2) and report the discovery of outer com- panions to two transiting WJs, Kepler-448b (Bourrier et al. 2015) and Kepler-693b (Morton et al. 2016). Based on TTVs and TDVs of the WJs, I find that the compan- ions are (sub-)stellar mass objects on highly eccentric, au-scale orbits (Sections 3 -- 5). In particular, I confirm a large mutual orbital inclination between the inner WJ and the companion in the latter system, which can in- duce a large amplitude of eccentricity oscillation and the tidal shrinkage of the inner orbit (Section 6) -- exactly as predicted in the "slow Kozai" scenario by a close com- panion. The companions' properties, however, are not fully compatible with such migration starting from be- yond the snow line. Thus I also assess the possibility of in situ formation (Section 6.3). I discuss possible follow- up observations in Section 7 and Section 8 concludes the paper. 2. SYSTEMATIC TTV SEARCH FOR SINGLY-TRANSITING WARM JUPITERS To identify the signature of outer companions, I ana- lyzed TTVs of 23 confirmed, singly-transiting WJs (Sec- tion 2.1) with the orbital period 7 days < P < 100 days and radius Rp > 8R⊕ in the DR24 of the KOI cata- log (Coughlin et al. 2016). Systems with multiple KOIs are all excluded, even though they consist of only one confirmed planet and false positives. I found clearly non-sinusoidal TTVs for Kepler-448/KOI-12b, Kepler- 693/KOI-824b, and Kepler-419/KOI-1474b. The result is consistent with the TTV search by Holczer et al. (2016), who reported significant long-term TTVs for the same three KOIs in our sample.1 Among those planets, Kepler-419b's TTVs have al- ready been analyzed by Dawson et al. (2014) with RV data, and the companion planet Kepler-419c was found to be a super Jupiter on an eccentric and mutually- aligned orbit with the inner one. Therefore in this pa- per, I focus on Kepler-448b and Kepler-693b, which both show clear non-sinusoidal TTVs, and masses and orbits of the perturbers can be well constrained. For the dynamical modeling taking into account the possible orbital misalignment, I reanalyze the transit light curves of Kepler-448b and Kepler-693b to derive both TTVs and TDVs (Section 2.2), consistently with the other transit parameters (Tables 1 -- 4). Here I also fit the planet-to-star radius ratio, Rp/R(cid:63), so that the possi- ble transit depth modulation due to the different crowd- ing depending on observing seasons does not mimic du- ration variations (cf. Mills & Fabrycky 2017). As shown in Figure 1, I find significant TTVs consisting of a long- term modulation and a short-term, sharp feature (top panels). For Kepler-448b, the spike-like feature is more clear than that reported in Holczer et al. (2016), pre- sumably owing to a more careful treatment of the local baseline modulation. No significant correlation is found between TTVs and the local light-curve slope or the fit- ted Rp/R(cid:63), which supports the physical origin rather than due to star spots (Mazeh et al. 2015). I show in Sections 4 and 5 that this unusual feature is reproduced by a periastron passage of an outer non-transiting com- panion in an eccentric orbit. I also identify a significant (∼ 5σ) linear trend in the transit duration of Kepler- 693b, which points to a mutual orbit misalignment. The duration change is also clear in Figure 2, where each de- 1 Due to the update in the stellar radius in the DR25 catalog, the WJ sample defined above now includes Kepler-522/KOI-318b and Kepler-827/KOI-1355b, for which Holczer et al. (2016) re- ported long-term TTVs. Interpretation of their TTVs is less clear than for the two systems discussed in the present paper, because of the lack of clear non-sinusoidal features (Kepler-522b) and a low signal-to-noise ratio (Kepler-827b). trended and normalized transit is overplotted around its center, along with the best-fit model. 2.2. Transit Times and Durations of Kepler-448b and Kepler-693b 3 2.1. Iterative Determination of Transit Times In the systematic TTV search for WJs, I analyze Pre- search Data Conditioning Simple Aperture Photometry (PDCSAP) fluxes obtained with the long-cadence (LC) mode, retrieved from the NASA exoplanet archive. I fol- low the iterative method as described by Masuda (2015) to derive the consistent transit times and light-curve shapes. The method consists of (1) the determination of the central time for each transit and (2) the refinement of the transit parameters by fitting the mean transit light curve. The fitting in this subsection is performed by minimizing the standard χ2 (Markwardt 2009) de- fined as the squared sum of the difference between the model and data divided by the PDCSAP-flux error. In the first step, I fit the data segments of three times the total duration centered at each transit. I adopt the Mandel & Agol (2002) model for the quadratic limb-darkening law generated with pyTransit package (Parviainen 2015), multiplied by a second-order poly- nomial function of time to take into account the local baseline modulation. The fitting is repeated iteratively removing 3σ outliers. Here a circular orbit is assumed and the values of central transit time tc and three coef- ficients of the polynomial are fitted, while the other pa- rameters (two limb-darkening coefficients q1 and q2 de- fined in Kipping (2013), mean stellar density ρ(cid:63), transit impact parameter b, planet-to-star radius ratio Rp/R(cid:63), and orbital period P ) are fixed. In the second step, I shift each transit by the value of tc to align all the transits around time zero. Then the data are averaged into bins of 0.05 times the transit duration (to reduce the computational time), where the value and error of each bin is set to be the median and 1.4826 times median absolute deviation of the flux values in the bin. The resulting "mean" light curve is again fitted with the same Mandel & Agol (2002) model for its central time t0, normalization constant, q1, q2, Rp/R(cid:63), ρ(cid:63), and b assuming a circular orbit and fixing the period at the value refined in step (1); only in the first iteration, I adopt a quadratic function of time as a baseline and fit three coefficients rather than one normalization. The values of q1, q2, Rp/R(cid:63), ρ(cid:63), and b obtained from this fitting is used for the first step of the next iteration. I perform five iterations for each KOI, starting with the second step based on the transit light curve phase- folded at the period given in the KOI catalog (Coughlin et al. 2016). I fit the resulting transit times with straight lines to search for any TTVs and identify the candidates mentioned above. For Kepler-448b and Kepler-693b, I perform more in- tensive analyses taking into account possible variations of transit durations using the short-cadence (SC) data if available. For Kepler-448, I use the SC data for all the quarters, while for Kepler-693 I combine both the LC (Q2, Q7, Q8, Q10 -- Q12, Q14 -- 16) and SC (Q14 -- Q16) data. The method is the same as in Section 2.1 except for the following differences. I additionally fit Rp/R(cid:63) and b in the first step and repeated 10 iterations. Here I fit Rp/R(cid:63) so that the seasonal variation in the transit depths not be misinterpreted as the duration variation (e.g., Van Eylen et al. 2013; Masuda 2015). The result- ing b and Rp/R(cid:63) are combined with ρ(cid:63) and P to yield the transit duration T as the average of the total and full durations (see equations 14 and 15 of Winn 2011). The errors in tc and T are scaled by the square root of the reduced chi-squared χ2 ν of the best-fit model to en- force χ2 ν = 1. When the mean light curve is fitted, the χ2 minimization is complemented by a short Markov Chain Monte Carlo (MCMC) chain (Foreman-Mackey et al. 2013), and the data are binned into one-minute bins for the SC data. I also tuned the parameters specifically to each system as follows: Kepler-448b -- I fit the data of 1.6 times duration for each transit. If more than 10% of data is missing in the segment, the transit is excluded from the analysis. I use fourth-order polynomial because the light curve shows short-term wiggles likely due to the stellar rota- tion (1.245 ± 0.124 days; McQuillan et al. 2013). Kepler-693b -- I fit the data of twice the duration for each transit and use second-order polynomial because the light curve shows smaller variability than Kepler- 448. I omit the transits with more than 10% gaps for the SC data, while for the LC data I omitted those with more than 30%. The SC and LC data are analyzed inde- pendently and the resulting transit times and durations are averaged to give one measurement if both are avail- SC/2 and half of the difference between the LC and SC values, and assign the larger of the two as its error. The former was the case for most of the points, while the latter pro- cess helped mitigate the effect of a few outliers caused presumably by local features in the light curve. able. For each data, I compute(cid:112)σ2 LC + σ2 The resulting transit times, durations, and radius ra- tios are summarized in Figure 1 and Tables 1 and 2. I also fit the mean transit light curve from the final iteration (binned version of those in Figure 2) with an MCMC including an additional Gaussian error in quadrature (denoted as "photometric jitter") and derive 4 Masuda (a) Kepler-448b (b) Kepler-693b Figure 1. Central times, durations, and planet-to-star radius ratios obtained by fitting each transit of (a) Kepler-448b (Table 1) and (b) Kepler-693b (Table 2). In the top panel, residuals from the linear fit to the observed transit times (i.e., TTVs) are shown for clarity. (a) Kepler-448b (b) Kepler-693b Figure 2. Detrended and normalized transits stacked around each center for (a) Kepler-448b (short-cadence data) and (b) Kepler-693b (long-cadence data). Filled circles show the data points, and the solid lines (only in panel b) show the best-fit transit models. The colors correspond to the mid-transit times, as shown in the right bar. Color gradations around the ingress and egress in panel (b) illustrate the systematic decrease in the transit duration (middle panel in the right column of Figure 1). 21012Transit timing variation (min)6.696.706.716.726.736.746.756.766.776.78Transit duration (hour)2004006008001000120014001600Transit time (BJD-2454833)0.0870.0880.0890.0900.0910.0920.093Planet-to-star radius ratio15105051015Transit timing variation (min)2.02.22.42.62.83.0Transit duration (hour)2004006008001000120014001600Transit time (BJD-2454833)0.100.110.120.130.14Planet-to-star radius ratio the posteriors for the transit parameters summarized in Tables 3 and 4. For Kepler-448b, the result agreed well with those from the least-square fit with its standard errors scaled by the square root of χ2 ν. For Kepler-693b, I obtain skewed posteriors and larger errors; the param- eters are better determined by the LC data that include more transits. Table 1. Transit Times, Durations, and Radius Ratios of Kepler-448b 5 Transit number Transit time (BJDTDB − 2454833) Transit duration (day) Planet-to-star radius ratio −1 ··· 128.7420 ± 0.0001 Note -- Quoted uncertainties are the standard errors derived from the covariance matrix scaled by(cid:112)χ2/d.o.f. of the 0.0920 ± 0.0004 0.2808 ± 0.0003 fit. This table is published in its entirety in the machine-readable format. A portion is shown here for guidance regarding its form and content. Table 2. Transit Times, Durations, and Radius Ratios of Kepler-693b Transit number Transit time (BJDTDB − 2454833) Transit duration (day) Planet-to-star radius ratio 0 ··· Note -- Quoted uncertainties are the standard errors derived from the covariance matrix scaled by (cid:112)χ2/d.o.f. of 173.609 ± 0.002 0.115 ± 0.007 0.115 ± 0.008 the fit, or are based on the combination of the LC and SC data (see Section 2.2). This table is published in its entirety in the machine-readable format. A portion is shown here for guidance regarding its form and content. Table 3. Transit Parameters of Kepler-448b Based on Short-cadence Data Parameter Value (Parameters from the Mean Transit Light Curve)a q1 = (u1 + u2)2 q2 = 0.199(7) u1 2(u1+u2) Center of the mean transit (day) Rp/R(cid:63) Normalization Transit impact parameter Mean stellar density (g cm−3) Photometric jitter (a/R(cid:63))e=0 u1 u2 0.39(2) −0.00001(2) 0.09044(7) 0.999999(3) 0.373(6) 0.393(3) 0.000036(2) 18.77(4) 0.348(9) 0.10(2) Table 3 continued Table 3 (continued) Parameter Value (Mean Orbital Period and Transit Epoch)b t0 (BJDTDB) P (day) 2454979.5961(2) 17.855234(4) aMedian and 68% credible interval of the MCMC posteri- ors from the mean transit light curve. Here the circular orbit is assumed to relate mean stellar density to the semi- major axis over stellar radius, a/R(cid:63). The limb-darkening coefficients u1 and u2 are converted from q1 and q2. b Obtained by linearly fitting the series of transit times in Table 1. Errors are scaled by(cid:112)χ2/d.o.f.. Note -- Parentheses after values denote uncertainties in the last digit. 6 Masuda Table 4. Transit Parameters of Kepler-693b Based on Long-cadence Data Parameter Value (Parameters from the Mean Transit Light Curve)a q1 = (u1 + u2)2 q2 = u1 2(u1+u2) Center of the mean transit (day) Rp/R(cid:63) Normalization Transit impact parameter Mean stellar density (g cm−3) Photometric jitter 0.6+0.3−0.2 0.4+0.3−0.2 0.0000(3) 0.117(+4) (−6) 1.00000(8) 0.63+0.06−0.15 3.0+1.0−0.5 0.00074(7) (a/R(cid:63))e=0 u1 u2 (Mean Orbital Period and Transit Epoch)b t0 (BJDTDB) P (day) 34+4−2 0.7+0.2−0.3 0.1+0.3−0.4 15.37566(3) 2455006.613(1) hierarchy of the system, I adopt the Jacobi coordinates to describe their orbits: the inner orbit (denoted by the subscript 1) is defined by the relative motion of the in- ner planet around the host star, while the outer one (denoted by the subscript 2) is the motion of the outer companion relative to the center of mass of the inner two bodies. The sky plane is chosen to be the reference plane, to which arguments of periastron and the line of nodes are referred. The direction of +Z-axis (which matters the definition of "ascending" node) is taken to- ward the observer. 3.2. Bayesian Framework I derive the posterior probability density function (PDF) for the set of system parameters θ conditioned on the observed data d: p(θd, I) = Z p(dθ, I) p(θI), (1) 1 where I denotes the prior information. The normaliza- tion factor Z ≡ p(dθ, I) p(θI)dθ (2) (cid:90) aMedian and 68% credible interval of the MCMC poste- riors from the mean transit light curve. Here the cir- cular orbit is assumed to relate mean stellar density to the semi-major axis over stellar radius, a/R(cid:63). The limb- darkening coefficients u1 and u2 are converted from q1 and q2. b Obtained by linearly fitting the series of transit times in Table 2. Errors are scaled by(cid:112)χ2/d.o.f.. Note -- Parentheses after values denote uncertainties in the last digit. 3. DYNAMICAL MODELING OF TTVS AND TDVS: METHOD 3.1. Model Assumptions The TTVs observed in the two systems (Figure 1) are clearly different from the sinusoidal signal due to a companion near a mean-motion resonance (Lithwick et al. 2012). In particular, rapid timing variations on a short timescale suggest that the perturbing compan- ions' orbits are eccentric. In addition, such a feature is observed only once for each system, and so the compan- ions must be far outside the WJs' orbits. Thus I assume a hierarchical three-body system as the only viable con- figuration and model the observed TTVs and TDVs to derive the outer companions' masses and orbits. I only consider the Newtonian gravity between the three bodies regarded as point masses (see Section 3.5.1 for justification), as well as the finite light-travel time in computing timings. To better take into account the is called the global likelihood or evidence, which rep- resents the plausibility of the model. The prior PDF p(θI) is given as a product of the prior PDFs for each parameter, which are assumed to be independent; they are presented in Tables 5 and 6. The likelihood L ≡ p(dθ, I) is defined and computed as described in Sections 3.4 and 3.5. To invert the observed signals, I utilize the nested- sampling algorithm MultiNest (Feroz et al. 2009; Feroz & Hobson 2008; Feroz et al. 2013) and its python inter- face PyMultiNest (Buchner et al. 2014), which allows us to sample the whole prior volume to identify multi- ple modes if any. I typically utilize 4800 live points and default sampling efficiency of 0.8, keep updating the live points until the evidence tolerance of 0.5 is achieved, and allow for the detection of multiple posterior modes. When multiple modes are detected, MultiNest also computes evidence corresponding to each mode, which is referred to as "local" evidence. 3.3. Procedure for Finding Solutions To check on the reliability of our numerical scheme and to find all the possible posterior modes, I adopt the following procedure. First, I use the analytic TTV for- mula for a hierarchical triple system (Borkovits et al. 2015, Appendix A) to fit only the TTV signal. This al- lows us to search a wide region of the parameter space and resulted in one global mode containing two peaks. The resulting solution was also found to be consistent with a rough analytical estimate based on the observed TTV features (see, e.g., Section 4.2). Then, I fit the same TTVs numerically adopting a slightly narrower prior range that well incorporates the global mode found from the analytic fit. The resulting posterior was found to be consistent with the one derived from the analytic fit. This agreement validates the numerical scheme I rely on. Finally, the same numerical method is used to model both TTVs and TDVs simultaneously to deter- mine the physical and geometric properties of the outer companions. In Sections 4 and 5, I mainly report and discuss the results from the final fit, while the analytic and numerical posteriors from TTVs alone are found in Appendix B for comparison. 3.4. TTV Modeling The likelihood for the TTV modeling is defined as follows: LTTV ≡(cid:89) (cid:113) 1 t,i + σ2 TTV) exp i 2π(σ2 (cid:34) − (ti − tmodel t,i + σ2 2(σ2 i )2 TTV) (cid:35) , where ti and σt,i are the transit times and their errors obtained in Section 2.2 (Tables 1 and 2). I also include σTTV as a model parameter that takes into account the additional scatter and marginalize over it to obtain more conservative constraint; it turns out that this parame- ter is not correlated with any other physical parameters (Figures C2 and C3) and hence does not affect the global shape of the joint PDF. i The model transit times tmodel are computed both an- alytically and numerically as mentioned in Section 3.3; this is for cross-validation, as well as to obtain insights into how the parameters are determined. In both cases, the model includes 14 parameters in addition to σTTV (see Tables B2 and B3): orbital period, orbital phase, ec- centricity, and argument of periastron for both inner and outer orbits; cosine of the outer orbit's inclination (inner one is fixed to be 0 for simplicity; this does not affect the result); difference in the longitudes of the ascending node; and masses of the host star and outer compan- ion (here I fix the inner planet's mass to be 1MJup as it is not determined from TTVs). While I use the time of inferior conjunction tic and orbital period P for the inner orbit, I choose the time of periastron passage τ2 and periastron distance relative to the inner semi-major axis q2/a1 to specify the outer orbital phase and period. This is because the latter two parameters are more di- rectly related to the position and duration of the "spike" in the observed TTVs than tic and P , and thus better determined. I fit the mass scale of the whole system as well because I include the light-travel time effect; in practice, however, the observed TTVs are insensitive to this parameter and its value is solely constrained by the prior knowledge. I first compute analytic transit times using the formula 7 by Borkovits et al. (2011, 2015) developed for hierarchi- cal planetary systems and triple-star systems (Appendix A). I include the light-travel time effect (LTTE) and P2- timescale dynamical effect due to the quadrupole com- ponent of the perturbing potential; I did not find no- table changes even with the octupole terms. The former is due to the motion of the inner binary (here the central star orbited by the inner planet), which changes finite time for the light emitted from the star and blocked by the planet to travel to us. The latter process involves the actual variation in the orbital period of the inner binary due to the tidal gravitational field exerted by the companion. Both of these effects are routinely observed in triple-star systems containing eclipsing binaries (e.g., Rappaport et al. 2013; Masuda et al. 2015). Numerical TTVs are computed using TTVFast code (Deck et al. 2014) by integrating the orbits of the three bodies and finding the times at which sky-projected dis- tance between the star and the inner planet, dsky, be- comes minimum. I choose the time step to be 0.3 days for Kepler-448 and 0.1 days for Kepler-693, which are roughly 1/60 and 1/150 of the inner orbital periods, re- spectively. To compute transit times, I modify the de- fault output of the TTVFast code to take into account the effect of light-travel time using the line-of-sight co- ordinate of the center of mass of the inner binary (i.e., central star and inner planet). 3.5. Joint TTV and TDV Modeling The joint TTV and TDV modeling was performed us- ing the following likelihood: L =LTTV (cid:34) − (Tj − T model T,j + σ2 2(σ2 j )2 dur) (cid:35) ×(cid:89) (cid:113) 1(cid:112)2πσ2 × j b 2π(σ2 1 T,j + σ2 exp dur) (cid:20) − (bmean − bmodel mean )2 (cid:21) , exp 2σ2 b where the second row is defined analogously to LTTV with the transit time t replaced by the transit duration T . I also include the constraint on the transit impact pa- rameter bmean from the mean transit light curve.2 The TDV modeling additionally requires the mean density of the host star ρ(cid:63) (equivalently stellar radius) and the "jitter" for durations σdur, and θ consists of 17 param- eters listed in Tables 5 and 6 as fitted parameters. Here I fit inner orbit's inclination and inner planet's mass, although the latter is not constrained at all by the data. 2 Here we can use bmean derived assuming a circular orbit be- cause the parameter is based on the shape of the light curve alone. A possible bias due to the non-zero eccentricity is absorbed in the fitted stellar density, which is not used in any of the following analyses. See, e.g., Eqn. (28) and (29) of Winn (2011). 8 Masuda This joint modeling is performed only numerically us- ing TTVFast. In addition to the transit times above, I use another default output, dsky at each transit cen- ter, to obtain the model transit impact parameter b = dsky/R(cid:63), where the stellar radius R(cid:63) is given by (3m(cid:63)/4πρ(cid:63))1/3; here I adopt b at the transit around the center of the observing duration as bmodel mean . The transit durations are obtained from yet another default output √ of the code vsky, the sky-projected relative star -- planet 1 − b2R(cid:63)/vsky. velocity at the transit center, via T = 2 3.5.1. Additional TDV Sources While I only consider the Newtonian gravitational in- teraction between point masses, transit durations may also be affected by secular orbit variation due to stel- lar quadrupole moment and/or general relativistic ef- fect. Indeed, we expect that Kepler-448 has a relatively large quadrupole moment J2 due to its rapid rotation (McQuillan et al. 2013; Bourrier et al. 2015), and the inner orbits are moderately eccentric in both systems as will be shown later. Here we show that those effects on transit durations are negligible (the effects on transit times are even smaller; see Miralda-Escud´e 2002). The duration drift due to the nodal precession is given by (Miralda-Escud´e 2002) T ≤ 3J2 b√ 1 − b2 sin λ. (3) where λ is the sky-projected stellar obliquity. For Kepler-448b with b (cid:39) 0.4, a/R(cid:63) (cid:39) 20, and λ (cid:39) 10◦, this results in the duration change over four years (cid:18) a (cid:19)−2 R(cid:63) ∆T (cid:46) 2 × 10−3 hr , (4) (cid:19) (cid:18) J2 10−4 where the quoted value of J2 is motivated by theoretical modeling and observational inference for a star with sim- ilar parameters (Szab´o et al. 2012; Masuda 2015). This ∆T is smaller than the measurement precision. The same is also true for Kepler-693b, for which J2 is likely smaller and measurements of T are less precise. The drift caused by general relativistic apsidal preces- sion is (Miralda-Escud´e 2002) (cid:18) a (cid:19)−1(cid:112) R(cid:63) T ≤ 4e or ∆T (cid:46) 2 × 10−3 hr (cid:17)2 (cid:16) na (cid:19) c 1 − b2 × 3 1 − e2 (cid:18) e 1 − e2 (5) (6) over four years; this is also negligible. Note that possible apsidal precession induced by the gravitational pertur- bation from the outer companion is already taken into account in our Newtonian model. 4. RESULTS: KEPLER-448/KOI-12 The resulting posterior PDF from the MultiNest analysis and the corresponding models are shown in Figures 3 (red solid lines) and Figure C2. The sum- mary statistics (median and 68% credible interval) of the marginal posteriors as well as the priors adopted for those parameters are given in Table 5. The solution consists of two separated posterior modes (denoted as Solution 1 and Solution 2) that reproduce the observed TTVs and TDVs equally well, without any significant difference in the local evidence computed for each mode. In both solutions, the outer companion (we tentatively call it Kepler-448c) is likely to have a mass in the brown- dwarf regime (22+7−5 MJup) and reside in a highly eccen- tric orbit close to the inner WJ. In spite of the small outer orbit, its pericenter distance 1.46+0.07−0.06 au is more than nine times larger than the inner semi-major axis and well within the stable regime (Mardling & Aarseth 2001). The system has one of the smallest binary sep- arations compared to the planet apastron 0.21+0.05−0.04 au among the known planetary systems with (sub-)stellar companions; see figure 8 of Triaud et al. (2017). I also find a modest eccentricity 0.34+0.08−0.07 for the in- ner WJ from the combination of timing and duration information. While the TTVs alone favor an even larger value (Appendix B), transit duration combined with the spectroscopic prior on ρ(cid:63) lowers it. The inner WJ mass is also floated between 0.1-10 MJup but no constraint better than that from the RV data (< 10 MJup as the 3σ upper bound; Bourrier et al. 2015) is found. This is because the TTVs of the inner WJ does not depend on its own mass to the first order. While the above constraints essentially come from TTVs, TDVs possibly allow us to constrain the mutual inclination i21 between the inner and outer orbits. In fact, the two solutions are different in this respect: So- lution 1 is "prograde", i.e., the two orbits are in the same directions, while Solution 2 corresponds to a "ret- rograde" configuration. Non-zero mutual inclinations are slightly favored in both solutions with the posterior probability that 39.◦2 < i21 < 140.◦8 being 27%. How- ever, it is more or less due to the large prior volume cor- responding to the misaligned solution and a wide range of the mutual inclination is allowed by the data (Fig- ure 4). In particular, the data are totally consistent with the coplanar configuration, as visually illustrated in the right panels of Figure 3. I found the evidence for the coplanar model is not significantly different from the misaligned case. Only the configuration with i21 (cid:39) 90◦ is slightly disfavored due to the lack of a large TDV expected from the strong perturbation in the direction perpendicular to the orbital plane. The joint TTV/TDV analysis also allows for con- straining the impact parameter during a possible occul- tation, bocc, based on that during the transit as well as the eccentricity and argument of periastron of the inner orbit. I find bocc = 0.22 ± 0.02, which means that the secondary eclipse should have been detected if Kepler- 448b was a star. The fact strengthens the planetary interpretation by Bourrier et al. (2015) who confirmed mb < 10 MJup at the 3σ level. The following subsections detail the prior information and interpretation of TTVs and TDVs. Table 5. Parameters of the Kepler-448 System from the Dynamical TTV and TDV Analysis Parameter Solution 1 Solution 2 Combined Prior 9 Fitted Parameters (Inner Orbit) 1. Time of inferior conjunction tic,1 (BJDTDB − 2454833) 2. Orbital period P1 (day) 3. Orbital eccentricity e1 4. Argument of periastron ω1 (deg) 5. Cosine of orbital inclination cos i1 (Outer Orbit) a 6. Time of the periastron passage τ2 (BJDTDB − 2454833) 7. Periastron distance over inner semi-major axis q2/a1 8. Orbital eccentricity e2 9. Argument of periastron ω2 (deg) 10. Cosine of orbital inclination cos i2 11. Relative longitude of a ascending node Ω21 (deg)a,b (Physical Properties) 12. Mass of Kepler-448 m(cid:63) (M(cid:12)) 13. Mean density of Kepler-448 ρ(cid:63) (g cm−3) 14. Mass of Kepler-448b mb (MJup)c 15. Mass of Kepler-448c mc (MJup) (Jitters) 16. Transit time jitter σTTV (10−4 day) 17. Transit duration jitter σdur (10−4 day) Derived Parameters Outer orbital period P2 (day) Inner semi-major axis a1 (au) Outer semi-major axis a2 (au) Periastron distance of the outer orbit a2(1 − e2) (au) Mutual orbital inclination i21 (deg) Physical radius of Kepler-448 (R(cid:12)) Physical radius of Kepler-448b (RJup)d 128.7418(+2) (−3) 128.7413(+2) (−2) 128.7415(+4) (−3) U(128.73, 128.75) 17.855183(+9) (−6) 17.855181(+7) (−5) 17.855182(+8) (−6) 0.34+0.07−0.06 −49+9−21 0.013(1) 0.35+0.09−0.08 −131+18−9 0.013(1) 0.34+0.08−0.07 ··· 0.013(1) Ulog(17.8551, 17.8553) U(0, 0.7) U(−180, 180) U(0, 0.02)a 1076+10−11 1080+10−11 1078+10−11 9.5+0.4−0.3 9.4+0.4−0.3 9.4+0.4−0.3 U(900, 1200) Ulog(5, 20) 0.61+0.11−0.07 −89+6−7 −0.3+0.3−0.3 3+4−4 1.5 ± 0.1 0.79+0.09−0.09 1.1+3.5−0.8 21+6−5 2.2 ± 0.3 3.8 ± 0.6 0.69+0.12−0.10 −89+5−6 −0.6+0.3−0.2 −178+5−4 1.5 ± 0.1 0.81+0.10−0.09 1.1+3.5−0.8 24+6−5 2.2 ± 0.3 3.7 ± 0.6 0.65+0.13−0.09 −89+5−6 −0.5+0.4−0.3 ··· 1.5 ± 0.1 0.80+0.10−0.09 1.1+3.5−0.8 22+7−5 2.2 ± 0.3 3.7 ± 0.6 U(0, 0.95) U(−180, 180) U(−1, 1) U(−180, 180) G(1.51, 0.09, 0.14) G(0.76, 0.11, 0.12) Ulog(0.1, 10) Ulog(0.001M(cid:12), 0.1M(cid:12)) Ulog(5 × 10−2, 5) Ulog(0.1, 10) (2.2+1.5−0.5) × 103 (2.9+3.3−1.0) × 103 (2.5+2.4−0.7) × 103 0.154(+4) (−3) 3.8+1.6−0.6 1.47+0.07−0.06 20+17−12 1.40 ± 0.06 1.24+0.06−0.05 0.154(+4) (−3) 5+3−1 1.45+0.06−0.06 146+19−16 1.39+0.07−0.06 1.23+0.06−0.06 0.154(+4) (−3) 4.2+2.4−0.9 1.45+0.07−0.06 ··· 1.40+0.07−0.06 1.23+0.06−0.05 ··· ··· ··· ··· ··· ··· ··· Table 5 continued 10 Masuda Figure 3. Dynamical models of the observed TTVs and TDVs of Kepler-448b. Red solid lines show 50 models randomly sampled from the joint posterior distribution obtained by fitting the data (black circles with error bars). The left column corresponds to the model which allows the mutual inclination to vary, while the right column shows the result when it is fixed to be zero. Table 5 (continued) Parameter Solution 1 Solution 2 Combined Transit impact parameter of Kepler-448b Occultation impact parameter of Kepler-448b 0.372(6) 0.22 ± 0.02 0.371(6) 0.21 ± 0.02 0.371(6) 0.22 ± 0.02 Prior ··· ··· Log evidence lnZ from Multinest −168.16 ± 0.09 −168.04 ± 0.09 −167.40 ± 0.09 Table 5 continued Table 5 (continued) 11 Parameter Solution 1 Solution 2 Combined Prior Note -- The elements of the inner and outer orbits listed here are Jacobian osculating elements defined at the epoch BJD = 2454833 + 120. The quoted values in the 'Solution' columns are the median and 68% credible interval of the marginal posteriors. Parentheses after values denote uncertainties in the last digit. The 'combined' column shows the values from the marginal posterior combining the two solutions; no value is shown if the combined marginal posterior is multimodal. In the prior column, U(a, b) and Ulog(a, b) denote the (log-)uniform priors between a and b, f (x) = 1/(b − a) and f (x) = x−1/(ln b − ln a), respectively; G(a, b, c) means the asymmetric Gaussian prior with the central value a and lower and upper widths b and c. aThere also exists a solution with negative cos i1. The solution is statistically indistinguishable from the one reported here, except that the signs of cos i2 and Ω21 are flipped (and thus i21 remains the same; see Eqn. A15). In principle, the TTVs for the solutions with cos i1 > 0 and cos i1 < 0 are not completely identical (see, e.g., A15 of Borkovits et al. 2015). However, the difference is in practice negligibly small for a transiting system because the effect is proportional to cot i1. I confirmed that this is indeed the case by performing the same numerical fit with the prior on cos i1 replaced by U(−0.02, 0). b Referenced to the ascending node of the inner orbit, whose direction is arbitrary. c This parameter is not constrained by the data; posterior is identical to the log-uniform prior. dDerived from the posterior of R(cid:63) and mean and standard deviation of Rp/R(cid:63) from individual transits (Table 3). posterior from the mean transit light curve (Table 3). 4.2. Constraints from TTVs The mass and some elements of the outer orbit (τ2, q2/a1, e2) are well determined from TTVs in spite of the non-transiting nature of the companion. The rela- tionship between these parameters and observed TTV features can roughly be understood as follows, with the help of the analytic formula in Appendix A. Most of the information comes from the spike-like fea- ture around BJD ∼ 2454833 + 1050, which is caused by the close encounter of the outer body and thus de- fines the time of its periastron passage τ2 (the effect is represented by S and C functions in Equations A9 and A10; see also Figure B1). In addition, its duration ∆t is sensitive to both a2 and e2: the former determines the overall orbital time scale and the latter changes the fraction of time spent around the periastron. The ∆t may be roughly estimated using Kepler's second law as follows: ∆t ∼ P2(1 − e2 2π 1 − 4 π 2)3/2 (cid:90) +π/2 (cid:18) (cid:19) (cid:21)3/2 ∼ P1 (cid:20) a2(1 − e2)8/3π −π/2 (cid:39) P2 2 (cid:39) P1 2 e2 a1 dv2 (1 + e2 cos v2)2 (cid:19)3/2 (cid:18) q2 a1 , (7) 2 where its order of magnitude is not so sensitive to the rather arbitrary choice of the interval of integration (−π/2 to π/2). Since ∆t ∼ 300 days and P1 ∼ 20 days, this estimate gives q2/a1 ∼ 10. Finally, its amplitude Figure 4. Marginal posterior PDFs for the mutual inclina- tion between the inner and outer orbits from the dynamical TTV and TDV modeling (Sections 4 and 5). Top -- Kepler- 448. Bottom -- Kepler-693. 4.1. Adopted Parameters ρ(cid:63) = (converted 1.51+0.14−0.09 M(cid:12) and from the I adopt m(cid:63) = 0.758+0.12−0.11 g cm−3 radius r(cid:63) = 1.41 ± 0.06 R(cid:12)) as the priors based on the spectroscopic values by Bourrier et al. (2015). The values agree with, and are more precise than, the latest KIC values (Mathur et al. 2016): m(cid:63) = 1.386+0.093 −0.076 M(cid:12), −0.2874 g cm−3. r(cid:63) = 1.367+0.267 Note that I do not use the values from the joint analysis in Bourrier et al. (2015) because they are derived assuming a circular orbit for the inner transiting planet, which turns out unlikely from our dynamical modeling. I adopt bmean = 0.373 and σb = 0.006 based on the −0.067 R(cid:12), and ρ(cid:63) = 0.7647+0.09607 12 Masuda ∆A is given by (Equations A2 and A3) (cid:20) q2(1 + e2) (cid:21)−3/2 mc ∆A (cid:39) P1 2π AL1 (cid:39) P1 15 16π (8) assuming mb, mc (cid:28) m(cid:63). Combining ∆A ∼ 2 min with the above estimate q2/a1 ∼ 10, we find mc/m(cid:63) ∼ 10−2. These numbers are in reasonable agreement with those from the full dynamical modeling. m(cid:63) a1 , In addition to the spike, a long-term modulation due to "tidal delay" (Borkovits et al. 2003; Agol et al. 2005, represented by M function in Equation A8) is also visi- ble (blue curves in Figure B1): tidal force from the outer companion delays the inner binary period, depending on the distance between the two. This effect, combined with the short-term spike, allows for further constraints on τ2, ω2, e2, and P2. In principle, this additional con- straint enables the separate determination of a2 and e2 rather than only q2, although a2 is not well constrained because it is not clear whether the whole cycle of the outer orbit is covered given the only one periastron pas- sage observed. The analytic expressions A2 -- A7 show that the short- term modulation represented by S or C function is pro- duced only when e1 (cid:54)= 0 or i21 (cid:54)= 0. In fact, the observed TTVs alone are found to be fitted well either by the non- zero e1 (δecc in Equation A2) or non-zero i21 (δnoncopl) effects. This causes the anti-correlated degeneracy be- tween e1 and i21. Since e1 also affects the spike ampli- tude, we have two classes of solutions: high-e1 -- low-i21 -- low-mc solution and low-e1 -- high-i21 -- high-mc solution. The joint TTV/TDV model slightly favors the latter. The directions of the orbits, ω1, ω2, cos i2, Ω21 are not well constrained from the TTVs alone. The complicated degeneracy between ω2 and Ω21 comes from the fact that TTVs are rather sensitive to such "dynamical" angles referred to the invariant plane as n1 and n2 in figure 1 of Borkovits et al. (2015). Nevertheless, I use cos i2 and Ω21 because of the simplicity in implementing the isotropic prior. The relationship between the two sets of angles are summarized in Appendix A.2. Finally, the LTTE effect (Equation A1) is about 0.1 min(a2/au) for mc/m(cid:63) ∼ 10−2. This term there- fore does not play a major role, as also seen in Figure B1. 4.3. Constraints from TTVs and TDVs In the joint analysis, I find a smaller e1 than that de- rived from the TTVs alone, because the combination of T , b, and ρ(cid:63) favors a small e1 sin ω1 (see equations 16, 18, and 19 of Winn 2011). This constraint is combined with those from TTVs to better determine e1 and ω1. The distribution of ω1 is still bimodal because no constraint is available for e1 cos ω1 from TDVs. These additional constraints on e1 and ω1 partly solve the degeneracies with other angles mentioned above. In particular, the value of mc from the joint TTV/TDV analysis is larger than from the TTVs alone, because the low-e1 -- high-i21 -- high-mc solution is favored. Absence of significant TDVs only weakly constrains the mutual inclination. As mentioned, the solution with i21 ∼ 90◦ is disfavored; if this were the case, a large tangential perturbation should have produced signifi- cant TDVs around the periastron passage, which are not present in the data. The solutions with Ω21 ∼ 0◦ and Ω21 ∼ 180◦ are basically indistinguishable and the log-likelihood is not sensitive to cos i2 either. In the TDV models (bottom left panel of Figure 3), both the positive and negative bumps are seen around the outer periastron because the inner inclination is perturbed in the opposite ways depending on the sign of cos i2. The current data do not favor either of the cases significantly. 5. RESULTS: KEPLER-693/KOI-824 The resulting posterior PDF from the MultiNest analysis and the corresponding models are shown in Figures 5 (red solid lines) and Figure C3. The sum- mary statistics (median and 68% credible region) of the marginal posteriors as well as the priors adopted for those parameters are given in Table 6. Again I found two almost identical solutions with prograde and retro- grade orbits. Similarly to Kepler-448c, the outer companion (we tentatively call it Kepler-693c) orbits close to the in- ner WJ in a highly eccentric orbit, except that it has a larger mass 1.5+0.6−0.4 × 102 MJup and can be a low-mass star. The outer pericenter distance of 1.5± 0.2 au is also similar to Kepler-448c and satisfies the stability condi- tion (Mardling & Aarseth 2001). Again non-zero eccen- tricity is favored for the inner orbit (e1 = 0.2+0.2−0.1), and its mass is not constrained at all from the data. A notable difference from the previous case is the clear TDV signal, which tightly constrains the mutual incli- nation to be i21 − 90◦ (cid:39) 40◦ (Figure 4, bottom). I checked that the data can never be explained by the aligned configuration (right column of Figure 5): I find the Bayesian evidence for the mutually-inclined model (lnZ = −89.18 ± 0.08) is larger than that of the copla- nar one (lnZcopl = −103.1 ± 0.07) by a factor of 106. The observed mutual inclination is the largest among those dynamically measured for planetary systems, and likely above the critical angle for the Kozai oscillation (posterior probability that 39.◦2 < i21 < 140.◦8 is 80%).3 3 Santerne et al. (2014) inferred the orbital inclination of the binary companion of the transiting WJ host KOI-1257/Kepler- 420 to be i2 = 18.◦2+18.◦0 −5.◦4 (68.3% interval), based on the trend in the SOPHIE RVs and bisector and full-width half maximum of In Section 6, I will show that the perturbation from the inclined companion does cause a large eccentricity oscil- lation of the inner WJ. Kepler-693b's mass is not measured, and the planet was statistically validated by Morton et al. (2016). A possible concern is that the absence of secondary eclipse, which the statistical validation partly relies on (in ad- dition to other factors including the transit shape and 13 non-detection of the companion via the AO imaging by Baranec et al. 2016), may lose its meaning if bocc is larger than unity due to the non-zero inner eccentricity. However, our dynamical modeling finds bocc = 0.64+0.08−0.09, which excludes the possibility. In addition, the derived mean stellar density is also compatible with the KIC value. Therefore, the low false positive probability (less than 1/3000) derived by Morton et al. (2016) is still valid in the light of our new constraints. Table 6. Parameters of the Kepler-693 System from the Dynamical TTV and TDV Analysis Parameter Solution 1 Solution 2 Combined Prior 173.611(+1) (−1) 173.610(+2) (−1) 173.610(+1) (−1) U(173.58, 173.64) Fitted Parameters (Inner Orbit) 1. Time of inferior conjunction tic,1 (BJDTDB − 2454833) 2. Orbital period P1 (day) 3. Orbital eccentricity e1 4. Argument of periastron ω1 (deg) 5. Cosine of orbital inclination cos i1 (Outer Orbit) a 6. Time of the periastron passage τ2 (BJDTDB − 2454833) 7. Periastron distance over inner semi-major axis q2/a1 8. Orbital eccentricity e2 9. Argument of periastron ω2 (deg) 10. Cosine of orbital inclination cos i2 11. Relative longitude of a ascending node Ω21 (deg)a,b (Physical Properties) 12. Mass of Kepler-693 m(cid:63) (M(cid:12)) 13. Mean density of Kepler-693 ρ(cid:63) (g cm−3) 15.37541(+10) (−7) 0.14+0.08−0.04 41+55−28 0.021(2) 15.37534(+10) (−8) 15.37537(+10) (−9) 0.3+0.2−0.1 174+13−39 0.022(3) 0.2+0.2−0.1 ··· 0.022(3) 640+20−17 640+22−18 640+22−17 13+2−1 13+1−2 13+2−2 0.48+0.12−0.06 41+20−20 −0.2+0.3−0.2 51+8−10 0.46+0.10−0.05 25+25−21 −0.3+0.2−0.1 −138+12−11 0.47+0.11−0.06 30+24−23 −0.3+0.2−0.2 ··· 0.80+0.03−0.03 2.2+0.3−0.2 0.80+0.04−0.03 2.2+0.3−0.3 0.80+0.04−0.03 2.2+0.3−0.3 Ulog(15.375, 15.376) U(0, 0.7) U (−180, 180) U(0, 0.04)a U(500, 900) Ulog(6, 25) U(0, 0.95) U(−180, 180) U(−1, 1) U (−180, 180) G(0.79, 0.03, 0.15) G(1.93, 0.18, 0.53) Ulog(0.1, 10) Ulog(0.001M(cid:12), 0.3M(cid:12)) Ulog(5 × 10−2, 50) Ulog(0.1, 102) 14. Mass of Kepler-693b mb (MJup)c 15. Mass of Kepler-693c mc (MJup) (Jitters) 16. Transit time jitter σTTV (10−4 day) 17. Transit duration jitter σdur (10−4 day) 0.8+2.7−0.6 167+59−43 2+4−1 9+14−8 1.0+3.2−0.7 136+50−34 1+4−1 9+13−8 0.9+3.0−0.7 145+58−37 1+4−1 9+14−8 Derived Parameters Outer orbital period P2 (day) (1.8+1.0−0.4) × 103 (1.8+0.6−0.3) × 103 (1.8+0.8−0.3) × 103 ··· Table 6 continued the stellar lines, combined with the Kepler light curve and stellar spectral energy distribution. If confirmed, the value translates into a large i21 (cid:38) 90◦ − i2 (Ω21 is not constrained at all) and a similar eccentricity evolution as discussed in the present paper is expected, although they caution that the outer binary inclination is poorly constrained, with the 99% limit on i2 being [8.◦2, 85.◦2]. 14 Masuda Figure 5. Dynamical models of the observed TTVs and TDVs of Kepler-693b. Red solid lines show 50 models randomly sampled from the joint posterior distribution obtained by fitting the data (black circles with error bars). The left column corresponds to the model which allows the mutual inclination to vary, while the right column shows the result when it is fixed to be zero. The coplanar model does not reproduce the observed transit durations. Table 6 (continued) Parameter Solution 1 Solution 2 Combined Inner semi-major axis a1 (au) Outer semi-major axis a2 (au) Periastron distance of the outer orbit a2(1 − e2) (au) Mutual orbital inclination i21 (deg) Physical radius of Kepler-693 (R(cid:12)) Physical radius of Kepler-693b (RJup)d Transit impact parameter of Kepler-693b Occultation impact parameter of Kepler-693b 0.112(+2) (−1) 2.9+1.1−0.5 1.5 ± 0.2 53+7−9 0.80 ± 0.03 0.91 ± 0.05 0.58 ± 0.04 0.68+0.06−0.06 0.112(+2) (−2) 2.8+0.7−0.4 1.5 ± 0.2 134+11−10 0.80 ± 0.04 0.91 ± 0.05 0.59 ± 0.05 0.62+0.09−0.08 0.112(+2) (−1) 2.8+0.8−0.4 1.5 ± 0.2 ··· 0.80 ± 0.04 0.91 ± 0.05 0.59 ± 0.05 0.64+0.08−0.09 Log evidence lnZ from Multinest −90.30 ± 0.08 −89.57 ± 0.08 −89.18 ± 0.08 Prior ··· ··· ··· ··· ··· ··· ··· ··· Table 6 continued Table 6 (continued) 15 Parameter Solution 1 Solution 2 Combined Prior Note -- The elements of the inner and outer orbits listed here are Jacobian osculating elements defined at the epoch BJD = 2454833 + 170. The quoted values in the 'Solution' columns are the median and 68% credible interval of the marginal posteriors. Parentheses after values denote uncertainties in the last digit. The 'combined' column shows the values from the marginal posterior combining the two solutions; no value is shown if the combined marginal posterior is multimodal. In the prior column, U(a, b) and Ulog(a, b) denote the (log-)uniform priors between a and b, f (x) = 1/(b − a) and f (x) = x−1/(ln b − ln a), respectively; G(a, b, c) means the asymmetric Gaussian prior with the central value a and lower and upper widths b and c. aThere also exists a solution with negative cos i1. The solution is statistically indistinguishable from the one reported here, except that the signs of cos i2 and Ω21 are flipped (and thus i21 remains the same; see Eqn. A15). In principle, the TTVs for the solutions with cos i1 > 0 and cos i1 < 0 are not completely identical (see, e.g., A15 of Borkovits et al. 2015). However, the difference is in practice negligibly small for a transiting system because the effect is proportional to cot i1. I confirmed that this is indeed the case by performing the same numerical fit with the prior on cos i1 replaced by U(−0.04, 0). b Referenced to the ascending node of the inner orbit, whose direction is arbitrary. c This parameter is not constrained by the data; posterior is identical to the log-uniform prior. dDerived from the posterior of R(cid:63) and that of Rp/R(cid:63) from the mean transit light curve (Table 4). 5.1. Adopted Parameters (corresponding I adopt m(cid:63) = 0.793+0.054 −0.1773 g cm−3 to −0.062 R(cid:12)) from KIC as the priors. −0.029 M(cid:12) and ρ(cid:63) = 1.931+0.526 r(cid:63) = 0.833+0.033 I adopt bmean = 0.6 and σb = 0.1 that roughly incorporates the 68% credible region of the posterior from the mean transit light curve (Table 4). 5.2. Constraints from TTVs The situation is basically similar to the Kepler-448 case, except that the TTV amplitude is larger by al- most an order of magnitude and so is the companion's mass. In addition, the whole shape of the short-term feature is not entirely observed due to the data gap, as shown in the large scatter of the models (Figure 5). This explains a weaker constraint on q2/a1, which is mainly determined by the duration of the feature (Section 4.2). 5.3. Constraints from TTVs and TDVs It is also the case in this system that the duration data point to a lower inner eccentricity than favored by the TTVs alone and increase the estimated mass. A striking difference is the clear trend in the duration that points to non-zero i21. The trend comes from the nodal precession of the inner orbit (Miralda-Escud´e 2002) and has also been observed in the Kepler-108 system recently (Mills & Fabrycky 2017). While a similar duration drift can also be caused by the apsidal precession of the ec- centric inner orbit, this is unlikely to be the case given the failure of the coplanar model. In fact, both pre- cession frequencies are basically comparable, but the ef- fect of the apsidal precession on the duration is smaller than that of the nodal precession by a factor of R(cid:63)/a1 (Miralda-Escud´e 2002). This difference can be signifi- cant for WJs with a/R(cid:63) > 10. 5.3.1. Mutual Inclination from the Likelihood Profile The constraint on the mutual inclination in Figure 4 and Table 6 is based on the one-dimensional Bayesian posterior marginalized over the other parameters. This depends not only on the likelihood and prior function but also on volume of the parameter space with high posterior probabilities. Thus, the resulting credible in- terval is generally not the same as the likelihood-based confidence interval even for a uniform prior, and a low value in the marginal posterior probability does not nec- essarily mean a poor fit to the data (Feroz et al. 2011). To see what value of the mutual inclination is "ex- cluded" by the data alone in the frequentist sense, I also estimate the confidence interval for i21 based on the likelihood profile. Specifically, I derive the maxi- mum likelihood L(i21) for each fixed value of i21 opti- mizing the other parameters, and examine its form as a function of i21. This has been done by searching for minimum χ2 solutions for a grid of cos i2 ∈ [−1, 1] and Ω21 ∈ [−90◦, 90◦] (prograde solutions). Here the con- straints on m(cid:63) and ρ(cid:63) are also incorporated in χ2, and σTTV and σdur were fixed to be zero. L), Figure 6 shows the resulting profile of −2 ln( L(i21)/ L denotes the maximum likelihood found by op- where timizing all the model parameters including i21. This is equivalent to the chi-squared difference from its min- imum value, ∆χ2, in our current setting. Here the ∆χ2 value is scaled so that the minimum χ2 solution has χ2 ν = 1. The resulting 1σ "confidence" interval is ◦ i21 = 47◦+17 ◦, and the 2σ lower bound is found to be −15 16 Masuda are set to be 1.25 days for Kepler-448, 10 days for Kepler- 693, and 1 day for the inner planets, although the spin evolution does not affect the result significantly. I adopt standard values for the dimensionless moments of iner- tia (0.059 and 0.25 for the star and inner planet) and the tidal Love numbers (0.028 and 0.5). The orbits are integrated for 10 Myr (sufficiently longer than the oscil- lation timescale; see below), starting from 1000 random sets of parameters sampled from the posterior distribu- tion from the dynamical analyses (Sections 4 and 5). Figure 7 shows the initial (black circles) and minimum (red diamonds) values of a1(1 − e2 1) over the course of evolution, the latter of which correspond to the max- ima of e1. In both systems, significant eccentricity os- cillations occur at least for some of the solutions. If we adopt a(1 − e2) < 0.1 au as a conventional thresh- old for the migrating WJs (Socrates et al. 2012b; Dong et al. 2014; Dawson et al. 2015), the periastrons be- come close enough to drive the tidal migration for 12% of the solutions for Kepler-448b and 96% for Kepler-693b, excluding the tidally-disrupted cases shown with trans- parent colors. If we choose a(1 − e2) < 0.05 au instead as a threshold (e.g., Anderson et al. 2016), the frac- tions become 5% and 33% for Kepler-448b and Kepler- 693b, respectively. Large eccentricity oscillations (and hence small minimum periastrons) are observed mainly for 40◦ (cid:46) i21 (cid:46) 140◦; this explains why a larger frac- tion of solutions have sufficiently small a(1 − e2) for the Kepler-693 system. The current eccentricities of the two WJs (1σ region shown with vertical dotted lines in Figure 8) also turn out to be the most likely values to be observed over the course of oscillation if we are random observers in time; they are around the peaks of the inner eccentricity distribution during the 10 Myr evolution averaged over the dynamical solutions (gray histograms in Figure 8). For highly inclined solutions, both libration and circu- lation of the argument of periastron (Kozai 1962), mod- ified by the octupole effects, are observed.4 In either of the two systems, the Kozai-Lidov timescale τKL (Kise- leva et al. 1998) and the octupole one τOct are both found to be much shorter than that of general rela- tivistic apsidal precession τGR for any solution found from TTVs and TDVs: for reference, I find typical τKL ∼ 10−2 Myr, τOct ∼ 10−1 Myr, and τGR ∼ 100.5 Myr for Kepler-448, and τKL ∼ 10−3 Myr, τOct ∼ 10−1.5 Myr, and τGR ∼ 100.5 Myr for Kepler-693. Thus the pertur- bation from the "close friends" is indeed strong enough to overcome general relativity (Dong et al. 2014). Note 4 On the other hand, the libration of the difference in the apsidal longitudes ∆inv, as discussed in Dawson & Chiang (2014), was not observed in the current simulations. Figure 6. Chi-squared difference from the global best-fit value as a function of the mutual inclination between the inner and outer orbits. The horizontal dashed lines show ∆χ2 = 1, 4, and 9, which correspond to 1σ, 2σ, and 3σ confidence intervals, respectively. i21 = 18◦. Thus I conclude that a high mutual inclina- tion is indeed favored by the data and the conclusion is insensitive to the prior information on i21. 6. LONG-TERM ORBITAL EVOLUTION A high mutual inclination confirmed for the Kepler- 693 system may result in a large eccentricity oscillation of the inner orbit as long as the perturbation is strong enough to overcome other short-range forces. If this is indeed the case, the system may serve as a direct piece of evidence that some WJs are undergoing eccentric- ity oscillations. Even in the Kepler-448 system where highly-inclined solutions are not necessarily favored, sig- nificant eccentricities of both the inner and outer orbits may still lead to the eccentricity excitation due to the octupole-level interaction. Given the full set of parameters constrained from the dynamical analysis, we can assess the future of the sys- tems rather reliably by extrapolating the dynamical so- lutions. In this section, we explore the effect of the sec- ular perturbation due to the outer companions on the inner planet's orbit and its tidal evolution. 6.1. Oscillation of the Inner Eccentricity I compute secular evolution of both the inner and outer orbits along with the spins of the star and inner planet. I use the code developed and utilized in Xue & Suto (2016) and Xue et al. (2017), which takes into account (i) gravitational interaction up to the octupole order and (ii) precessions due to general relativity as well as tidal and rotational deformation of the bodies. Here I neglect magnetic braking of the star and tidal dis- sipation inside the bodies, and assume zero stellar and planetary obliquities for simplicity. The rotation periods that this is the case for the octupole effect as well, which explains why some low-i21 solutions in the Kepler-448 system lead to a large eccentricity oscillation. 6.2. Migration Timescales Are the two WJs currently migrating into HJs due to the eccentricity oscillation? If this is the case, the current migration timescale τmig needs to be compara- If τmig (cid:29) τage they are ble to the system age τage. unlikely to be migrating, while if τmig (cid:28) τage the WJs should have evolved into HJs rapidly and we are un- likely to observe the system in the current state. Here we perform an order-of-magnitude comparison between the two timescales, given their large observational and theoretical uncertainties. Figure 9 shows the distribution of tidal migration timescale given by equation 2 of Petrovich & Tremaine (2016) at the minimum periastron distance computed in Section 6.1. Strictly speaking, the relevant timescale is expected to be longer because oscillation of e1 slows down the migration (Petrovich 2015), but in our case the maximum eccentricity is not so close to unity (since a1 is already small) that the modification is unlikely large (cf. figure 2 of Petrovich 2015); thus we simply neglect the correction. The timescales are computed for three different values of the viscous time of the planet, which gives a characteristic timescale for dissipation: tV = 0.015 yr, tV = 1.5 yr, and tV = 150 yr, while the dissipation inside the star is neglected (see Socrates et al. (2012a) for comparison with different parameter- izations). The three values roughly correspond to (i) very efficient tidal dissipation required for some high- eccentricity migration scenarios to explain the observed HJs (Petrovich 2015; Hamers et al. 2017), (ii) dissipation required to circularize the orbits of HJs with P ≤ 5 days within 10 Gyr (Socrates et al. 2012a), and (iii) values cal- ibrated based on a sample of eccentric planetary systems (Hansen 2010; Quinn et al. 2014), while the limit from the Jupiter -- Io system (tV (cid:38) 15 yr) lies in between the latter two. I also indicate (rough) estimates for the ages of the two systems with vertical dashed lines: 1.5 Gyr for Kepler-448 based on spectroscopy (Bourrier et al. 2015) and 5 Gyr for Kepler-693 as a tentative value given that the host star has dimensions of a K dwarf. The comparison between the histogram and the dashed line shows that the migrating solutions with τmig ∼ τage exist for a wide range of tV for the Kepler- In particular, the eccentricity oscillation 693 system. plays a crucial role for tV (cid:38) 1.5 yr so that such solu- tions realize. The migrating solutions also exist for the Kepler-448 system, though they seem plausible only for a small fraction of significantly misaligned solutions that lead to the large eccentricity oscillation, or require effi- cient tidal dissipation with tV (cid:46) 1.5 yr. 17 6.2.1. Possible Fates of the Inner Planets The timescale arguments above indicate the inner WJs may evolve into HJs within the lifetime of the system. If this is the case, HJ systems with close substellar com- panions as found by a long-term RV monitoring (Triaud et al. 2017) may have been WJs like ours in the past. As a proof of concept, I compute the evolution of the two systems for 1 Gyr, including the tidal dissipation with the planetary viscous timescale tV = 1.5 yr for an illustration. I stop the calculation when the inner orbit is circularized (both a1 < 0.1 au and e1 < 0.01 are achieved) before 1 Gyr. If we change tV, we expect things just happen on a differ- ent timescale corresponding to the change. I fix tV = 50 yr for the star. Figure 10 compares the initial and final semi-major axes from those simulations. We see that some solu- tions do survive the tidal disruption and evolve into HJs within 1 Gyr. Such an outcome is rarer in the Kepler-693 system than in the Kepler-448 system. This is consistent with the expectation that the former system is likely older than the latter at least by a factor of a few. In fact, this kind of path may be even rarer than it seems in the right histograms of Figure 10, because some of the survived HJs (shown with bluer colors) have expe- rienced the circularization on a much shorter timescale compared to the system age due to the rapid eccentricity surge: if we are random observers in time, it is a priori unlikely to observe a system with such a short remaining lifetime compared to the current age (Gott 1993). How- ever, I do not attempt to correct for this effect here, given that the outcome will be sensitive to the uncer- tain tidal parameter in any case. If correctly taken into account, this kind of argument will potentially allow for better constraints on the system parameters. 6.3. Implications for the In Situ Formation Scenario While the observed properties of the two inner planets are consistent with those of migrating WJs, presence of the outer (sub)stellar companions as close as 1.5 au chal- lenges the high-eccentricity migration scenario from be- yond the snow line. Planets in S-type orbits around tight binaries with periastron distances less than 10 au have also been reported around KOI-1257 (Santerne et al. 2014), Kepler-444 (Dupuy et al. 2016), HD 59686 (Ortiz et al. 2016), and possibly ν Octantis (Ramm et al. 2016). If confirmed to be a low-mass star, Kepler-693c has the smallest peristron among such stellar companions. A similar issue has also been discussed for WJs with outer planetary-mass companions (Antonini et al. 2016): the outer orbits in these systems, if primordial, are in most cases too small for the inner WJs to have migrated from (cid:38) 1 au. In addition, population synthesis simula- tions of high-eccentricity migration from (cid:38) 1 au, either 18 Masuda (a) Kepler-448 (b) Kepler-693 Figure 7. Effect of secular eccentricity oscillation on the semi-latus rectum of the inner orbit, a1(1 − e2 1). This distance corresponds to the final semi-major axis of the HJ if the orbit is circularized at a fixed angular momentum. The black circles and histograms are randomly sampled from the posterior of the dynamical TTV/TDV modeling and represent the current values. The red diamonds and histograms are the minimum values over many oscillation timescales (Section 6.1). Solutions with minimum periastron distance a1(1 − e1) less than the Roche limit (here chosen to be 2.7 R(cid:63) based on Guillochon et al. 2011, simply assuming the planetary mass 10−3 M(cid:63) and radius 0.1 R(cid:63)) are plotted with transparent colors and not included in the histogram. The horizontal dashed line indicates a1(1 − e2 1) = 0.1 au, which is the conventional threshold for possible tidal circularization accepted by Socrates et al. (2012b), Dong et al. (2014), and Dawson et al. (2015). (a) Kepler-448b (b) Kepler-693b Figure 8. The distributions of the inner eccentricities (e1) during the secular oscillation averaged over the dynamical solutions. The vertical dotted lines (shaded region) show the median and 68% credible interval of the current e1 measured from the TTVs and TDVs. via the companion on a wide orbit (Anderson et al. 2016; Petrovich & Tremaine 2016) or secular chaos in multiple systems (Hamers et al. 2017), have difficulty in produc- ing a sufficient number of WJs relative to HJs. These may also argue for the WJ formation via disk migra- tion or in situ. Considering the prevalence of compact super-Earth systems revealed by Kepler, the latter can well be possible theoretically (Lee et al. 2014) and may also have observational supports (Huang et al. 2016). The companions discovered in the Kepler-448 and Kepler-693 systems may further argue for the in situ origin. Such low-mass stellar or brown-dwarf compan- ions on au-scale orbits may be formed via fragmenta- tion at a larger separation followed by the orbital decay due to dissipative dynamical interactions involving gas accretion and disks, which proceed in (cid:46) 1 Myr (Bate et al. 2002; Stamatellos & Whitworth 2009; Bate 2012). This implies that giant-planet formation and migration 19 (a) Kepler-448 (b) Kepler-693 Figure 9. Migration timescales of the inner WJs for three different viscous timescales tV, computed with equation 2 of Petrovich & Tremaine (2016). The red and black histograms show the values at the minimum periastron over the course of eccentricity oscillation (Section 6.1 and Figure 7) and at the initial (current) periastron, respectively. The vertical dashed lines indicate the current age of the system estimated from spectroscopy (for Kepler-448) and speculated from the host-star property (for Kepler-693). (a) Kepler-448 (b) Kepler-693 Figure 10. Inner semi-major axes versus current mutual inclinations after 1 Gyr evolution in the presence of tidal dissipation with the planetary tV = 1.5 yr and stellar tV = 50 yr (diamonds and red histograms), along with their initial values (black circles and histograms). The color of each point shows the stopping time of the simulations, tend, defined as the shorter of 1 Gyr and the time when the inner orbit is circularized (a1 < 0.1 au and e1 < 0.01). Models with minimum periastrons inside the Roche limit (2.7 R(cid:63)) are shown with transparent colors and not included in the histograms. must have completed very quickly if they preceded those of the outer companion. Alternatively, the companions may have arrived at the current orbit well after disk mi- gration and disk dispersal via chaotic dissolution of an initial triple-star system or an impulsive encounter with a passing star (Marzari & Barbieri 2007; Mart´ı & Beaug´e 2012). While these scenarios are compatible with the ec- centric and inclined outer orbit, they may suffer from the fine-tuning problem. In the former scenario, for exam- ple, a binary orbit typically shrinks only by a factor of a few, limited by energy conservation (Marzari & Bar- bieri 2007). The outcomes are likely more diverse in the latter, but only a small fraction of them usually consti- tutes a suitable solution (Mart´ı & Beaug´e 2012), and such close encounters as to alter the binary orbit signifi- cantly are likely rare when the planet formation is com- Masuda 20 pleted, even in a cluster with ∼ 103 stars (Adams et al. 2006). Also note that tidal friction associated with the close encounter with the primary (e.g., Kiseleva et al. 1998) is unavailable to shrink the binary orbit in the presence of the inner planet. Considering these possible difficulties of the alternative scenarios, in situ formation in a tight binary seems to be an attractive possibility that provides simple solutions both for our two systems and for other theoretical and observational issues, al- though the disk migration followed by rearrangement of the outer orbit cannot be excluded. This motivates us to examine whether the moderate eccentricities of our WJs can be explained in the in-situ framework, in which a near-circular orbit is normally ex- pected. Here we consider a specific form of question, in the same spirit as Anderson & Lai (2017):5 suppose that the inner WJs were produced into circular orbits with the current semi-major axes, can their observed non-zero eccentricities be explained by the perturbation from the detected companions? To see this, I perform a similar set of simulations as in Section 6.1 setting e1 = ω1 = 0 initially, while sampling the other parameters from the posterior. Note that this experiment is applicable to the disk migration case as well, whose outcome is also a short-period planet on a circular orbit. 7. DISCUSSIONS 7.1. Follow-up Observations While the future observations of transit times and du- rations will surely improve the constraint on the outer period and eccentricity, I did not find any systematic de- pendence of the future TTV and TDV evolutions on the current mutual orbital inclination at least within about 10 years. Moreover, it is challenging to observe a full transit from the ground, due to the long transit dura- tion ((cid:39) 7 hr) for Kepler-448b and faintness (V = 16.8) of the host star for Kepler-693b. How about RV observations? Based on the constraints from TTVs and TDVs, the RV semi-amplitude due to the outer companion is expected to be 2.7+0.7−0.6×102 m s−1 for Kepler-448 and 2.7+0.9−0.6 km s−1 for Kepler-693. The variation may be observable for Kepler-693 with a large telescope, while the detection is implausible for Kepler- 448 with v sin i(cid:63) (cid:39) 60 km s−1. Instead, Gaia astrometry (Perryman et al. 2001) is promising to detect the outer binary motion and bet- ter determine the mutual inclination. Since the mis- alignment in the sky plane Ω21 is dynamically well constrained in both systems (and the inner planets are transiting), if measured, significantly improve the constraint on i21. Based on the dynamical modeling, the expected astro- metric displacements due to the outer companions are 1.4+1.2−0.5×102 µas for Kepler-448 with V = 11.4 (assuming the distance d = 426 ± 40 pc from Bourrier et al. 2015) and 4.4+3.4−1.5 × 102 µas for Kepler-693 with V = 16.8 (as- suming d = 1110 pc from the isochrone fit). They are both well above the astrometric precision expected after the nominal five-year mission (Perryman et al. 2014). inclinations of the outer orbits i2, 7.2. Frequency of Close and Massive Companions to WJs It is admittedly difficult to evaluate the completeness of our TTV search due to the complex dependence of the signal on the system parameters. Nevertheless, our detection of the close and massive companions in two systems, among the sample of 23 WJs, suggests such an architecture is not extremely uncommon. We also need to take into account that the two systems would not have been detected if the periastron passage did not occur during the Kepler mission. The ratios of the outer orbital periods (P2 ∼ 2500 days for Kepler-448c and P2 ∼ 1800 days for Kepler-693c) to the Kepler ob- serving duration (∼ 1400 days) suggest that there may be one or two more WJs with similar companions in the Kepler sample that have evaded our search. This crude estimate seems compatible with the theoretical argument by Petrovich & Tremaine (2016) that roughly 20% of WJs can be accounted for by high-eccentricity (cid:113) 1,max (cid:112)3/5 = cos i21,init. Figure 11 summarizes the maximum inner eccentric- ities achieved during the 10 Myr evolution against the current mutual inclination. The maximum value ex- ceeds the current best-fit value (horizontal dashed line) for 14% and 76% of the solutions that did not lead to tidal disruption in the Kepler-448 and Kepler-693 systems, respectively. Thus the current architecture can indeed be compatible with the initially circular orbits. The sequences of the maximum e1 (e1,max) and initial i21 (i21,init) roughly follows the relation 1 − e2 Considering the arguments in Section 6.2, it is also conceivable that the inner WJs were formed into circu- lar orbits at larger semi-major axes than observed now (but still inside the snow line) and have migrated to the current orbits via the tidal migration. In this case, ex- citation of eccentricity should have been easier because the gravitational interaction with the companion was initially stronger. This process might serve as yet an- other path of HJ formation: some HJs may have been isolated WJs formed in situ or via disk migration, whose orbits were shrunk due to the tidal high-eccentricity mi- gration driven by a close companion. 5 See Mustill et al. (2017) for other possible pathways of eccen- tric WJ formation. 21 (a) Kepler-448 (b) Kepler-693 Figure 11. Maximum eccentricities of the inner WJs achieved during secular evolution started from circular orbits against the current mutual orbital inclinations. The horizontal dashed lines show the median eccentricities from the dynamical modeling. Models with minimum periastrons inside the Roche limit are shown with transparent colors and not included in the histograms. migration, although the observed architectures of our systems may not support the migration from (cid:38) 1 au via this process as assumed in Petrovich & Tremaine (2016). 8. SUMMARY AND CONCLUSION This paper reported the discovery of close companions to two transiting WJs via their TTVs and TDVs. The companions have masses comparable to a brown dwarf or a low-mass star (22+7−5 MJup and 150+60−40 MJup), and they are on highly eccentric orbits (e (cid:38) 0.5) with small periastron distances ((cid:39) 1.5 au). For the companion of Kepler-693b, a large mutual orbital inclination ((cid:39) 50◦) with respect to the inner planetary orbit is indicated by TDVs, while the constraint on the mutual inclina- tion is weak for the Kepler-448 system. They are among the few systems with constraints on mutual inclinations, and that inferred for the Kepler-693 system is the largest ever determined dynamically for planetary systems. The value is indeed large enough for the eccentricity oscilla- tion via the Kozai mechanism to occur: more than 90% of the solutions inferred for Kepler-693b (and some 10% for Kepler-448b) imply that the inner WJs' eccentrici- ties exhibit a large enough oscillations for tidal dissipa- tion to affect the inner orbits significantly, by bringing a(1 − e2) less than 0.1 au. The corresponding migration timescales can be compatible with the hypothesis that the inner WJs are tidally migrating to evolve into HJs, for a wide range of viscous timescales. The architectures of the two systems support the sce- nario that eccentric WJs are currently undergoing ec- centricity oscillation induced by a close companion and experiencing the slow tidal migration where the orbit shrinks only at the high-eccentricity phase. On the other hand, the origin of the current highly eccentric/inclined configuration is still unclear. Specifically, they may not fit into the classical picture that close-in gas giants have migrated from beyond the snow line, given the close and (sub-)stellar nature of the outer companions. Forma- tion of gas giants within the snow line onto circular or- bits, followed by eccentricity excitation by the compan- ion, therefore seems another viable option to be pur- sued, although the companion may instead have arrived at the current orbit after disk migration of the inner WJ. Regardless of the origin of the current configura- tion, the long-term evolution simulation demonstrates a new pathway of HJ formation via "high-eccentricity" migration of a WJ formed in situ or via disk migration. The author is grateful to the entire Kepler team for making the revolutionary data available. The au- thor thanks Hajime Kawahara, Yasushi Suto, Yuxin Xue, Saul Rappaport, Josh Winn, Sho Uehara, Sarah Ballard, Gongjie Li, Tim Morton, Dong Lai, Kaloyan Penev, Adrian Price-Whelan, Chelsea Huang, George Zhou, Bekki Dawson, Cristobal Petrovich, Amaury Tri- aud, Kengo Tomida, Kaitlin Kratter, Eve Ostriker, and Ji-Ming Shi for helpful conversations and insightful discussions; and Tam´as Borkovits for answering ques- tions on the analytic ETV formula. Comments on the manuscript from Dong Lai, Josh Winn, and an anonymous referee were also very helpful in improving the paper and enriching its content. K.M. acknowl- edges the support by JSPS Research Fellowships for Young Scientists (No. 26-7182) and also by the Lead- 22 Masuda ing Graduate Course for Frontiers of Mathematical Sci- ences and Physics. This work was performed in part under contract with the California Institute of Technol- ogy (Caltech)/Jet Propulsion Laboratory (JPL) funded by NASA through the Sagan Fellowship Program exe- cuted by the NASA Exoplanet Science Institute. This paper includes data collected by the Kepler mis- sion. Funding for the Kepler mission is provided by the NASA Science Mission directorate. Some of the data presented in this paper were obtained from the Mikul- ski Archive for Space Telescopes (MAST). STScI is op- erated by the Association of Universities for Research in Astronomy, Inc., under NASA contract NAS5-26555. Support for MAST for non-HST data is provided by the NASA Office of Space Science via grant NNX09AF08G and by other grants and contracts. This research has made use of the NASA Exoplanet Archive, which is op- erated by the California Institute of Technology, under contract with the National Aeronautics and Space Ad- ministration under the Exoplanet Exploration Program. This research has made use of NASA's Astrophysics Data System Bibliographic Services. APPENDIX A. ANALYTIC TTV FORMULA FOR HIERARCHICAL TRIPLE SYSTEMS In a part of the TTV analysis, we adopt the analytic timing formula for hierarchical three-body systems by Borkovits et al. (2015), which takes into account the eccentricities of both inner and outer orbits and arbitrary mutual inclination between them (see also Borkovits et al. 2011, 2015, 2016, for its applications). Among various effects that could possibly affect the observed transit times, we include two of the most important effects: light-travel time effect (LTTE) and P2-timescale dynamical effects up to the quadrupole order. The other effects including the octupole-level dynamical effects and short-term perturbations are at least an order-of-magnitude smaller than the quadrupole terms and thus neglected (see Borkovits et al. 2015, for a complete discussion of these effects). Note that, in this appendix, +Z-axis is taken away from the observer's direction; this definition is opposite to the one in the main text, and so arguments of periastron in the formulae below differ by π from the ones in the main text. We model the timing variations from the linear ephemeris due the LTTE effect and the P2-timescale dynamical A.1. Formula effect. The LTTE term is given by (A1) where v2 is the true anomaly and u2 ≡ ω2 + v2 is the true longitude. The P2-timescale dynamical effect is modeled up to the quadrupole order as follows (Borkovits et al. 2015, equation 5): m(cid:63) + mb + mc c , ∆LTTE = − a2 sin i2 mc (1 − e2 2) sin u2 1 + e2 cos v2 Here the overall amplitude is fixed by the factor ∆1 = P1 2π AL1(1 − e2 1)1/2 [δtidal + δecc1 + δecc2 + δnoncopl] . and each TTV component is given by δtidal = AL1 = 15 8 m(cid:63) + mb + mc mc (cid:20) 8 (1 − e2 2)−3/2, P1 P2 (cid:21) f1 + 4 5 K1 15 M, δecc1 = (1 + cos i21){K11S [2u2 − 2(n2 − n1)] − K12C [2u2 − 2(n2 − n1)]} , δecc2 = (1 − cos i21){K11S [2u2 − 2(n2 + n1)] + K12C [2u2 − 2(n2 + n1)]} , (cid:18) δnoncopl = sin2 i21 where K11 cos 2n1 + K12 sin 2n1 − 2 5 f1 − 3 5 K1 [2M − S(2u2 − 2n2)] , M = v2 − l2 + e2 sin v2, (cid:20) S(2u2) = sin 2u2 + e2 sin(u2 + ω2) + sin(3u2 − ω2) 1 3 (cid:19) (cid:21) , (A2) (A3) (A4) (A5) (A6) (A7) (A8) (A9) C(2u2) = cos 2u2 + e2 cos(u2 + ω2) + f1 = 1 + 25 8 e2 1 + 15 8 e4 1 + cos(3u2 − ω2) 1 3 1 + O(e8 e6 1), (cid:21) , (cid:19) (cid:20) (cid:18) 3 4 e6 1 (cid:18) 23 (A10) (A11) (A12) (A13) (A17) (A18) (A19) (A20) (A21) (A22) 95 64 (cid:19) (cid:19) (cid:18) 1 7 64 (cid:18) 51 40 K1(e1, ω1) = − e1 sin ω1 + (cid:18) 5 − e4 1 + 3 16 16 (cid:19) 1 8 e4 1 + 3 e2 1 + 64 cos 4ω1 − 3 16 e6 1 cos 2ω1 + e5 1 sin 5ω1 + 3 16 sin 3ω1 e5 1 e3 1 + 2 1 cos 6ω1 + O(e7 e6 1), (cid:19) K11(e1, ω1) = K12(e1, ω1) = − e2 1 + 3 4 − 3 16 e4 1 + 3 3 16 32 1 sin 3ω1 − e3 (cid:18) e1 − 1 2 1 − 1 e3 4 1 cos 3ω1 − e3 − 3 16 e6 1 + 16 (cid:18) 1 (cid:19) (cid:18) 1 e5 1 e1 + 1 2 1 − 1 e4 16 cos ω1 + e6 1 e3 1 + (cid:19) (cid:18) 51 (cid:19) 40 241 640 cos 2ω1 e6 1 e4 1 + 1 cos 6ω1 + O(e7 e6 1), 37 80 3 64 sin ω1 + e5 1 1 4 cos 4ω1 − 1 16 e5 1 sin 5ω1 + e2 1 + (cid:19) e2 1 + 87 80 e4 1 + 541 640 e6 1 sin 2ω1 e4 1 + 5 32 e6 1 16 sin 4ω1 + 1 16 e5 1 cos 5ω1 + 1 sin 6ω1 + O(e7 e6 1). 3 64 (A14) The angles n1 and n2 are the directions of Z > 0 part of the line of intersection of the inner and outer orbits measured from the ascending nodes of the two orbits, defined as in figure 1 of Borkovits et al. (2015) between [0, π], and l2 is the mean anomaly. Note that K11(e1, π − ω1) = K11(e1, ω1) and K12(e1, π − ω1) = −K12(e1, ω1). A.2. Conversion of the Angles In the main body of the paper, we did not use the physically most natural parametrization of the angles because it complicates the assignment of the prior in the MultiNest fitting. Here we summarize how to relate our set of fitted angles (i1, i2, Ω21) to that of physical angles (i21, n1, n2) in the analytic formula by Borkovits et al. (2015), since this case is not covered in their Appendix D. The mutual inclination i21 can be computed as usual: (A15) Note that sin i21 (cid:54)= 0 for Ω21 (cid:54)= 0 since (i1, i2) (cid:54)= (0, 0) nor (π, π) for transiting systems as discussed in this paper. For cos Ω21 = ±1, the above equation reduces to cos i21 = cos i1 cos i2 + sin i1 sin i2 cos Ω21. (A16) Let us first consider the case when sin i21 (cid:54)= 0. If sin Ω21 (cid:54)= 0, the sine and cosine rules of the spherical trigonometry yield cos i21 = cos(i2 ∓ i1). sin i21 sin Ω21 = sin i2 sin n1 = sin i1 sin n2 and cos n1 = cos Ω21 cos n2 + sin Ω21 sin n2 cos i2, cos n2 = cos Ω21 cos n1 + sin Ω21 sin n1 cos(π − i1), for sin Ω21 < 0 case. For sin Ω21 > 0, the cosine rules are replaced by cos n1 = cos Ω21 cos n2 + sin Ω21 sin n2 cos(π − i2), cos n2 = cos Ω21 cos n1 + sin Ω21 sin n1 cos i1. In fact, the two cases can be written in a single form as cos n1 = cos Ω21 cos n2 − sin Ω21 sin n2 cos i2, 24 Masuda Thus, for a non-zero mutual inclination we have cos n2 = cos Ω21 cos n1 + sin Ω21 sin n1 cos i1. and sin n1,2 = sgn(sin Ω21) sin i2,1 sin i21 sin Ω21 = sin i2,1 sin i21 sin Ω21 cos n1 = sgn(sin Ω21) sin i21 (− sin i1 cos i2 + cos Ω21 cos i1 sin i2) , (A23) (A24) (A25) sgn(sin Ω21) (cos i1 sin i2 − cos Ω21 sin i1 cos i2) . cos n2 = (A26) If sin i21 = 0, we may just define n2 − n1 = 0 for i21 = 0 and n2 + n1 = π for i21 = π to correctly compute the non-vanishing terms of TTVs. Other than i1 = i2 = 0 or π, this includes the two cases: (i) i1 − i2 = 0 and Ω21 = 0 and (ii) i1 + i2 = π and Ω21 = π. As shown in Table A1, we have n1 − n2 = 0 for either i2 − i1 → ±0 in case (i). Similarly in case (ii), we have n1 + n2 = π for either i2 + i1 − π → ±0. Although the other combination is ambiguous, it appears only in the vanishing terms of the TTV formula. sin i21 cos Ω21 → 1 cos Ω21 → −1 sgn(sin Ω21) sgn(sin Ω21) (n1, n2) (0, 0) or (π, π) for i2 − i1 → 0 (0, π) or (π, 0) for i2 + i1 → π Table A1. cos n1,2 for sin Ω21 → 0 (cos n1, cos n2) (cid:16) sin(i2−i1) (cid:16)− sin(i2+i1) sin i21 sin i21 , sin(i2−i1) , sin(i2+i1) sin i21 sin i21 (cid:17) (cid:17) B. RESULTS OF ANALYTIC AND NUMERICAL TTV ANALYSES B.1. Comparison of the Two Results For TTVs, I performed both an analytic fit with a wider prior range and a numerical fit with a narrower prior range. As shown in Tables B2 and B3, I found consistent posteriors from the two analyses; this agreement validates our numerical procedure. The best-fit models are basically indistinguishable from those in Figures 3 and 5. Table B2. Parameters of the Kepler-448 System from the Analytical and Numerical TTV Analyses Parameter Posterior Prior Posterior Prior Analytic Numerical Fitted Parameters (Inner Orbit) 1. Time of inferior conjunction tic,1 (BJDTDB − 2454833) 2. Orbital period P1 (day) 3. Orbital eccentricity e1 4. Argument of periastron ω1 (deg) 5. Cosine of orbital inclination cos i1 (Outer Orbit) 6. Time of the periastron passage τ2 (BJDTDB − 2454833) 7. Periastron distance over inner semi-major axis q2/a1 128.7414(+7) (−3) U(128.7, 128.8) 128.7418(+2) (−2) U(128.73, 128.75) 17.855219(+6) (−13) 0.5+0.3−0.2 −43+158−104 0 (fixed) 1076+10−10 9.3+0.4−0.3 Ulog(17.85, 17.86) U(0, 0.95) U(−180, 180) ··· U(120, 1600) Ulog(1, 50) 17.855179(+7) (−6) 0.5+0.1−0.2 −15+166−140 0 (fixed) 1072+12−10 9.4+0.4−0.4 Ulog(17.8551, 17.8553) U(0, 0.7) U(−180, 180) ··· U(900, 1200) Ulog(5, 20) Table B2 continued Table B2 (continued) Parameter 8. Orbital eccentricity e2 9. Argument of periastron ω2 (deg) 10. Cosine of orbital inclination cos i2 11. Relative longitude of ascending node Ω21 (deg)a (Physical Properties) 12. Mass of Kepler-448 m(cid:63) (M(cid:12)) 13. Mass of Kepler-448b mb (MJup) 14. Mass of Kepler-448c mc (MJup) (Jitters) 15. Transit time jitter σTTV (10−4 day) Posterior 0.54+0.10−0.05 −91+52−29 0.1+0.7−0.8 −5+162−155 1.54+0.11−0.09 1 (fixed) 12+10−4 2.2 ± 0.3 Analytic Prior U(0, 0.95) U(−180, 180) U(−1, 1) U(−180, 180) G(1.51, 0.09, 0.14) ··· Ulog(10−4M(cid:12), 0.3M(cid:12)) Ulog(0.05, 5) 25 Numerical Prior U(0, 0.95) U(−180, 180) U(−1, 1) U(−180, 180) G(1.51, 0.09, 0.14) ··· Ulog(10−3M(cid:12), 0.1M(cid:12)) Ulog(0.05, 5) Posterior 0.60+0.12−0.07 −116+58−34 0.1+0.6−0.7 1+135−131 1.54+0.11−0.09 1 (fixed) 15+7−4 2.2 ± 0.3 Derived Parameters Outer orbital period P2 (day) Inner semi-major axis a1 (au) Outer semi-major axis a2 (au) Periastron distance of the outer orbit a2(1 − e2) (au) Mutual orbital inclination i21 (deg) (1.6+0.8−0.2) × 103 0.154(3) 3.1+0.9−0.3 1.44+0.07−0.06 118+31−68 ··· ··· ··· ··· ··· (2.0+1.4−0.4) × 103 0.154(3) 3.6+1.5−0.5 1.45+0.07−0.06 54+81−28 ··· ··· ··· ··· ··· Note -- The elements of the inner and outer orbits listed here are Jacobian osculating elements defined at the epoch BJD = 2454833 + 120. The quoted values in the 'Solution' columns are the median and 68% credible interval of the marginal posteriors. Parentheses after values denote uncertainties in the last digit. The 'combined' column shows the values from the marginal posterior combining the two solutions; no value is shown if the combined marginal posterior is multimodal. In the prior column, U(a, b) and Ulog(a, b) denote the (log-)uniform priors between a and b, f (x) = 1/(b − a) and f (x) = x−1/(ln b − ln a), respectively; G(a, b, c) means the asymmetric Gaussian prior with the central value a and lower and upper widths b and c. aReferenced to the ascending node of the inner orbit, whose direction is arbitrary. Table B3. Parameters of the Kepler-693 System from the Analytical and Numerical TTV Analyses Parameter Posterior Prior Posterior Prior Analytic Numerical Fitted Parameters (Inner Orbit) 1. Time of inferior conjunction tic,1 (BJDTDB − 2454833) 2. Orbital period P1 (day) 3. Orbital eccentricity e1 4. Argument of periastron ω1 (deg) 5. Cosine of orbital inclination cos i1 173.611(+2) (−1) 15.37565(+3) (−7) 0.6 ± 0.2 64+72−158 0 (fixed) U(173.55, 173.65) 173.609(+1) (−2) U(173.58, 173.64) Ulog(15.37, 15.38) U(0, 0.95) U(−180, 180) ··· 15.37549(+11) (−7) 0.5 ± 0.1 84+62−93 0 (fixed) Ulog(15.375, 15.376) U(0, 0.7) U(−180, 180) ··· Table B3 continued 26 Masuda Table B3 (continued) Parameter Posterior Prior Posterior Prior Analytic Numerical (Outer Orbit) 6. Time of the periastron passage τ2 (BJDTDB − 2454833) 7. Periastron distance over inner semi-major axis q2/a1 8. Orbital eccentricity e2 9. Argument of periastron ω2 (deg) 10. Cosine of orbital inclination cos i2 11. Relative longitude of ascending node Ω21 (deg)a (Physical Properties) 12. Mass of Kepler-693 m(cid:63) (M(cid:12)) 13. Mass of Kepler-693b mb (MJup) 14. Mass of Kepler-693c mc (MJup) (Jitters) 15. Transit time jitter σTTV (10−4 day) 636+28−16 13+2−2 0.50+0.18−0.08 60+58−62 0.0 ± 0.3 −6+167−157 0.80 ± 0.03 1 (fixed) 57+41−21 U(170, 1560) Ulog(1, 50) U(0, 0.95) U(−180, 180) U(−1, 1) U(−180, 180) G(0.79, 0.03, 0.05) ··· Ulog(10−4M(cid:12), 0.3M(cid:12)) 660+23−20 15+2−2 0.7+0.1−0.1 86+42−47 −0.1 ± 0.4 −21+178−139 0.80 ± 0.03 1 (fixed) 94+37−23 U(500, 900) Ulog(6, 25) U(0, 0.95) U(−180, 180) U(−1, 1) U(−180, 180) G(0.79, 0.03, 0.05) ··· Ulog(10−3M(cid:12), 0.3M(cid:12)) 6+4−3 Ulog(0.05, 50) 6+4−3 Ulog(0.05, 50) Derived Parameters Outer orbital period P2 (day) Inner semi-major axis a1 (au) Outer semi-major axis a2 (au) Periastron distance of the outer orbit a2(1 − e2) (au) Mutual orbital inclination i21 (deg) (2.0+2.5−0.6) × 103 0.112(+2) (−1) 3.0+2.1−0.6 1.5 ± 0.2 133+30−115 ··· ··· ··· ··· ··· (5.3+8.0−2.4) × 103 0.112(+2) (−1) 5.8+4.9−1.9 1.6 ± 0.2 133+30−105 ··· ··· ··· ··· ··· Note -- The elements of the inner and outer orbits listed here are Jacobian osculating elements defined at the epoch BJD = 2454833 + 170. The quoted values in the 'Solution' columns are the median and 68% credible interval of the marginal posteriors. Parentheses after values denote uncertainties in the last digit. The 'combined' column shows the values from the marginal posterior combining the two solutions; no value is shown if the combined marginal posterior is multimodal. In the prior column, U(a, b) and Ulog(a, b) denote the (log-)uniform priors between a and b, f (x) = 1/(b − a) and f (x) = x−1/(ln b − ln a), respectively; G(a, b, c) means the asymmetric Gaussian prior with the central value a and lower and upper widths b and c. aReferenced to the ascending node of the inner orbit, whose direction is arbitrary. B.2. Decomposition of the TTV Solutions Using the Analytic Formula The analytic formula allows us to understand how each physical effect in Equation A2 contributes to the observed TTVs. Figure B1 shows the decomposed sig- nals for each of the (i) "LTTE" ∆LTTE, (ii) "tidal" δtidal, (iii) "eccentric" δecc1 + δecc2, and (iv) "non-coplanar" δnoncopl terms for 10 solutions randomly sampled from the posterior obtained in the previous section. The plot shows that the δecc terms play a crucial role in producing the short-term feature, especially for Kepler-448b. C. CORNER PLOTS FOR THE POSTERIORS FROM DYNAMICAL ANALYSES Figures C2 and C3 show the corner plots of the pos- terior distributions obtained from the dynamical TTV and TDV analyses in Sections 4 and 5. The figures are generated using corner.py by Foreman-Mackey (2016). Adams, F. C., Proszkow, E. M., Fatuzzo, M., & Myers, P. C. Agol, E., Steffen, J., Sari, R., & Clarkson, W. 2005, MNRAS, 2006, ApJ, 641, 504 359, 567 REFERENCES 27 (a) Kepler-448b (b) Kepler-693b Figure B1. Decomposition of the analytic TTV models into individual terms in Equations A1 and A2. Anderson, K. R., & Lai, D. 2017, ArXiv e-prints, Feroz, F., Hobson, M. P., Cameron, E., & Pettitt, A. N. 2013, arXiv:1706.00084 ArXiv e-prints, arXiv:1306.2144 Anderson, K. R., Storch, N. I., & Lai, D. 2016, MNRAS, 456, Foreman-Mackey, D. 2016, The Journal of Open Source 3671 Antonini, F., Hamers, A. S., & Lithwick, Y. 2016, AJ, 152, 174 Baranec, C., Ziegler, C., Law, N. M., et al. 2016, AJ, 152, 18 Bate, M. R. 2012, MNRAS, 419, 3115 Bate, M. R., Bonnell, I. A., & Bromm, V. 2002, MNRAS, 336, 705 Batygin, K., Bodenheimer, P. H., & Laughlin, G. P. 2016, ApJ, 829, 114 Software, 24, doi:10.21105/joss.00024 Foreman-Mackey, D., Hogg, D. W., Lang, D., & Goodman, J. 2013, PASP, 125, 306 Goldreich, P., & Tremaine, S. 1980, ApJ, 241, 425 Gott, III, J. R. 1993, Nature, 363, 315 Guillochon, J., Ramirez-Ruiz, E., & Lin, D. 2011, ApJ, 732, 74 Hadden, S., & Lithwick, Y. 2016, ArXiv e-prints, arXiv:1611.03516 Boley, A. C., Granados Contreras, A. P., & Gladman, B. 2016, Hamers, A. S., Antonini, F., Lithwick, Y., Perets, H. B., & ApJL, 817, L17 Borkovits, T., Csizmadia, S., Forg´acs-Dajka, E., & Hegedus, T. 2011, A&A, 528, A53 Borkovits, T., ´Erdi, B., Forg´acs-Dajka, E., & Kov´acs, T. 2003, A&A, 398, 1091 Borkovits, T., Hajdu, T., Sztakovics, J., et al. 2016, MNRAS, Portegies Zwart, S. F. 2017, MNRAS, 464, 688 Hansen, B. M. S. 2010, ApJ, 723, 285 Hirano, T., Narita, N., Sato, B., et al. 2012, ApJL, 759, L36 Holczer, T., Mazeh, T., Nachmani, G., et al. 2016, ApJS, 225, 9 Huang, C., Wu, Y., & Triaud, A. H. M. J. 2016, ApJ, 825, 98 Jontof-Hutter, D., Ford, E. B., Rowe, J. F., et al. 2016, ApJ, 455, 4136 820, 39 Borkovits, T., Rappaport, S., Hajdu, T., & Sztakovics, J. 2015, MNRAS, 448, 946 Kipping, D. M. 2013, MNRAS, 435, 2152 Kiseleva, L. G., Eggleton, P. P., & Mikkola, S. 1998, MNRAS, Bourrier, V., Lecavelier des Etangs, A., H´ebrard, G., et al. 2015, 300, 292 A&A, 579, A55 Knutson, H. A., Fulton, B. J., Montet, B. T., et al. 2014, ApJ, Bryan, M. L., Knutson, H. A., Howard, A. W., et al. 2016, ApJ, 785, 126 821, 89 Buchner, J., Georgakakis, A., Nandra, K., et al. 2014, A&A, 564, A125 Kozai, Y. 1962, AJ, 67, 591 Lee, E. J., Chiang, E., & Ormel, C. W. 2014, ApJ, 797, 95 Lin, D. N. C., Bodenheimer, P., & Richardson, D. C. 1996, Coughlin, J. L., Mullally, F., Thompson, S. E., et al. 2016, Nature, 380, 606 ApJS, 224, 12 Dawson, R. I., & Chiang, E. 2014, Science, 346, 212 Dawson, R. I., Murray-Clay, R. A., & Johnson, J. A. 2015, ApJ, 798, 66 Dawson, R. I., Johnson, J. A., Fabrycky, D. C., et al. 2014, ApJ, 791, 89 Deck, K. M., Agol, E., Holman, M. J., & Nesvorn´y, D. 2014, ApJ, 787, 132 Dong, S., Katz, B., & Socrates, A. 2014, ApJL, 781, L5 Dupuy, T. J., Kratter, K. M., Kraus, A. L., et al. 2016, ApJ, 817, 80 Feroz, F., Cranmer, K., Hobson, M., Ruiz de Austri, R., & Lithwick, Y., Xie, J., & Wu, Y. 2012, ApJ, 761, 122 Mandel, K., & Agol, E. 2002, ApJL, 580, L171 Mardling, R. A., & Aarseth, S. J. 2001, MNRAS, 321, 398 Markwardt, C. B. 2009, in Astronomical Society of the Pacific Conference Series, Vol. 411, Astronomical Data Analysis Software and Systems XVIII, ed. D. A. Bohlender, D. Durand, & P. Dowler, 251 Mart´ı, J. G., & Beaug´e, C. 2012, A&A, 544, A97 Marzari, F., & Barbieri, M. 2007, A&A, 467, 347 Masuda, K. 2015, ApJ, 805, 28 Masuda, K., Uehara, S., & Kawahara, H. 2015, ApJL, 806, L37 Mathur, S., Huber, D., Batalha, N. M., et al. 2016, ArXiv Trotta, R. 2011, Journal of High Energy Physics, 6, 42 e-prints, arXiv:1609.04128 Feroz, F., & Hobson, M. P. 2008, MNRAS, 384, 449 Feroz, F., Hobson, M. P., & Bridges, M. 2009, MNRAS, 398, 1601 Mazeh, T., Holczer, T., & Shporer, A. 2015, ApJ, 800, 142 McQuillan, A., Mazeh, T., & Aigrain, S. 2013, ApJL, 775, L11 28 Masuda Figure C2. Joint posterior distribution from the dynamical analysis of Kepler-448b's TTVs and TDVs (Table 5 and Figure 3). Mills, S. M., & Fabrycky, D. C. 2017, AJ, 153, 45 Miralda-Escud´e, J. 2002, ApJ, 564, 1019 Morton, T. D., Bryson, S. T., Coughlin, J. L., et al. 2016, ApJ, 822, 86 Mustill, A. J., Davies, M. B., & Johansen, A. 2017, MNRAS, 468, 3000 Ortiz, M., Reffert, S., Trifonov, T., et al. 2016, A&A, 595, A55 Parviainen, H. 2015, MNRAS, 450, 3233 Perryman, M., Hartman, J., Bakos, G. ´A., & Lindegren, L. 2014, ApJ, 797, 14 Perryman, M. A. C., de Boer, K. S., Gilmore, G., et al. 2001, Petrovich, C. 2015, ApJ, 799, 27 Petrovich, C., & Tremaine, S. 2016, ApJ, 829, 132 Petrovich, C., Tremaine, S., & Rafikov, R. 2014, ApJ, 786, 101 Quinn, S. N., White, R. J., Latham, D. W., et al. 2014, ApJ, 787, 27 Ramm, D. J., Nelson, B. E., Endl, M., et al. 2016, MNRAS, 460, 3706 Rappaport, S., Deck, K., Levine, A., et al. 2013, ApJ, 768, 33 Rasio, F. A., & Ford, E. B. 1996, Science, 274, 954 Sanchis-Ojeda, R., Fabrycky, D. C., Winn, J. N., et al. 2012, A&A, 369, 339 Nature, 487, 449 Figure C3. Joint posterior distribution from the dynamical analysis of Kepler-693b's TTVs and TDVs (Table 6 and Figure 5). Santerne, A., H´ebrard, G., Deleuil, M., et al. 2014, A&A, 571, Triaud, A. H. M. J., Neveu-VanMalle, M., Lendl, M., et al. 2017, A37 Socrates, A., Katz, B., & Dong, S. 2012a, ArXiv e-prints, arXiv:1209.5724 MNRAS, arXiv:1612.04166 Van Eylen, V., Lindholm Nielsen, M., Hinrup, B., Tingley, B., & Kjeldsen, H. 2013, ApJL, 774, L19 Winn, J. N. 2011, in Exoplanets, ed. S. Seager (Tucson, AZ: Socrates, A., Katz, B., Dong, S., & Tremaine, S. 2012b, ApJ, University of Arizona Press), 55 -- 77 750, 106 Stamatellos, D., & Whitworth, A. P. 2009, MNRAS, 392, 413 Szab´o, G. M., P´al, A., Derekas, A., et al. 2012, MNRAS, 421, Wu, Y., & Lithwick, Y. 2011, ApJ, 735, 109 Wu, Y., & Murray, N. 2003, ApJ, 589, 605 Xue, Y., Masuda, K., & Suto, Y. 2017, ApJ, 835, 204 Xue, Y., & Suto, Y. 2016, ApJ, 820, 55 L122
1102.0511
1
1102
2011-02-02T17:32:35
Radial Velocity Survey of Low-Mass Companions to sdB Stars
[ "astro-ph.EP", "astro-ph.SR" ]
The origin of subdwarf B (sdB) stars is not fully understood yet since it requires high mass loss at the red giant stage. SdBs in close binary systems are formed via common envelope ejection, but the origin of apparently single sdB stars remains unclear. Substellar companions may be able to trigger common envelope ejection and help forming sdBs that appear to be single. Using a sample of high resolution spectra we aim at detecting small radial velocity (RV) shifts caused by such low mass (sub-)stellar companions. The RVs are measured with high accuracy using sharp metal lines. Our goal is to test the theoretical predictions and put constraints on the population of the lowest mass companions to sdB stars.
astro-ph.EP
astro-ph
RV survey of low-mass companions to sdB stars Lew Classen∗, Stephan Geier∗, Ulrich Heber∗ and Simon J. O'Toole† ∗Dr. Karl Remeis -- Observatory & ECAP, Astronomical Institute, Friedrich-Alexander University Erlangen-Nuremberg, Sternwartstr. 7, D-96049 Bamberg, Germany †Australian Astronomical Observatory, PO Box 296, Epping, NSW, 1710, Australia Abstract. The origin of subdwarf B (sdB) stars is not fully understood yet since it requires high mass loss at the red giant stage. SdBs in close binary systems are formed via common envelope ejection, but the origin of apparently single sdB stars remains unclear. Substellar companions may be able to trigger common envelope ejection and help forming sdBs that appear to be single. Using a sample of high resolution spectra we aim at detecting small radial velocity (RV) shifts caused by such low mass (sub-)stellar companions. The RVs are measured with high accuracy using sharp metal lines. Our goal is to test the theoretical predictions and put constraints on the population of the lowest mass companions to sdB stars. Keywords: subdwarf B stars, spectroscopic binaries, low-mass companions PACS: 97.80.Fk, 97.82.Fs, 97.20.Vs INTRODUCTION Subluminous stars are considered to be core helium-burning objects with a thin hy- drogen shell. In a Hertzsprung-Russell diagram these stars are situated on the Extreme Horizontal Branch (EHB). The evolution of these stars is far from understood since the formation of an sdB requires extreme mass loss at the tip of the RGB. The mechanism leading to this mass loss is still under debate [3, 4]. Formation scenarios can be roughly divided into single star and binary channels. About 50% of the known sdBs are found to be in close binaries with periods of few hours to several days consistent with the predictions of the common envelope (CE) ejection scenario [2]. Two main sequence stars evolve in a binary system. The more massive one becomes a red giant first and fills its Roche Lobe. Unstable mass transfer leads to the formation of a CE. The core of the red giant and the immersed secondary experience a loss of orbital energy which is deposited in the CE causing a spiral-in towards the center of mass. As soon as enough energy has been deposited in the envelope, it is ejected. The inwards migration stops and the stars end up in a close binary. This scenario is well suited to explain the known population of close sdB binaries. But how are the apparently single sdBs formed? Soker [10] suggested that substellar companions like planets or brown dwarfs may be able to trigger CE ejection as well. In order to survive the common envelope and avoid evaporation inside the envelope the secondary should have a minimum mass of more than 10 MJ. A similar scenario has been proposed for the formation undermassive single white dwarfs by Nelemans and Tauris [7]. Several substellar objects have recently been detected around hot subdwarfs in wide orbits, showing that planets can survive the red giant phase of their host star [9, 5, 8]. 0.4 0.3 0.2 0.1 0 x u l f d e z i l a m r o n He H e I Hd metal lines −0.1 3900 3950 4000 4050 l [Å] 4100 4150 4200 FIGURE 1. The plot displays a section of a FEROS spectrum of HD 205805 featuring exemplary sharp metal absorption lines. Recently, Geier et al. [1] discovered a candidate substellar companion in close orbit around the bright sdB HD 149382, which perfectly fits into the pattern predicted by Soker. Systems like that may be quite common and could have remained unnoticed up to now. The goal of our project is to put constraints on the population of the lowest mass companions to sdB stars by measuring accurate radial velocities from multi-epoch high resolution spectra. OBSERVATIONS AND DATA REDUCTION Up to now our survey consists of 23 bright sdB stars with visual magnitudes ranging from 10 to 14. We used high resolution spectra obtained with the FEROS instrument mounted at the ESO/MPG-2.2m telescope. The spectra have a resolution of 48 000. Spectra were reduced, calibrated and corrected for earth's orbital motion with the MI- DAS package using the FEROS reduction pipeline. All programme stars are single-lined objects without visible spectral features of a companion. Furthermore, those binaries with high radial velocity variability have been excluded because their unseen companions must be quite massive. Each star has been observed several times (from 2 to 15) on timescales ranging from one day to some years. As an example we display a section of the FEROS spectrum of HD 205805 in Fig 1. 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 B B B B B B B B B B B B B B B B B B B B B S S S S S S S S S S S S S S S S S S S S S T T T T T T T T T T T T T T T T T T T T T I I I I I I I I I I I I I I I I I I I I I F F F F F F F F F F F F F F F F F F F F F y y y y y y y y y y y y y y y y y y y y y b b b b b b b b b b b b b b b b b b b b b t t t t t t t t t t t t t t t t t t t t t i i i i i i i i i i i i i i i i i i i i i f f f f f f f f f f f f f f f f f f f f f a a a a a a a a a a a a a a a a a a a a a t t t t t t t t t t t t t t t t t t t t t a a a a a a a a a a a a a a a a a a a a a d d d d d d d d d d d d d d d d d d d d d 7 7 7 7 7 7 7 7 7 7 7 7 7 7 7 7 7 7 7 7 7 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 6 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 5 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 c c c c c c c c c c c c c c c c c c c c c + + + + + + + + + + + + + + + + + + + + + x x x x x x x x x x x x x x x x x x x x x u u u u u u u u u u u u u u u u u u u u u l l l l l l l l l l l l l l l l l l l l l f f f f f f f f f f f f f f f f f f f f f d d d d d d d d d d d d d d d d d d d d d e e e e e e e e e e e e e e e e e e e e e z z z z z z z z z z z z z z z z z z z z z i i i i i i i i i i i i i i i i i i i i i l l l l l l l l l l l l l l l l l l l l l a a a a a a a a a a a a a a a a a a a a a m m m m m m m m m m m m m m m m m m m m m r r r r r r r r r r r r r r r r r r r r r o o o o o o o o o o o o o o o o o o o o o n n n n n n n n n n n n n n n n n n n n n 1.5 1.5 1.5 1.4 1.4 1.4 1.3 1.3 1.3 1.2 1.2 1.2 1.1 1.1 1.1 1 1 1 0.9 0.9 0.9 FeIII 5833.938 Å FeIII 5833.938 Å FeIII 5833.938 Å FeIII 5833.938 Å FeIII 5833.938 Å FeIII 5833.938 Å FeIII 5833.938 Å FeIII 5833.938 Å FeIII 5833.938 Å FeIII 5833.938 Å FeIII 5833.938 Å FeIII 5833.938 Å FeIII 5833.938 Å FeIII 5833.938 Å FeIII 5833.938 Å FeIII 5833.938 Å FeIII 5833.938 Å FeIII 5833.938 Å FeIII 5833.938 Å FeIII 5833.938 Å FeIII 5833.938 Å SiIII 5739.734 Å SiIII 5739.734 Å SiIII 5739.734 Å SiIII 5739.734 Å SiIII 5739.734 Å SiIII 5739.734 Å SiIII 5739.734 Å SiIII 5739.734 Å SiIII 5739.734 Å SiIII 5739.734 Å SiIII 5739.734 Å SiIII 5739.734 Å SiIII 5739.734 Å SiIII 5739.734 Å SiIII 5739.734 Å SiIII 5739.734 Å SiIII 5739.734 Å SiIII 5739.734 Å SiIII 5739.734 Å SiIII 5739.734 Å SiIII 5739.734 Å NII 5710.766 Å NII 5710.766 Å NII 5710.766 Å NII 5710.766 Å NII 5710.766 Å NII 5710.766 Å NII 5710.766 Å NII 5710.766 Å NII 5710.766 Å NII 5710.766 Å NII 5710.766 Å NII 5710.766 Å NII 5710.766 Å NII 5710.766 Å NII 5710.766 Å NII 5710.766 Å NII 5710.766 Å NII 5710.766 Å NII 5710.766 Å NII 5710.766 Å NII 5710.766 Å NII 5679.558 Å NII 5679.558 Å NII 5679.558 Å NII 5679.558 Å NII 5679.558 Å NII 5679.558 Å NII 5679.558 Å NII 5679.558 Å NII 5679.558 Å NII 5679.558 Å NII 5679.558 Å NII 5679.558 Å NII 5679.558 Å NII 5679.558 Å NII 5679.558 Å NII 5679.558 Å NII 5679.558 Å NII 5679.558 Å NII 5679.558 Å NII 5679.558 Å NII 5679.558 Å FeIII 5156.111 Å FeIII 5156.111 Å FeIII 5156.111 Å FeIII 5156.111 Å FeIII 5156.111 Å FeIII 5156.111 Å FeIII 5156.111 Å FeIII 5156.111 Å FeIII 5156.111 Å FeIII 5156.111 Å FeIII 5156.111 Å FeIII 5156.111 Å FeIII 5156.111 Å FeIII 5156.111 Å FeIII 5156.111 Å FeIII 5156.111 Å FeIII 5156.111 Å FeIII 5156.111 Å FeIII 5156.111 Å FeIII 5156.111 Å FeIII 5156.111 Å CII 5145.165 Å CII 5145.165 Å CII 5145.165 Å CII 5145.165 Å CII 5145.165 Å CII 5145.165 Å CII 5145.165 Å CII 5145.165 Å CII 5145.165 Å CII 5145.165 Å CII 5145.165 Å CII 5145.165 Å CII 5145.165 Å CII 5145.165 Å CII 5145.165 Å CII 5145.165 Å CII 5145.165 Å CII 5145.165 Å CII 5145.165 Å CII 5145.165 Å CII 5145.165 Å NII 5005.150 Å NII 5005.150 Å NII 5005.150 Å NII 5005.150 Å NII 5005.150 Å NII 5005.150 Å NII 5005.150 Å NII 5005.150 Å NII 5005.150 Å NII 5005.150 Å NII 5005.150 Å NII 5005.150 Å NII 5005.150 Å NII 5005.150 Å NII 5005.150 Å NII 5005.150 Å NII 5005.150 Å NII 5005.150 Å NII 5005.150 Å NII 5005.150 Å NII 5005.150 Å OII 4649.135 Å OII 4649.135 Å OII 4649.135 Å OII 4649.135 Å OII 4649.135 Å OII 4649.135 Å OII 4649.135 Å OII 4649.135 Å OII 4649.135 Å OII 4649.135 Å OII 4649.135 Å OII 4649.135 Å OII 4649.135 Å OII 4649.135 Å OII 4649.135 Å OII 4649.135 Å OII 4649.135 Å OII 4649.135 Å OII 4649.135 Å OII 4649.135 Å OII 4649.135 Å NII 4630.539 Å NII 4630.539 Å NII 4630.539 Å NII 4630.539 Å NII 4630.539 Å NII 4630.539 Å NII 4630.539 Å NII 4630.539 Å NII 4630.539 Å NII 4630.539 Å NII 4630.539 Å NII 4630.539 Å NII 4630.539 Å NII 4630.539 Å NII 4630.539 Å NII 4630.539 Å NII 4630.539 Å NII 4630.539 Å NII 4630.539 Å NII 4630.539 Å NII 4630.539 Å SiIII 4574.757 Å SiIII 4574.757 Å SiIII 4574.757 Å SiIII 4574.757 Å SiIII 4574.757 Å SiIII 4574.757 Å SiIII 4574.757 Å SiIII 4574.757 Å SiIII 4574.757 Å SiIII 4574.757 Å SiIII 4574.757 Å SiIII 4574.757 Å SiIII 4574.757 Å SiIII 4574.757 Å SiIII 4574.757 Å SiIII 4574.757 Å SiIII 4574.757 Å SiIII 4574.757 Å SiIII 4574.757 Å SiIII 4574.757 Å SiIII 4574.757 Å SiIII 4567.840 Å SiIII 4567.840 Å SiIII 4567.840 Å SiIII 4567.840 Å SiIII 4567.840 Å SiIII 4567.840 Å SiIII 4567.840 Å SiIII 4567.840 Å SiIII 4567.840 Å SiIII 4567.840 Å SiIII 4567.840 Å SiIII 4567.840 Å SiIII 4567.840 Å SiIII 4567.840 Å SiIII 4567.840 Å SiIII 4567.840 Å SiIII 4567.840 Å SiIII 4567.840 Å SiIII 4567.840 Å SiIII 4567.840 Å SiIII 4567.840 Å SiIII 4552.622 Å SiIII 4552.622 Å SiIII 4552.622 Å SiIII 4552.622 Å SiIII 4552.622 Å SiIII 4552.622 Å SiIII 4552.622 Å SiIII 4552.622 Å SiIII 4552.622 Å SiIII 4552.622 Å SiIII 4552.622 Å SiIII 4552.622 Å SiIII 4552.622 Å SiIII 4552.622 Å SiIII 4552.622 Å SiIII 4552.622 Å SiIII 4552.622 Å SiIII 4552.622 Å SiIII 4552.622 Å SiIII 4552.622 Å SiIII 4552.622 Å NII 4530.410 Å NII 4530.410 Å NII 4530.410 Å NII 4530.410 Å NII 4530.410 Å NII 4530.410 Å NII 4530.410 Å NII 4530.410 Å NII 4530.410 Å NII 4530.410 Å NII 4530.410 Å NII 4530.410 Å NII 4530.410 Å NII 4530.410 Å NII 4530.410 Å NII 4530.410 Å NII 4530.410 Å NII 4530.410 Å NII 4530.410 Å NII 4530.410 Å NII 4530.410 Å FeIII 4419.596 Å FeIII 4419.596 Å FeIII 4419.596 Å FeIII 4419.596 Å FeIII 4419.596 Å FeIII 4419.596 Å FeIII 4419.596 Å FeIII 4419.596 Å FeIII 4419.596 Å FeIII 4419.596 Å FeIII 4419.596 Å FeIII 4419.596 Å FeIII 4419.596 Å FeIII 4419.596 Å FeIII 4419.596 Å FeIII 4419.596 Å FeIII 4419.596 Å FeIII 4419.596 Å FeIII 4419.596 Å FeIII 4419.596 Å FeIII 4419.596 Å SIII 4284.979 Å SIII 4284.979 Å SIII 4284.979 Å SIII 4284.979 Å SIII 4284.979 Å SIII 4284.979 Å SIII 4284.979 Å SIII 4284.979 Å SIII 4284.979 Å SIII 4284.979 Å SIII 4284.979 Å SIII 4284.979 Å SIII 4284.979 Å SIII 4284.979 Å SIII 4284.979 Å SIII 4284.979 Å SIII 4284.979 Å SIII 4284.979 Å SIII 4284.979 Å SIII 4284.979 Å SIII 4284.979 Å SIII 4253.589 Å SIII 4253.589 Å SIII 4253.589 Å SIII 4253.589 Å SIII 4253.589 Å SIII 4253.589 Å SIII 4253.589 Å SIII 4253.589 Å SIII 4253.589 Å SIII 4253.589 Å SIII 4253.589 Å SIII 4253.589 Å SIII 4253.589 Å SIII 4253.589 Å SIII 4253.589 Å SIII 4253.589 Å SIII 4253.589 Å SIII 4253.589 Å SIII 4253.589 Å SIII 4253.589 Å SIII 4253.589 Å FeIII 4137.764 Å FeIII 4137.764 Å FeIII 4137.764 Å FeIII 4137.764 Å FeIII 4137.764 Å FeIII 4137.764 Å FeIII 4137.764 Å FeIII 4137.764 Å FeIII 4137.764 Å FeIII 4137.764 Å FeIII 4137.764 Å FeIII 4137.764 Å FeIII 4137.764 Å FeIII 4137.764 Å FeIII 4137.764 Å FeIII 4137.764 Å FeIII 4137.764 Å FeIII 4137.764 Å FeIII 4137.764 Å FeIII 4137.764 Å FeIII 4137.764 Å NII 4043.532 Å NII 4043.532 Å NII 4043.532 Å NII 4043.532 Å NII 4043.532 Å NII 4043.532 Å NII 4043.532 Å NII 4043.532 Å NII 4043.532 Å NII 4043.532 Å NII 4043.532 Å NII 4043.532 Å NII 4043.532 Å NII 4043.532 Å NII 4043.532 Å NII 4043.532 Å NII 4043.532 Å NII 4043.532 Å NII 4043.532 Å NII 4043.532 Å NII 4043.532 Å NII 4041.310 Å NII 4041.310 Å NII 4041.310 Å NII 4041.310 Å NII 4041.310 Å NII 4041.310 Å NII 4041.310 Å NII 4041.310 Å NII 4041.310 Å NII 4041.310 Å NII 4041.310 Å NII 4041.310 Å NII 4041.310 Å NII 4041.310 Å NII 4041.310 Å NII 4041.310 Å NII 4041.310 Å NII 4041.310 Å NII 4041.310 Å NII 4041.310 Å NII 4041.310 Å NII 3994.997 Å NII 3994.997 Å NII 3994.997 Å NII 3994.997 Å NII 3994.997 Å NII 3994.997 Å NII 3994.997 Å NII 3994.997 Å NII 3994.997 Å NII 3994.997 Å NII 3994.997 Å NII 3994.997 Å NII 3994.997 Å NII 3994.997 Å NII 3994.997 Å NII 3994.997 Å NII 3994.997 Å NII 3994.997 Å NII 3994.997 Å NII 3994.997 Å NII 3994.997 Å CII 3920.681 Å CII 3920.681 Å CII 3920.681 Å CII 3920.681 Å CII 3920.681 Å CII 3920.681 Å CII 3920.681 Å CII 3920.681 Å CII 3920.681 Å CII 3920.681 Å CII 3920.681 Å CII 3920.681 Å CII 3920.681 Å CII 3920.681 Å CII 3920.681 Å CII 3920.681 Å CII 3920.681 Å CII 3920.681 Å CII 3920.681 Å CII 3920.681 Å CII 3920.681 Å OI 6363.780 Å OI 6363.780 Å OI 6363.780 Å OI 6300.304 Å OI 6300.304 Å OI 6300.304 Å OI 5577.340 Å OI 5577.340 Å OI 5577.340 Å −150 −150 −150 −150 −150 −150 −150 −150 −150 −150 −150 −150 −150 −150 −150 −150 −150 −150 −150 −150 −150 −100 −100 −100 −100 −100 −100 −100 −100 −100 −100 −100 −100 −100 −100 −100 −100 −100 −100 −100 −100 −100 −50 0 −50 −50 −50 −50 −50 −50 −50 −50 −50 −50 −50 −50 −50 −50 −50 −50 −50 −50 −50 −50 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 Vrad [kms−1] Vrad [kms−1] Vrad [kms−1] Vrad [kms−1] Vrad [kms−1] Vrad [kms−1] Vrad [kms−1] Vrad [kms−1] Vrad [kms−1] Vrad [kms−1] Vrad [kms−1] Vrad [kms−1] Vrad [kms−1] Vrad [kms−1] Vrad [kms−1] Vrad [kms−1] Vrad [kms−1] Vrad [kms−1] Vrad [kms−1] Vrad [kms−1] Vrad [kms−1] 50 50 50 50 50 50 50 50 50 50 50 50 50 50 50 50 50 50 50 50 50 0.8 0.8 0.8 −60 −50 −40 −30 −20 −10 0 10 20 30 −60 −50 −40 −30 −20 −10 0 10 20 30 −60 −50 −40 −30 −20 −10 0 10 20 30 Vrad [kms−1] Vrad [kms−1] Vrad [kms−1] FIGURE 2. Spectral features of HD 205805 fitted with FITSB2. The left hand panel displays metal absorption lines suitable for RV measurements, in the right hand panel exemplary night-sky emission lines used to check the wavelength calibration are shown. The latter are shifted in RV due to heliocentric correction applied to the spectra. MEASUREMENT Subdwarf B stars are hot (Teff = 20 000 − 35 000 K) and their atmospheres radiative. The dominant features in their spectra are the hydrogen Balmer and several helium lines. But the spectra also show up to ≃ 50 weak metal absorption lines which are very sharp. These features are best suited to measure radial velocities. The atmospheres of sdB stars are considered to be stable. The only known effect other than a close companion that may lead to RV variability are pulsations. Pulsating sdB stars do indeed exist with periods between 90 s and 3 h and RV amplitudes of few km s−1 at most [4]. For our analysis we chose a set of sharp, unblended metal lines situated between 3600 and 6600 Å (see Fig. 2, left hand panel). Accurate rest wavelengths were taken from the NIST database. Depending on the atmospheric parameters of the star and the quality of the data a subset was used. The RV measurement consisted of two steps. First, we fitted Gaussian and Lorentzian profiles to each line separately using the SPAS routine (Hirsch priv. comm.). The re- sulting mean RVs had standard deviations ranging from 0.8 to 2.7 km s−1. Second, we D D D D D D D D D D D D D D D D D D D D D D D D used the FITSB2 routine [6] to perform a simultaneous fit of all suitable lines. Uncer- tainties were calculated using a bootstrapping algorithm and found to range from 0.1 to 0.5 km s−1, while both methods result in consistent mean RVs. The scatter of individual RV values measured with SPAS around the mean value was reviewed for possible systematic effects like correlations with certain elements. No such effects could be found. To check the wavelength calibration for systematic errors we used telluric features as well as night-sky emission lines (see Fig. 2, right hand panel). Having their origin on earth these lines should have zero RV. The FEROS instrument turned out to be very stable. Usually corrections of less than 0.5 km s−1 had to be applied. In order to derive the degree of precision we can achieve with our method, we gen- erated synthetic model spectra with realistic atmospheric parameters including Poisson noise with S/N = 1000 and determined the RV in the way described above. The un- certainties turned out to range from 0.1 to 0.4 km s−1. The accuracy is limited by the resolution of the spectrograph, the number of features and the intrinsic thermal broad- ening of the lines. RESULTS 0 −5 −10 ] 1 − s m k [ V R −15 −20 −25 3800 4000 4200 4400 4600 5000 5200 5400 5600 5800 6000 4800 l [Å] FIGURE 3. The diagram shows the result of RV determination for two spectra of EC 20106−5248 taken 6 months apart. Each measuring point represents a metal absorption line. The horizontal lines marking mean RV values of the spectra indicate a significant shift in radial velocity. TABLE 1. Summary of RV results for our sample obtained using two different algorithms. RVs for each line were measured independently and averaged with SPAS (left hand column) whereas FITSB2 uses a simultaneous fit (right hand column). SPAS FITSB2 star synthetic EC 00042−2737 CD-3515910 Feige 38 EC 01120−5259 PG 1432+004 EC 02542−3019 PG 1505+074 EC 20229−3716 EC 21043−4017 EC 03591−3232 HD 4539 HD 205805 PHL 334 EC 11349−2753 PG 0342+026 EC 14345−1729 HE 0151−3919 PG 2151+100 EC 15103−1557 PHL 44 JL 36 EC 13506−3137 EC 20106−5248 D Vrad max [km s−1 ] 0.06 0.08 0.17 0.21 0.23 0.32 0.35 0.40 0.41 0.50 0.53 0.64 0.71 0.84 1.04 1.06 1.10 1.29 2.08 2.60 3.16 5.38 5.64 12.43 standard deviation 0.28 1.21 1.11 0.93 1.23 1.11 2.68 0.64 0.79 1.30 1.19 0.91 1.00 1.08 0.95 0.94 1.30 1.65 1.36 1.58 1.76 1.96 1.52 1.61 D Vrad max [km s−1 ] mean error 0.00 0.10 0.62 1.02 0.80 3.20 12.65 0.09 0.13 0.52 0.49 0.22 Four stars were found to be significantly (another two marginally) RV variable. The shift in RV ranges from 0.5 to 12.7 km s−1 (see Tab. 1). As an example we display two RV measurements for EC 20106−5248 in Fig. 3. Follow-up photometry and high resolution spectroscopy are necessary to exclude pulsational variability and derive the orbital parameters of these binaries. Constraints can then be put on the companion masses. Even if there are no low mass companions we would expect to find a certain fraction of systems with small RV shifts caused by more massive secondaries seen at low inclination. A larger sample is needed to exclude this case. No RV variations were found in the rest of the sample. Nevertheless, the errors allow to derive upper limits for Mcomp sini of undetected substellar companions in close orbits from the binary mass function (Eq. 1). f (M) = M3 comp sin3 i (Mcomp + MsdB)2 = PK3 2p G (1) Soker [10] suggested that these objects should be more massive than 10 MJ and should have orbital periods shorter than 10 d. Adopting this maximum period and assuming a ] r e t i p u J M [ i n s i p m o c M m u m x a m i 20 18 16 14 12 10 8 6 4 2 0 SPAS FITSB2 6 1 7 3 − 9 2 2 0 2 C E 4 7 0 + 5 0 5 1 G P t h n y s 2.5 2.0 1.5 1.0 0.5 0 1 9 5 1 5 3 − D C 9 3 5 4 D H 7 3 7 2 − 2 4 0 0 0 C E 4 0 0 + 2 3 4 1 G P 9 5 2 5 − 0 2 1 1 0 C E 2 3 2 3 − 1 9 5 3 0 C E 3 5 7 2 − 9 4 3 1 1 C E 6 2 0 + 2 4 3 0 G P 5 0 8 5 0 2 D H 7 1 0 4 − 3 4 0 1 2 C E 9 2 7 1 − 5 4 3 4 1 C E 9 1 9 3 − 1 5 1 0 E H 9 1 0 3 − 2 4 5 2 0 C E 7 5 5 1 − 3 0 1 5 1 C E 4 3 3 L H P 8 3 e g e F i ] 1 − s m k [ K FIGURE 4. The histogram shows upper mass limits for unseen companions in close orbits around sdBs, where no significant RV shifts could be detected (left hand axis). These are derived from upper limits for the RV semi-amplitudes (right hand axis). canonical mass of 0.47 M⊙ for the sdB, the errors of the RV measurements can be used as upper limits for the RV semi-amplitudes (K). As can be seen in Fig. 4, tight constraints can be put on possible substellar compan- ions, if the measurement error is smaller than ≃ 1.0 km s−1. Given data of perfect quality, close-in planets with masses as low as 0.3 MJ can be firmly excluded if our assumption about the orbital period is correct. In conclusion ≃ 26% of the apparently single stars in our sample show small RV variations. The true fraction is expected to be higher, because the accuracy is limited by the S/N in most cases. Companions with masses higher than 10 MJ in close orbits (P < 10 d) can be excluded for 61% of our sample. One possible explanation for the fractions found is that the assumed CE scenario is not the only channel for sdB formation. However, even with CE ejection being the only channel, the companion may be destroyed during the CE phase either by evaporation or a merger with the stellar core leaving behind a single sdB. Better statistics and reliable predictions from theoretical calculations are required to solve this problem. ACKNOWLEDGMENTS S.G. is supported by DFG through grant He 1353/49-1. REFERENCES 1. S. Geier, H. Edelmann, U. Heber, and L. Morales-Rueda, ApJ 702, L96 -- L99 (2009). 2. Z. Han, P. Podsiadlowski, P. F. L. Maxted, T. R. Marsh, and N. Ivanova, MNRAS 336, 449 -- 466 (2002). 3. U. Heber, A&A 155, 33 -- 45 (1986). 4. U. Heber, ARA&A 47, 211 -- 251 (2009). 5. J. W. Lee, S. Kim, C. Kim, R. H. Koch, C. Lee, H. Kim, and J. Park, AJ 137, 3181 -- 3190 (2009). 6. R. Napiwotzki, L. Yungelson, G. Nelemans, T. R. Marsh, B. Leibundgut, R. Renzini, D. Homeier, D. Koester, S. Moehler, N. Christlieb, D. Reimers, H. Drechsel, U. Heber, C. Karl, and E. Pauli, "Dou- ble degenerates and progenitors of supernovae type Ia," in Spectroscopically and Spatially Resolving the Components of the Close Binary Stars, edited by R. W. Hilditch, H. Hensberge, & K. Pavlovski, 2004, vol. 318 of Astronomical Society of the Pacific Conference Series, pp. 402 -- 410. 7. G. Nelemans, and T. M. Tauris, A&A 335, L85 -- L88 (1998). 8. S. Qian, L. Zhu, S. Zola, W. Liao, L. Liu, L. Li, M. Winiarski, E. Kuligowska, and J. M. Kreiner, ApJ 695, L163 -- L165 (2009). 9. R. Silvotti, S. Schuh, R. Janulis, J. Solheim, S. Bernabei, R. Østensen, T. D. Oswalt, I. Bruni, R. Gualandi, A. Bonanno, G. Vauclair, M. Reed, C. Chen, E. Leibowitz, M. Paparo, A. Baran, S. Charpinet, N. Dolez, S. Kawaler, D. Kurtz, P. Moskalik, R. Riddle, and S. Zola, Nature 449, 189 -- 191 (2007). 10. N. Soker, AJ 116, 1308 -- 1313 (1998).
1808.00052
1
1808
2018-07-31T20:00:13
Understanding WASP-12b
[ "astro-ph.EP" ]
The orbital period of the hot Jupiter WASP-12b is apparently changing. We study whether this reflects orbital decay due to tidal dissipation in the star, or apsidal precession of a slightly eccentric orbit. In the latter case, a third body or other perturbation would be needed to sustain the eccentricity against tidal dissipation in the planet itself. We have analyzed several such perturbative scenarios, but none is satisfactory. Most likely therefore, the orbit really is decaying. If this is due to a dynamical tide, then WASP-12 should be a subgiant without a convective core as Weinberg et al. (2017) have suggested. We have modeled the star with the MESA code. While no model fits all of the observational constraints, including the luminosity implied by the GAIA DR2 distance, main-sequence models are less discrepant than subgiant ones.
astro-ph.EP
astro-ph
MNRAS 000, 1 -- 12 (2017) Preprint 2 August 2018 Compiled using MNRAS LATEX style file v3.0 Understanding WASP 12b Avery Bailey1(cid:63) and Jeremy Goodman1 1Department of Astrophysical Sciences, Princeton University, Princeton, NJ 08540, USA Accepted XXX. Received YYY; in original form ZZZ ABSTRACT The orbital period of the hot Jupiter WASP-12b is apparently changing. We study whether this reflects orbital decay due to tidal dissipation in the star, or apsidal preces- sion of a slightly eccentric orbit. In the latter case, a third body or other perturbation would be needed to sustain the eccentricity against tidal dissipation in the planet itself. We have analyzed several such perturbative scenarios, but none is satisfactory. Most likely therefore, the orbit really is decaying. If this is due to a dynamical tide, then WASP-12 should be a subgiant without a convective core as Weinberg et al. (2017) have suggested. We have modeled the star with the mesa code. While no model fits all of the observational constraints, including the luminosity implied by the GAIA DR2 distance, main-sequence models are less discrepant than subgiant ones. Key words: planet -- star interactions -- stars: individual: WASP-12 -- planets and satellites: individual: WASP-12b -- planets and satellites: dynamical evolution and stability 1 INTRODUCTION Much circumstantial evidence indicates that tidal dissipa- tion sculpts the orbits of short-period binary stars and ex- oplanets. First-principles tidal theories often have difficulty explaining the observations quantitatively, however. For ex- ample, among low-mass main-sequence binaries, the period below which orbits circularize appears to increase with sys- tem age up to periods ∼ 20 d, whereas standard dissipation mechanisms become ineffective beyond ∼ 10 d (Zahn 2013). Transiting exoplanets offer the prospect of testing tidal dissipation in real time. Massive exoplanets with very short periods are expected to exhibit orbital decay due to tidal dissipation in their host stars, whose rotation is usually sub-synchronous, on timescales short compared to the star's main-sequence lifetime (Levrard et al. 2009). (This should not occur for stellar binaries because of the much greater an- gular momentum in the orbit, only a small fraction of which is needed to bring the stars into synchronous rotation.) In favorable cases where the inspiral time is (cid:46) 107 yr, transit timing with sub-minute accuracy may be expected to detect the period change after a decade or so. Currently the most promising tentative detection has been made for WASP-12b, a planet with mass mb ≈ 1.5 MJ in a 1.0914 d orbit around a main-sequence F star (Hebb et al. 2009). Highly statistically significant departures from a linear transit ephemeris have been measured by Maciejewski et al. (2016) and recently confirmed by Patra et al. (2017). (cid:63) E-mail: [email protected] © 2017 The Authors According to the latter authors, the measured rate of change of orbital period is (cid:219)P = −29 ± 3 ms yr−1, and P/ (cid:219)P = 3.2 Myr. Three hypotheses for the orbital period change have been discussed. One is orbital decay. A second is precession of the periapse of a slightly eccentric orbit with a period ∼ 10 yr (Maciejewski et al. 2016). The required eccentricity is on the order of 10−3, well below the limit e < 0.05 set by Husnoo et al. (2012). Patra et al. (2017) find that this explanation is disfavored by times of planetary occultation (secondary eclipse) as measured with Spitzer: an eccentric orbit would tend to displace the times of primary and sec- ondary eclipses in opposite directions, whereas the data seem to prefer an advance of both. Furthermore, it seems unlikely that even such a small eccentricity could have survived tidal dissipation in the planet. Nevertheless, Patra et al. (2017) conclude that apsidal precession cannot yet be definitively ruled out on the basis of the timing data. The third possible explanation for (cid:219)P is acceleration by a companion. In fact WASP-12 is accompanied by a pair of M stars at projected separation ≈ 1(cid:48)(cid:48), (Bechter et al. 2014). Given that the estimated mass of this pair is ≈ 0.75 M(cid:12) and the distance to WASP-12 is 432.5±6.1 pc (Gaia Collaboration et al. 2018), the maximum line of sight acceleration is ≈ 0.33 m s−1 yr−1, corresponding to (cid:219)P < 0.1 ms yr−1, far smaller than the observed value. More to the point -- because there might be unseen massive planets closer in -- Knutson et al. (2014) have used their radial-velocity data to place a limit (cid:46) 4 m s−1 yr−1 on this acceleration, and this is still almost an order of magnitude too small to explain (cid:219)P. In the absence of a plausible fourth hypothesis, orbital 8 1 0 2 l u J 1 3 . ] P E h p - o r t s a [ 1 v 2 5 0 0 0 . 8 0 8 1 : v i X r a 2 A. Bailey and J. Goodman decay would therefore seem to be the best explanation for the observed departures from a linear ephemeris. There are, however, reasons for doubt. If the orbital decay timescale is in fact only ∼ 3 Myr, whereas the main-sequence lifetime of the host star is (cid:38) 1 Gyr (see §2), we must be viewing the system at a special time. On the other hand, WASP-12 is perhaps the best current candidate for measurable orbital decay out of hundreds of hot Jupiters, so perhaps such a "coincidence" should be less surprising. A potential concern is the small measured rotation: v sin i < 2.2 km s−1 (Hebb et al. 2009), v sin i < 5.1 km s−1 (Fossati et al. 2010b), or v sin i = 3.4 ± 0.9 km s−1 (Torres et al. 2012). If the planetary orbit has donated much of its original angular momentum to the star, one might expect the star to have a larger v sin i: the converse argument has been used by Penev et al. (2016) to suggest orbital decay in the HATS-18 system. In §3 and §4.2 however, we demonstrate that this expectation is in- correct and that tidal mechanisms are insufficient to bring WASP-12 to full synchronous rotation. Instead, tidal mecha- nisms should spin-up only a small core region of the star -- the observational effect of which we explore. The orbital decay explanation has been previously in- vestigated by Weinberg et al. (2017) who offer the novel sug- gestion that WASP-12 is a subgiant star. But because this particular system holds a unique and valuable place within the context of tidal theories and planet-star interaction, we felt it necessary to investigate this system further. We make a more thorough examination of stellar models before in- dependently coming to similar interpretations as Weinberg et al. (2017). Our analysis also benefits from the most recent luminosity estimates for WASP-12 (see §2.3) and while the results are inconclusive, this new luminosity favors a higher mass main-sequence model. Bearing in mind this preference for a main-sequence model, we present a comprehensive in- vestigation of alternative explanations for the observed pe- riod change in §5. 2 STELLAR MODELS The tidal dissipation mechanisms discussed here are sensi- tive to the internal structure of the star, particularly the existence and extent of convection zones. Therefore, we be- gin by selecting a fiducial model for the WASP-12 host. 2.1 Properties of the WASP-12 star Table 1 summarizes properties of WASP-12 as indepen- dently determined by the studies cited. The effective temper- ature and metallicity are in principle directly determinable by comparison of spectra with atmospheric models, while the mean stellar density ¯ρ ≡ 3M∗/4πR3∗ follows from the or- bital period and the fractional width R∗/a of the planetary transit. We have chosen not to use spectroscopic determi- nations of surface gravity, as some of these studies regard log g as problematic unless constrained by the mean den- sity. The adopted values on the last line were obtained as a straightforward weighted average of the values shown: −2 σ = Xk σ2 k , k 1 σ2 k (1) −2 X = σ  k the original errors {σk} being symmetrized where necessary, e.g. 6300+200−100 → 6300 ± 150. The adopted error σ is opti- mistic, especially for Teff and [Fe/H], since the original errors are probably dominated by systematics of the atmospheric models. 2.2 Interior models Models for WASP-12 were constructed using the 10108 re- lease of the publicly available 1-D stellar evolution code mesa (Paxton et al. 2011, 2013, 2015, 2018). This version of mesa includes an improved prescription for determining radiative-convective boundaries, the locations of which can sensitively alter the strength of tidal effects. We arrived at a fiducial model for WASP-12 after searching the param- eter space over mass M∗ and initial metallicity Zinit with bounds 1.15 M(cid:12) < M∗ < 1.4 M(cid:12) and 0.01 < Zinit < 0.033. An unweighted sum of χ2 statistics of the adopted properties listed in Table 1 was chosen as the goodness-of-fit statistic to be minimized over the searched parameter space. Though we conducted searches including all three observables in the goodness-of-fit statistic χ2 ( ¯ρ, Teff, [Fe/H]), here we focus on the results of searches for the statistic χ2 ( ¯ρ, Teff) which omit [Fe/H]. Both statistics give similar results, but the latter lends itself to analyzing the observables in an individual sense instead of a combined one. For our calculation of the model [Fe/H] ≡ log10(Zsurf/Xsurf)−log10(Z(cid:12)/X(cid:12)), we adopted the Asplund et al. (2009) value, Z(cid:12)/X(cid:12) = 0.0181. For each combination of M∗ and Zinit, a stellar model was evolved from pre-main sequence with our goodness-of-fit statistic evalu- ated at each timestep until the statistic moved far enough from a local minimum to trigger the stopping conditions for that evolutionary run. For all parameter space searches we adopted the physics of Choi et al. (2016), but tested two mixing length parame- ter values of α = 1.9 and α = 2.3. The results of a series of evolutionary runs according to both several grid and simplex searches for χ2 ( ¯ρ, Teff) are displayed in Fig. 1 for α = 1.9 and Fig. 2 for α = 2.3. We highlight three models in particular between Figs. 1 & 2 and provide additional details in Ta- ble 2. The first model, model A, is representative of a class of models that are on the main-sequence, plus are able to adequately fit the observed ¯ρ, Teff, and [Fe/H]. Most signif- icantly, models of this type have a small convective core. If WASP-12 had no convective core, gravity waves excited at the outer convective -- radiative boundary might deposit their angular momentum by breaking non-linearly at the inner turning point where the Brunt -- Vaisala N equals the tidal frequency ω (Goodman & Dickson 1998; Terquem et al. 1998; Barker & Ogilvie 2010; Weinberg et al. 2017). The con- vective core in these models removes the possibility of this dissipation mechanism by moving the turning point outward to a region where the gravity waves lack the amplitude to break. As we show in §3 and §4.1, models with the structure of model A are unable to explain the observed tidal decay of WASP-12b. This motivated the search for an additional class of models that lack a convective core. At lower masses, the convective core of models that still fit the observed ¯ρ and Teff shrinks. This continues until, as is displayed in Figs. 1 & 2, the convective core disappears entirely around a mass of ≈ 1.2 M(cid:12) where the well-fitting models become subgiants. MNRAS 000, 1 -- 12 (2017) Table 1. Observed and adopted properties of WASP-12 Understanding WASP 12b 3 Teff K 6300+200−100 6250 ± 100 6118 ± 64 6313 ± 52 -- -- [Fe/H] dex 0.30.05 0.15 0.32 ± 0.12 0.07 ± 0.07 0.21 ± 0.04 -- -- 6241 ± 36 0.198 ± 0.032 ¯ρ ¯ρ(cid:12) 0.35 ± 0.03 Reference Hebb et al. (2009) -- -- Fossati et al. (2010b) Torres et al. (2012) Southworth (2012) Mortier et al. (2013) Maciejewski et al. (2013) -- 0.325 ± 0.016 0.315 ± 0.007 0.3181 ± 0.0063 Adopted We focus on two representative subgiant models that we re- fer to as model B and C, differentiated by having α = 1.9 and α = 2.3 respectively. Unfortunately, each of these models suffers from inconsistencies: model B having too low surface metallicity to match observations and model C having an unrealistically high α. While our fiducial value for the error on [Fe/H] is un- doubtably optimistic, past analyses have all concurred that WASP-12 has supersolar surface abundance, whereas our model B is near solar or slightly subsolar. As the surface metallicity is dependent upon the prescription for elemen- tal diffusion and stellar rotation, we also tested an alter- native prescription more tailored to WASP-12 than Choi et al. (2016), assuming a rotation rate of ≈ 10 km s−1. This rotation rate was informed by measurements of the Rossiter -- McLaughlin effect in this system that indicate a strong spin -- orbit misalignment of 59+15◦ −20 and that v sin i∗ = 1.6+0.8−0.4 km s−1 (Albrecht et al. 2012). Corresponding changes to the surface metallicity were minimal and it seems unlikely that any rotation or diffusion mechanism would be able to signif- icantly enhance the surface metallicity above Zinit to repro- duce the observed [Fe/H]. Though model C fits well for all three observables, it assumes α = 2.3, in what is probably an unrealistically high choice of α. As Procyon is a spectral neighbor to WASP- 12 and is particularly well-constrained, it provides reason- able calibration for α. Assuming a mass of 1.478 ± 0.012M(cid:12) (Bond et al. 2015), setting initial [Fe/H] equal to the ob- served [Fe/H] = −0.05 ± 0.03 (Allende Prieto et al. 2002), models with α ≥ 2.2 can not simultaneously reproduce the observed Teff = 6516 ± 87 K (Aufdenberg et al. 2005), lumi- nosity log10(L/L(cid:12)) = 0.84 ± 0.018 (Jerzykiewicz & Molenda- Zakowicz 2000), and ¯ρ = 0.1725 ± 0.0007 ρ(cid:12) (Bedding et al. 2010). Instead, the best-fitting models occur in the range 1.8 ≤ α ≤ 2.1. Solar calibrations corroborate this, suggesting α = 1.93 (van Saders & Pinsonneault 2012), α = 1.82 (Choi et al. 2016), etc. While we certainly have not exhausted the parameter space of possible subgiant models, and there are enough tun- able parameters in stellar modeling that it may be possible to construct a subgiant model that fits the observables, with standard assumptions it is difficult to do so. On the other hand, main-sequence models that fit the observables (ex- cept the luminosity -- see below) are generic and easy to find, which is likely why previous studies estimate the mass of WASP-12 to be near 1.4M(cid:12) (Collins et al. 2017; Southworth 2012). Though the convective core in these models inhibits MNRAS 000, 1 -- 12 (2017) Figure 1. The tested parameter space over initial mass and metallicity colored according to the goodness-of-fit statistic un- der a cubic interpolation. Lines are plotted for the adopted [Fe/H]=0.198 (blue) and the solar value (red). Solid black con- tours for luminosity are placed at the level: log10(L∗/L(cid:12)) = 0.65 with ±10% error bars. Models with masses lower than the black dashed line have a radiative core whereas higher masses have a small convective core. These models assume mixing length pa- rameter α = 1.9. The location of models A & B are marked as cyan points tidal decay via breaking of gravity waves, the presence of a convective envelope allows for damping of the equilibrium tide (§3) and dynamical tide (§4.1) by turbulent viscosity (Zahn 1977). In these later sections however, we adopt model A as a fiducial model to show that both the equilibrium tide and dynamical tide in main-sequence models are unable to explain WASP-12b's decay. 2.3 WASP-12 luminosity In addition to using ¯ρ and Teff to constrain stellar models, we made some investigation of whether our models A,B,C could be constrained via luminosity measurements. The most re- cent parallax measurements made by the GAIA mission (Gaia Collaboration et al. 2016, 2018) place WASP-12 at a distance of 432.5 ± 6.1 pc, significantly further than past determinations. Adapting the work of Stassun et al. (2017), 1.151.201.251.301.35M∗(M(cid:12))0.0150.0200.0250.030Zinit10−1100101102χ2(cid:0)¯ρ,Teff(cid:1) 4 A. Bailey and J. Goodman Table 2. Fiducial WASP-12 Models Name α Zinit A B C 1.9 1.9 2.3 0.0223 0.0162 0.0234 M∗ M(cid:12) 1.34 1.20 1.24 Age Gyr 2.72 4.24 4.51 log10(L∗/L(cid:12)) 0.55 0.52 0.53 Teff K 6250 6242 6245 [Fe/H] dex 0.20 -0.03 0.20 ¯ρ ¯ρ(cid:12) 0.3182 0.3185 0.3181 χ2 ( ¯ρ, Teff) χ2 ([Fe/H]) 0.064 0.006 0.017 0.031 49.06 0.006 et al. (2018)1 predicts E(B − V) = 0.070.02 0.03 mag at 440 pc in the direction of WASP-12, which would correspond to AV ≈ 0.21 ± 0.09 mag for a normal extinction curve. Query- ing the Gaia DR2 catalog for stars within one degree of WASP-12, parallaxes ≥ 2.28 m.a.s., and Teff > 5500 K yields 101 results, of which 78 have AG values listed. There is no clear trend with distance, but the median AG for the more distant half of this sample is 0.23 mag. These are slightly lower than the Stassun et al. (2017) estimate, but consistent within the uncertainties. So it seems that WASP-12 is at least ∼ 10 - 30% more luminous than any of the models in Table 2. Considering these three independently determined ob- servables ( ¯ρ, Teff, L) are incompatible with one another, we also ran a chi-square model search including the luminosity, with 10% errors on L∗. As one can infer from inspection of the luminosity contours in Figs. 1 & 2, including L∗ in the search moves the track of well-fitting models towards lower metallicity such that it lies between the low χ2( ¯ρ, Teff) track and the log10(L∗/L(cid:12)) = 0.65 contour. As the three observ- ables are incompatible, χ2( ¯ρ, Teff, L∗) becomes significantly more nonzero, with minimum χ2 ≈ 3 -- 5, depending on α and assuming 10% error bars on the luminosity. Ultimately, it seems that because lines of constant ¯ρ, Teff lie nearly par- allel to lines of constant L∗ in the Zinit − M∗ plane, that luminosity measurements offer little guidance in choosing between models of type A, B, or C (at least in the mass range 1.15 M(cid:12) < M∗ < 1.4 M(cid:12)). 3 EQUILIBRIUM TIDE The adiabatic equilibrium tide describes the hydrostatic tidal response of the host star to a perturbing body in the ab- sence of dissipation. In this hydrostatic limit where the tidal frequency ω goes to zero, the functional relationship between density ρ, pressure p, and potential Φ is preserved. As a re- sult, density and pressure are constant along equipotentials and have the same value on a given equipotential as they would on the same equipotential absent the tide. If one ne- glects composition gradients, entropy S is a function of ρ and P only and would consequently follow the equipotentials. In regions with a nonzero entropy gradient, i.e. radiative zones, the adiabatic condition would require fluid elements r ∝ Φ1, the subscript 1 eq also stay tied to equipotentials (ξ indicating an Eulerian perturbation). The equilibrium tide is also incompressible, ∇∇∇· ξξξeq = 0. In stably stratified regions the radial fluid displacement is explicitly described by the 1 which can be queried at argonaut.skymaps.info MNRAS 000, 1 -- 12 (2017) Figure 2. The tested parameter space for models with mixing length parameter α = 2.3. Plotted according to the caption in Fig. 1 with the location of model C marked as a cyan point who estimate the extinction to WASP-12 at AV = 0.29 mag., to this updated distance measurement gives a luminosity of log10 (L∗/L(cid:12)) = 0.65. Taking L∗ together with our se- lected values of Teff and ¯ρ uniquely determines the mass at M∗ ≈ 1.9 M(cid:12). Such a high mass would seem to favor the higher mass main-sequence models for WASP-12 but evolu- tionary runs at 1.9 M(cid:12) fail to simultaneously fit these three observables. Figs. 1 & 2 similarly suggest that any model sig- nificantly larger than 1.4 M(cid:12) and having the requisite Teff, ¯ρ would require an unrealistically high metallicity. For exam- ple, even the best-fitting highest metallicity 1.9 M(cid:12) model tested ([Fe/H]≈ 0.4) had a value χ2 ( ¯ρ, Teff, L∗) > 100. The incompatibility of these three observables is also visible in Figs. 1 & 2 as lines of constant L∗ lie parallel to the track of constant ¯ρ, Teff. This tension between measured observ- ables is alleviated as one goes to either higher Teff or lower luminosity. While GAIA DR2 lists a very precise parallax for WASP-12, 2.3122 ± 0.0325 m.a.s., the extinction (AG) is not reported. Without correction for extinction, the reported lu- minosity is 3.435±0.075 L(cid:12), which is entirely compatible with the models in Table 2. One might therefore worry that Stas- sun et al. (2017) have overestimated the extinction or the flux -- the former perhaps because some of the photometric data they used were published before it was recognized that the star has two M-dwarf companions within 1(cid:48)(cid:48) (Bergfors et al. 2013). But for comparison, the dust map of Green 1.151.201.251.301.35M∗(M(cid:12))0.0150.0200.0250.030Zinit10−1100101102χ2(cid:0)¯ρ,Teff(cid:1) Þ 1 equation, eq ξ r = Φ1 dΦ/dr . (2) With the addition of composition gradients, though the above entropy argument no longer holds, a similar result can be derived. Namely, the fluid displacements are still de- scribed by eq. (2) and are still incompressible where the square Brunt-Vaisala frequency N2 (cid:44) 0 . In convective re- gions where the entropy gradient vanishes, and N2 = 0, fluid displacements may not necessarily follow the above eq. (2) but one defines the equilibrium tide such that eq. (2) is sat- isfied. In convective regions of a star, turbulent viscous forces facilitate the cascading of bulk kinetic energy to smaller scales where it is dissipated. The form of this dissipative system and it's action on the equilibrium tide in stars with a convective region was developed by Zahn (1966) and dis- sipates the energy on a timescale (Remus et al. 2012), R∗ M∗ = 4π R+/R∗ 6264 35 ρνt x8dx 1 (3) tdiss where x ≡ r/R∗ is the fractional radius, R+ is the radius of the outermost radiative-convective boundary, and νt is the convective viscosity. Here we have restricted the limits of in- tegration to extend only over the convective envelope as the equivalent contribution due to the convective core is heav- ily suppressed by the x8 dependence within the integrand. Eq. (3) operates under the assumption of a thin convective envelope. Stated more precisely: (i) the mass of the convective region is negligible (M∗ ≈ M+), (iii) the stellar invariant U is small compared to unity (ii) the self-interacting perturbation to the potential caused by equilibrium tide displacements Φ1,∗ is small com- pared to the perturbation to the potential caused by the star's external companion Φ1,b (Φ1,∗ + Φ1,b ≈ Φ1,b), (U ≡ d ln M/d ln R (cid:28) 1). Our fiducial WASP-12 models satisfy the above criteria for a thin convective envelope with the convective envelope con- taining < 0.2 per cent of the mass of the star, Φ1,∗ < 0.03Φ1,b, and the stellar invariant U < 0.025. In calculating the dissipation rate associated with the equilibrium tide, we assume a viscosity of the form, (cid:113) νt = lc vc 1 + (τc/Ptide)2 (4) where lc is the mixing length, vc is the r.m.s. vertical con- vective velocity, τc ≡ 2lc/vc is twice the local convective turnover time and Ptide = 2π/ω is the tidal period. This form of the viscosity is a heuristic that reproduces the tur- bulent viscosity formalism of Zahn (1966) in the limit that Ptide (cid:29) τc and Ptide (cid:28) τc. The suppression of viscosity that arises in Zahn's formalism for Ptide (cid:28) τc comes from the fact that for large tidal frequencies, eddies are unable to travel a full mixing length. One then supposes that the mean free path of such an eddy should be replaced by the distance an eddy travels in something like half a tidal period result- ing in a suppression by a factor of Ptide/τc. Others such as Goldreich & Nicholson (1977) have argued that eddies MNRAS 000, 1 -- 12 (2017) Understanding WASP 12b 5 with turnover times greater than the tidal period do not 'exchange momentum with the mean flow on this time scale [the tidal period]' and therefore are not necessary in evaluat- ing the diffusivity. The result is a suppression of the viscosity that is quadratic in Ptide/τc rather than linear. This uncer- tainty in the form of the viscosity remains an outstanding problem in tidal theory but here we adopt Zahn's formalism partially because simulations done by Penev et al. (2007) recover a suppression in the vertical component of the vis- cosity that scales most closely to linear. Taking radial profiles for the mixing length and con- vective velocity from our fiducial models of WASP-12, as- suming Zahn's scaling of the viscosity, and integrating over the convective envelope, yields tdiss ≈ 300 yr. From the rate of dissipation, an estimate for the orbital semi-major axis a ≈ 0.0234 au and the stellar moment of inertia I∗, one can also estimate the synchronization time (Zahn 2013) m2 bR2∗ M∗I∗ , = a 1 tdiss 1 (5) tsync which yields tsync ≈ 11 Gyr. This suggests that viscous dis- sipation of the equilibrium tide is too weak to have signifi- cantly spun up the star, a result consistent with observations assuming a low initial rotation rate. The corresponding or- bital decay rate however, tsync ≈ 1.2 Gyr, (cid:18) mba2 (cid:19) (6) = P(cid:219)P 2I∗ (cid:18) R∗ (cid:19)6 is several orders of magnitude too long to explain the ob- served decay. 4 DYNAMICAL TIDE In addition to the hydrostatic tidal response of the equilib- rium tide, there must also exist a low frequency dynami- cal response that mathematically arises from a condition for the continuity of fluid displacements across the radiative- convective boundary of the star. Dubbed the dynamical tide, this fluid response results in the excitation of internal grav- ity waves at the star's radiative-convective boundary that propagate inwards to be damped by radiative diffusion. The dynamical tide couples to the star's natural eigenfrequencies, potentially dissipating the tide at a rate orders of magnitude above the equilibrium rate if the system lies close to reso- nance. Provided the damping mechanisms acting on the dy- namical tide are efficient to the point where waves are being damped before returning to the radiative-convective bound- ary, the resonances are broadened to the point of overlap. Under the assumption that the resonances overlap, the dissi- pation rate is estimated as a frequency average over the res- onances. Adapting an expression for the frequency-averaged torque ¯τ from Kushnir et al. (2017), to WASP-12's outer convective boundary and using quantities obtained from our fiducial models, (cid:19)6(cid:32) (cid:18) R+ (cid:32) Gm2 a b a R3 + GM+ (cid:33) ¯τ ≈ 2Gm2 R+ b ≈ 2 × 10−7 (cid:33)1/2 (cid:18) ρ+ ¯ρ+ 1 − ρ+ ¯ρ+ (cid:19)2 ωorb (7) 6 A. Bailey and J. Goodman where ρ+ is the mass density at R+, ¯ρ+ is the mean mass density interior to R+, M+ is the mass interior to R+, and ωorb = 6.67 × 10−5 s−1 is the orbital angular frequency. mb 2 ¯τ √ GM∗a ≈ − 1 1 Myr Though there is an analogous torque caused by waves ex- cited at the inner radiative-convective boundary for model A, the frequency averaged torque is some seven orders of magnitude smaller as ¯τ ∝ r13/2. The circular orbital decay rate corresponding to the above torque is, (cid:219)a a = − which is fairly close to the observationally inferred decay rate (cid:219)a/a = (4.8 Myr)−1. This picture of tidal decay via the dynamical tide is a natural explanation for any models with a radiative core such as B and C. Tidally excited internal gravity waves would freely propagate inwards and break near the center of the star, resulting in the above decay rate. But as these models have their drawbacks (see §2), we now ask whether models with a convective core such as model A have a mechanism to recover the frequency-averaged torque. (8) 4.1 Damping rates The frequency-averaged torque can be recovered by model A if the radiative diffusion timescale or the viscous damping timescale in the convective envelope is comparable to the propagation time for a gravity wave. Because the Kelvin- Helmholtz timescale in roughly solar mass stars is relatively long, we operate under a quasi-adiabatic assumption in cal- culating these damping rates. The relevant linearized work integral for calculating the radiative diffusion timescale is, W ≈ − Þ R+ (∇∇∇ · FFF1) d3rrr, (9) δT T R− where δT is the Lagrangian perturbation to temperature and the Lagrangian heat flux perturbation δFFF is replaced with FFF1, its Eulerian perturbation, because the star is approxi- mately in nuclear equilibrium on timescales short compared to the main-sequence lifetime. We allow the integral to range from the inner radiative-convective boundary (R−) to the outer one (R+) rather than the whole of the star. Though there is a small positive contribution to the work integral from the convective regions, the contribution is small be- cause the work integral ends up being proportional to the superadiabatic gradient (∇ad − ∇) (cid:28) 1. Because the wave- length of these modes is small compared to a pressure scale height Hp, the opacity can be approximated as roughly con- stant and the above integral simplifies to, l(l + 1) W ≈ KT∇ad (∇ad − ∇) Þ R+ ξr2 (10) (cid:35) dr, + (cid:34)(cid:12)(cid:12)(cid:12)(cid:12) dξr dr (cid:12)(cid:12)(cid:12)(cid:12)2 r2 R− where l is a mode's angular order, K is the thermal conduc- tivity, ∇ is the temperature gradient, and ∇ad is the adiabatic temperature gradient. To determine the linear eigenfunc- tions ξr, ξh we numerically integrated the well known fourth- order set of stellar structure equations for linear, adiabatic, non-radial perturbations by shooting to a fitting point at the outer radiative-convective boundary. This yielded a ra- diative damping rate γrad ≈ (300 yr)−1. Because this is orders of magnitude slower than the propagation time Þ (cid:12)(cid:12)(cid:12)(cid:12) ∂kr ∂ω (cid:12)(cid:12)(cid:12)(cid:12) dr ≈ (cid:112)l(l + 1) Þ R+ ω2 R− dr ≈ 9 d, N r (11) tprop = radiative diffusion by itself is not significant enough to broaden the resonant peaks to the point of overlap. Although convective viscosity is not effective at dissi- pating the equilibrium tide in this system, convective viscos- ity could also damp internal gravity waves as they evanesce in convective regions. The total viscous diffusion associated with shear tensor σi j and dynamic viscosity µ in Einstein notation is (cid:219)Evisc = µ i j − 1 σ2 3 σ2 ii d3rrr ≈ µσ2 i j d3rrr. (12) Þ (cid:18) (cid:19) Solving for the squared components of the shear tensor in spherical polar coordinates, we find that the viscous work due to convection is of the form, Þ (cid:16) (cid:219)Evisc = ω2Þ −5l(l + 1)r−2(cid:60)(cid:16) r2dr µ (cid:34)(cid:12)(cid:12)(cid:12)(cid:12) dξr (cid:17) dr ∗ ξ hξr (cid:12)(cid:12)(cid:12)(cid:12)2 (cid:17) + 2 l2 + l + 1 r−2 ξr2 + l(l + 1)(l2 + l + 1)r−2 ξh2 (cid:35) . (13) For model A, this work integral corresponds to a damping rate of γvisc ≈ (300 yr)−1, still substantially long to inhibit averaging over the resonances. Though we can't justify the use of a frequency aver- aged dissipation to explain the observed decay, it's possible that we are observing this system sufficiently close to res- onance to produce a high decay rate. Adopting a damping rate, 1/γ ≡ 1/γrad + 1/γvisc, a circular orbit (consistent with observations) and uniform observation in time, the proba- bility of seeing the system with observed decay rate with its 1σ errors, −32 < (cid:219)P < −26 ms yr−1, is at the level of ≈ 10−7. Of course this probability should not be accepted in a rigorous sense because it ignores selection biases, but it is still instruc- tive to share how truly little the resonances are broadened by our selected damping mechanisms. Without an effective damping mechanism, it is still pos- sible to recover a frequency-averaged decay rate if the modes excited in the star are able to overturn stratification and break at some radius in their zone of propagation, thus de- positing all their energy. In linear theory, this criterion for breaking is simply ∆ ≡ r−1∂r (rξr) > 1. Fig. 3 shows the max- imal value of ∆ in model A for a range of frequencies. Be- cause ∆max > 1 only for ω close to resonance, wave-breaking in model A does not provide a natural explanation for the tidal decay. 4.2 Rotational effects Assuming that WASP-12b's signature is indeed due to decay via the dynamical tide, as is suspected for model B or C, the extent to which the star should have been spun up to syn- chronous rotation can be explicitly calculated. Even though WASP-12 is observed to have small surface rotation, at least some part of the core of the star should have been spun up from internal gravity waves breaking and depositing their angular momentum. We make the approximation that the relevant torque is changed appreciably only by changes to the orbit and not by changes to the star itself, so that τ ∝ aη. Given the period of time which the dynamical tide has been acting ∆t, the moment of inertia of the synchronously rotat- MNRAS 000, 1 -- 12 (2017) Understanding WASP 12b 7 Figure 4. Low-order frequency splittings due to a synchronously spinning stellar core versus its radius Rspin, or equivalently, versus the duration ∆t of the tidal torque. Color denotes the radial order for a mode while line style denotes the angular order l = 1 (solid) or l = 2 (dashed). (cid:104) Þ R∗ 0 order n, angular order l, r + l(l + 1)ξ2 ξ2 h Knl ≡ (cid:104) Þ R∗ 0 − 2ξr ξh − ξ2 h ρr2dr r + l(l + 1)ξ2 ξ2 h (cid:105) (cid:105) ρr2dr . (16) We present several of these low-order rotational splittings in Fig. 4 as a function of the size of the spinning core, and find that they are on the order of a few µHz. Compare this to the corresponding rotational splittings in a model rotat- ing uniformly at the measured v sin i∗ = 1.6 km s−1; though the splittings would be enhanced by non-zero contributions from the entire star, ultimately they would remain orders of magnitude smaller owing to a much lower νrot. 5 DISCUSSION We have seen that the apparent period change ( (cid:219)P) observed in the transits of WASP-12b cannot easily be explained as secular orbital decay. Standard mechanisms of tidal dissi- pation are too slow, unless the orbit happens to be close to resonance with a global g mode. We have estimated the probability for this to be quite small. The leading alternative explanation for the anomalous transit times is that the planetary orbit is slightly eccen- tric, e ≈ 2 × 10−3. In this interpretation, the true orbital pe- riod is constant, but the transit times depart slightly from a linear ephemeris due to precession of periastron at a rate (cid:219)ω ≈ 26 deg yr−1 (Maciejewski et al. 2016; Patra et al. 2017). The latter authors estimate that for a reasonable tidal qual- ity factor of the planet itself, Qp ≤ 106, any primordial ec- centricity would have decayed to e < 10−3 after a few million years, whereas the system age appears to be > 1 Gyr. There- fore, the eccentricity would have to be recently excited or continually forced. Figure 3. The breaking criterion in linear theory as a function of tidal frequency. The wave can sufficiently overturn stratification and break when(cid:12)(cid:12)r−1∂r (r ξr)(cid:12)(cid:12)max > 1 (cid:19) ing core is, (cid:34)(cid:18) Ispin = mba2 1 + ∆t(1 − 2η)τ √ GM∗a mb (cid:35) , (14) 1 1−2η − 1 where the τ and a refer to present day values. In the fol- lowing, we adopt the form of the torque in eq. 7 (η = −10) and use model B to estimate relevant stellar quantities. For model B, the time between the best fit model and the dis- appearance of a convective core is 6 Myr (cid:38) ∆t (cid:38) 3 Myr. But because the precursor to model B that has not yet lost its convective core still manages to fit the observables well, the actual ∆t could be made arbitrarily small and it's better to take 6 Myr (cid:38) ∆t (cid:38) 0 Myr . On the one hand, we could be seeing this system 100 years after the convective core disap- peared, in which case the the core has not been spun up sig- nificantly, but probabilistically it's most likely we're seeing this system on the order of millions of years after the con- vective core disappeared. Even models 10-100 Myr after the core disappeared don't fit the observables terribly, but using one of these values doesn't change the radial extent of the spun up core due to the steep dependence of τ on a. This fact is shown in Fig. 4 where we scale the upper abcissa with the value of the radius of the synchronously rotating core Rspin to the corresponding ∆t on the lower abcissa. This insensi- tivity of Rspin on ∆t provides a potentially testable prediction of the dynamical tide explanation -- if WASP-12b's decay is an effect of the dynamical tide, the innnermost ≈ 0.2R(cid:12) of WASP-12 itself should be rapidly rotating. Because the linear rotational frequency of this core νrot = 66 µHz is significantly less than the linear eigenfre- quencies of our subgiant model, the rotational splittings can be estimated in a perturbative manner as in Aerts et al. (2010). For azimuthal order m, the splittings δν can be writ- ten: δν = m Knl(r)νrot(r)dr, (15) Þ R∗ 0 where Knl is the unnormalized rotational kernel for radial MNRAS 000, 1 -- 12 (2017) 10.511.011.512.0ω(day−1)10−310−210−1100101(cid:12)(cid:12)(cid:12)r−1∂r(cid:0)rξr(cid:1)(cid:12)(cid:12)(cid:12)max10−610−510−410−310−210−1100101102103∆t(Myr)2468δν/m(µHz)0.010.050.100.150.200.250.28Rspin(R(cid:12))n=0n=1n=2n=3 8 A. Bailey and J. Goodman We now briefly examine mechanisms for forcing the ec- centricity or modulating the period of the orbit via changes in host star or third bodies. In the following, unless otherwise noted, we take M∗ = 1.4 M(cid:12), which is slightly higher than any of the values in Table 2. Therefore R∗ = 1.64 R(cid:12) based on the mean den- sity adopted in Table 1. With Southworth (2012)'s result that Rb/R∗ ≈ 0.1159 ± 0.0033, we then have Rb ≈ 1.89RJ for the planetary radius. Both radii would scale ∝ M1/3∗ to other assumed values of the stellar mass. Adopting the ra- dial velocity amplitude K = 221.9 ± 3.1 m s−1 from Knut- son et al. (2014) and the inclination I = (83 ± 0.5)◦ from Maciejewski et al. (2013), and since mb(M∗ + mb)−2/3 = K∗ sec I(P/2πG)1/3, we then have mb = 1.41 MJ; this scales approximately as M1/3∗ a = (P/2π)2/3[G(M∗ + mb)]1/3 ≈ 0.02322 au. . Finally the semimajor axis becomes 5.1 Eccentricity from convection Phinney (1992) proposed that the small measured eccentrici- ties of binary millisecond pulsars with white-dwarf compan- ions can be explained by potential fluctuations associated with convection in the envelope of the companion when on the giant or asymptotic-giant branch. With few exceptions, the orbits of subsequently discovered binary millisecond pul- sars have conformed well to the predictions of this model (Lorimer 2008). mass µ ≈ 1.4 MJ, Phinney's equation (7.33) reads Adapted to the WASP-12 system, so that the reduced (cid:18) L∗Renv 5 L(cid:12) R(cid:12) · 1.4M(cid:12) M∗ (cid:19)1/3(cid:18) Menv 0.0004 M(cid:12) (cid:19)1/6 , (cid:104)e2(cid:105)1/2 ≈ 2 × 10−5 (17) in which Renv ≈ 1.4 R(cid:12) is the radius at the base of the outer convection zone in our preferred model for WASP-12. Menv, the mass of that zone, is sensitive to the effective tempera- ture, metallicity, and evolutionary state of the star, but in view of the sixth root, no plausible value of Menv could make up the two orders of magnitude by which the r.m.s. eccen- tricity predicted by eq. (17) falls short of the value required to explain the quadratic term in the transit ephemeris. Fur- thermore, as Phinney remarks, his eq. (7.33) probably over- estimates the eccentricity expected when the turnover time of the largest convective eddies exceeds the tidal period, as occurs in WASP-12 by at least one order of magnitude. 5.2 The Applegate effect Applegate & Patterson (1987) and Applegate (1992) sug- gested that long-term modulations observed in the eclipse times of some close stellar binaries, including V471 Tau and Algol, are caused by slow changes in the quadrupole moment of one or both stars induced by their magnetic cycles. In the later version of this idea, the magnetic stress is not large enough to distort the equilibrium shape of the star directly, but rather slowly redistributes angular momentum within the star(s), leading to changes in the rotationally-induced quadrupole. Because the changes are slow, they would not excite the eccentricity of the orbit, but the quadrupole con- tributes to the central force between the stars and hence to the orbital period itself. Watson & Marsh (2010, here- after WM10) have scaled Applegate (1992)'s model to sev- eral exoplanet systems. For WASP-12b, they estimate that the anomaly in the transit time (O − C, observed minus cal- culated) could be as much as 42 (T/50 yr)3/2, where T is the period on which the dynamo modulates the internal differ- ential rotation. This last could be the same as the period of the magnetic dipole, or half that, depending on the type of dynamo. MW10's predicted variation is not a great deal smaller than the ∼2-minute departure from a linear transit ephemeris found by Patra et al. (2017). It depends on several several uncertain parameters besides the dynamo period T, so one ought to consider whether the uncertainties in these parameters might allow the Applegate effect to explain the WASP 12 data. The relevant parameters are the rotation period of the star, for which WM10 take Prot = 36 d, the fractional mass of the convection zone, for which they take Menv/M∗ = 0.1, and the portion of the mean luminosity that is converted to mechanical form to change the differential rotation. For the latter they take ∆L = 0.1 L; this seems large, but perhaps not in direct conflict with observations because, as they point out, the luminosity variation at the photosphere could be much smaller due to the thermal iner- tia of the convection zone (i.e., the ratio of its total thermal energy to the luminosity of the star; this is about 300 yr for WASP 12). WM10's equations imply that the transit-time anomalies scale with these parameters as follows: (O − C)max ∝ T3/2P−1 rot (18) (cid:18) Menv M∗ (cid:19)1/2 (∆L)1/2 . The mass of the convective envelope of WASP 12 is probably (cid:46) 10−3 M∗, as remarked above; following eq (18), this would reduce the predicted O−C by an order of magnitude. On the other hand, the rotation period may be rather less than the assumed value if the star is viewed near pole on, as Rossiter- MacLaughlin measurements suggest (Albrecht et al. 2012). The median rotation period for main-sequence F8 stars2 is ≈ 8 d (Nielsen et al. 2013). Since dynamo periods appear to correlate positively with stellar rotation periods (Saar & Brandenburg 1999; Bohm-Vitense 2007), however, the pos- itive scaling with T seems likely to overwhelm the negative scaling with Prot in eq. (18). dict a variation ∆J2 (cid:38) 5 × 10−8 in its rotationally-induced If MW10's scalings are applied to the Sun, they pre- dimensionless quadrupole moment over the dynamo cycle. The internal differential rotation of the Sun has been di- rectly constrained by helioseismology, and for a significant fraction of a cycle. Antia et al. (2008) have used these data to estimate that (cid:104)J2(cid:105)(cid:12) = 2.2 ± 0.01 × 10−7, and the variation over a nine-year period to be (cid:46) 1 × 10−10, i.e. several or- ders of magnitude smaller than MW10's assumptions would predict. For these reasons (i.e., both our estimates of the ac- tual parameters of WASP-12, as well as comparision with heliosesimological inferences for the Sun), it is unlikely that the Applegate effect explains the transit-time anomolies of WASP-12b. 2 Hebb et al. (2009) classify WASP 12 as F9V MNRAS 000, 1 -- 12 (2017) 5.3 Bow shock b + v2 b ρw(v2 Ultraviolet absorption is seen just before each transit of WASP-12b and has been interpreted as evidence for mass loss from the planet through its inner Lagrange point (Fos- sati et al. 2010a). Alternatively, this could be the signature of a bow shock ahead of the planet encountering a wind from the star (Lai et al. 2010; Vidotto et al. 2010). Such a shock would exert a drag on WASP-12b's orbit. As shown here, however, an improbably dense wind would be required to The torque exerted on the planet by the shock is timescale is then mb b ρw(v2 explain the observed (cid:219)P. w)1/2vba, in which CD is a factor of order CD πR2 unity (the drag coefficient), ρw the pre-shock density of the wind, vw the wind velocity, Rb ≈ 1.9 RJ the radius of the planet, and vb ≈ (GM∗/a)1/2 the orbital velocity. The decay P(cid:219)P For the numerical estimate, we have taken CD = 0.3, and ρw = 2×10−18 g cm−3 (i.e. nH = 1.5×106 cm−3); the latter fol- of 10−12.3(vwind/100 km s−1) M(cid:12) yr−1. In order to explain the apparent decay rate (P/ (cid:219)P ≈ 3 Myr), the wind density would have to increase some six orders of magnitude, making the mass-loss timescale of the star (cid:46) 10 Myr. This is unreason- able as the star is probably older than 1 Gyr. lows Vidotto et al. (2010) and implies a stellar mass-loss rate ≈ 4 × 1012 yr. 3πCD R2 b + v2 w)1/2 (19) = 5.4 Kozai-Lidov oscillations We consider the possibility that a non-transiting third body in the system continuously excites a small eccentricity in the orbit of WASP-12b so that, as suggested by Maciejewski et al. (2016), the transit-time anomolies result from apsidal precession of the slightly elliptical orbit. Apsidal precession itself imposes a lower bound on the perturbations that such a hypothetical companion must ex- ert to excite WASP-12b's eccentricity. Let the companion have mass mc, semimajor axis ac, and orbital eccentricity ec, and let {mb, ab, eb} be those of WASP-12b itself. By a standard calculation in secular perturbation theory, one can show that if eb (cid:28) 1 initially, then eb will grow by the Kozai- Lidov mechanism (hereafter KLM) only if M2∗ R5 b mba8 b , 10 3 k2b (20) c)3/2 > mc c(1 − e2 a3 in which Rb is the radius of WASP-12b and k2b its Love number, these two quantities being important for the apsidal precession rate. The inequality (20) assumes that the orbital planes of mc and mb are orthogonal, which maximizes the efficiency of the KLM. We are also assuming ac > ab, i.e. the third body's orbit is exterior to that of WASP-12b. The orbits should not cross, whence ac(1− ec) > ab, and therefore abac. With k2b ≈ 0.6, the lower bound on ac(1 − e2 the companion's mass for the KLM becomes c)1/2 > √ mc > 77. M⊕ (21) An upper bound on mc follows from the published radial-velocity data (Hebb et al. 2009; Husnoo et al. 2011; Albrecht et al. 2012; Bonomo et al. 2017). After subtraction MNRAS 000, 1 -- 12 (2017) (cid:19)3/2 (cid:18) ac ab (cid:113) (cid:113) 1 − e2 1 − e2 c Understanding WASP 12b 9 of the WASP 12b signal3 and correction for the nominal measurement errors, these data have variance ≈ (9 m s−1)2. The RV signal of the hypothetical WASP-12c should be no larger than this. Therefore mc < 18 f −1/2 M⊕ , (22) (cid:19)1/2 (cid:18) ac ab with f being a geometrical factor that determines the mean- square projection of the orbital velocity onto the line of sight: f (ec, ωc, Ic) = sin2 ωc + c cos2 ωc sin2 Ic . (23) 1 + (cid:112)3/5 of the two planetary orbits must be greater than sin−1(cid:112)2/5 ≈ In order that the KLM operate, the relative inclination 39.2◦, so cos Ic cos Ib + sin Ic sin Ib cos(Ωc − Ωb) < with Ωb,c being the longitudes of the ascending nodes. Since the inclination of WASP 12b is Ib ≈ (83± 0.5)◦ (Maciejewski et al. 2013), the above constraint is compatible with Ic ≈ 0, and of course also with any eccentricity ec or argument of periastron ωc. So the factor f could be arbitrarily small. The two inequalities (21) & (22) could therefore both be satisfied by an exterior perturber (ac > ab), although this becomes less probable as the separation between the orbits increases because of the different scalings with ac/ab. Fur- thermore, eqs. (22)-(23) suppose that the radial velocity is measured continuously, whereas in fact it is sampled some- what sparsely and irregularly: nearly half of the ∼ 90 mea- surements were made by Albrecht et al. (2012) in a single night. If WASP 12c's orbit were highly eccentric, and thus hovering usually near apastron, its full radial-velocity am- plitude might not be sampled. We have not systematically investigated the probabil- ity that both of the mutually antagonistic bounds (21) and (22) could be satisfied. Nevertheless, the Kozai-Lidov mech- anism does not seem to provide a natural explanation for the quasi-secular transit-time anomalies of WASP 12b. The hypothesis is attractive only in comparison to all of the other possibilities that we have investigated. 5.5 Resonance We have considered the possibility that the orbital variations of WASP 12b are caused by resonant interactions with an unseen planet. We focus on mean-motion resonances. Suppose first a 1:1 resonance, in other words, a small trojan planet librating around the stable Lagrange points of the WASP 12+WASP 12b system.4 The inferred amplitude of the period variation is 29 ± 3 ms yr−1 (Patra et al. 2017), amounting to ∆ ln P ≈ 3 × 10−6 over the 9 years that tran- sits have been monitored. We estimate that a roughly lunar 3 We subtract an optimally scaled multiple of the photometric ephemeris of Patra et al. (2017), including their secular period derivative (cid:219)P = (0.92 ± 0.01) × 10−9. Thus this limit applies to com- panions with periods less than the span, of the data, ∼ 7 yr. 4 We thank Scott Tremaine for suggesting that we look into this. 10 A. Bailey and J. Goodman mass in a "horseshoe" 1:1 resonant libration could modu- late WASP 12b's period at this amplitude. This would eas- ily satisfy the limit mc < 34 M(cid:12) on Trojan companions to WASP-12b found by Lillo-Box et al. (2018), who based their analysis on archival radial velocities. The difficulty, how- ever, is in the period of the modulation. It is well known that small-amplitude librations around the Lagrange points Plib = Porb ×2(1 + q)/√ in the coplanar restricted three-body problem have period 27q, where Porb is the orbital period of the massive bodies and q < 0.04 is their mass ratio. In the present case where Porb = 1.09 d and q ≈ 10−3, Plib ≈ 13 d. A large-amplitude libration can have a somewhat longer pe- riod than this, but not by more than a factor ∼ 2 unless very close to the separatrix between libration and circulation, as we have convinced ourselves by numerical experiments. Such a Plib is far too short to mistaken for a secular trend over 9 yr unless severely aliased, which seems unlikely in view of the density of transit observations [see the tabulation in Patra et al. (2017)]. We have also examined first-order mean motion reso- nances Pc : Pb ≈ (j +1) : j, with j ≥ 1 an integer. Our analysis is restricted to coplanar, near-circular cases, but the main conclusions would probably be similar even for strongly mis- aligned orbits. The unseen body WASP-12c is presumed to be much less massive than WASP-12b. Close to such a resonance, the jth azimuthal harmonic of the potential of the orbit of b directly forces the eccentricity of c's orbit (ec), and the (j+1)th harmonic of c forces eb. In the first case, or "exterior" resonance, eb is neglected to leading order, while ec is neglected for the interior resonance (e.g. Murray & Dermott 2000). The forced eccentricities depend not only on the masses mc and mb but also on the distances from exact resonance; these differ because of the unforced apsidal precession rates of the two planets. As already noted in §5.4, the apsidal precession of b is dominated by its tidal distortion: b0 ≈ 3.9 × 10−4nb, with nb = 2πP−1 b being its mean motion. If c is a smaller body such as a super-earth, its apsidal motion is dominated by the axisymmetric potential of b's orbit. Near the 2:1 resonances, we estimate that (cid:219)c0 ≈ 3.8 × 10−4nb. Because of the coincidence that (cid:219)b ≈ (cid:219)b, the slow frequencies that measure the distance from resonance, namely νb ≡ jnb −(j + 1)nc − (cid:219)b0 and νc ≡ jnb −(j + 1)nc − (cid:219)c0 will usually be nearly equal, at least for the 2:1 resonances (j = 1). Tidal dissipation within the planets damps the forced eccentricity at the rate (cid:18) d ln e (cid:19) dt γp ≡ − 63 4Q(cid:48) p M∗ mp = tide (cid:19)5 (cid:18) Rp ap np (24) where Q(cid:48) p is the tidal quality factor of planet p corrected for its Love number. On short timescales ∼ ν−1, an equilibrium holds between forcing and damping. Secularly however, at second order in eccentricity and first order in the damping rate (24), there is a transfer of orbital energy and angular momentum between planets. The transfer is always outward, i.e. from b to c in our case, but in the proportion ∆E = nc∆J for the interior resonance (where the orbit of c is approxi- mated as circular), and ∆E = nb∆J for the exterior resonance (where ∆eb is neglected). The rate of transfer of angular mo- mentum is related to the tidal dissipation rates (cid:219)Eb,c > 0 by b/4 γbmb A2 ν2 b + γ2 b (cid:20) where Ab = Gmc abac (cid:21) dJc dt = − dJb dt = (cid:219)Eb + (cid:219)Ec nb − nc (25) The effect of this torque is to increase the slow frequencies νb and νc, and hence to increase the distance from resonance if these frequencies are already positive. If body c is a super-earth, we estimate that (cid:219)Eb > (cid:219)Ec by a factor of at least a few at the first few mean-motion resonances (j (cid:46) 6): (cid:219)Eb = (26) , (cid:33)1/3 (cid:32) 106 Q(cid:48) b d d ln α (j+1) 1/2 (α) + 2jb b (j+1) 1/2 (α) , α=ab/ac (27) (j+1) 1/2 (α) are the usual Laplace coef- in which the functions b ficients. The last equation follows from first-order epicyclic theory if the damping term is inserted by hand. (These equa- tions also determine (cid:219)Ec if all subscripts "b" and "c" are in- terchanged and j + 1 is replaced by j.) For definiteness, let us focus on the 2:1 resonance, j = 1, so that νb ≈ νc by the numerical coincidence noted above. Presuming that mc (cid:28) mb, the increase in ν due to the torque (25) is dominated by the change in the mean motion of c, but dEc/dJc ≈ nc because dissipation occurs mainly in body b. Hence dnc/dJc ≈ −3/mca2 c . In the relevant regime where ν (cid:29) γ, dν/dt ∝ ν−2 because of the denominator in eq. (27), the other terms in eqs. (26)-(27) being effectively constant when ν (cid:28) nb,c. Integrating this relation with the constants included yields ν ≈ 0.02 mc M⊕ T Gyr nb , (28) presuming that the system started from exact resonance at time T in the past. The quantities in parentheses in eq. (28) are uncertain, but because of the cube root, it is unlikely that the distance from resonance (ν/nb) is much less than 10−2. Now at a (j+1) : j resonance, the combination (j+1)nc−jnb is the forced (cid:219)b. Therefore νb = (cid:219)b0 − (cid:219)b. Since apsidal precession rate, we have previously estimated that (cid:219)b0 ≈ 4×10−4nb, it follows from eq. (28) that (cid:219)b < 0, with a period ∼ 50 × Pb ≈ 55 d. Thus while it is possible to choose mc so that the amplitude of the forced eccentricity eb = 2 × 10−3, the period of the quasi-secular (cid:219)P. apsidal precession is much too rapid to explain the observed 6 SUMMARY We have revisited the possible causes of WASP-12b's depar- ture from a linear ephemeris. Either the orbit is decaying, or some dynamical perturbation maintains a small eccen- tricity and the apsides precess on some period longer than a decade. We have considered various perturbations induced by unseen third bodies or distortions of the star WASP-12 itself, but none is consistent with all of the observational constraints, at least not without fine tuning. MNRAS 000, 1 -- 12 (2017) The conclusion therefore seems inescapable that the or- bit is indeed decaying, presumably because of tidal dissipa- tion in the star. Indeed, the dynamical tide -- computed for a circular orbit and a negligibly rotating star -- naturally yields an orbital lifetime comparable to what is inferred from tran- sit timing. But this requires that the star has evolved onto the subgiant branch and lost its convective core, as Wein- berg et al. (2017) have suggested. In that case, the g modes excited at the base of WASP-12's thin surface convection zone might be just strong enough to damp nonlinearly in the core, which would broaden the g-mode resonances so that they overlapped. If WASP-12 were still on the main- sequence and still had its convective core, the resonances would be very sharp, and the orbit would have to be im- plausibly close to resonance to explain the current rate of orbital evolution. Alternatively, if the star had a rapidly ro- tating core, with a rotation period as short or shorter than the period of the orbit, then the tidally excited g-modes would be absorbed at the critical (corotation) layer (Barker & Ogilvie 2010); the torque applied by absorption of the in- going waves would then maintain the rapid rotation of the layer and presumably of the core beneath it. This begs the question how the core could have started out with such rapid rotation, however. Moreover, unlike the subgiant hypothesis, it does not naturally explain why the decay timescale is so much shorter than the age of the star. The observational constraints on WASP-12 itself, when fit to theoretical models for its structure made with the mesa code, favor a main-sequence star rather than a subgiant. Actually, we have not been able to find any mesa model that fits all of the observations comfortably: the spectroscopically inferred Teff and [Fe/H] are in tension with the luminosity inferred from the GAIA-DR2 distance and Stassun et al. (2017)'s bolometric flux. This problem would exist even if there were no evidence for orbital decay, though the transit light curves are essential for constraining the star's mean density. We thank Josh Winn for introducing us to this problem and for much helpful advice and conversation. ACKNOWLEDGEMENTS This work has made use of data from the European Space Agency (ESA) mission Gaia (https://www.cosmos.esa. int/gaia), processed by the Gaia Data Processing and Analysis Consortium (DPAC, https://www.cosmos.esa. int/web/gaia/dpac/consortium). Funding for the DPAC has been provided by national institutions, in particular the institutions participating in the Gaia Multilateral Agree- ment. REFERENCES Aerts C., Christensen-Dalsgaard J., Kurtz D. W., 2010, Astero- seismology. Springer Albrecht S., et al., 2012, ApJ, 757, 18 Allende Prieto C., Asplund M., Garc´ıa L´opez R. J., Lambert Understanding WASP 12b 11 Applegate J. H., Patterson J., 1987, ApJ, 322, L99 Asplund M., Grevesse N., Sauval A. J., Scott P., 2009, ARA&A, 47, 481 Aufdenberg J. P., Ludwig H.-G., Kervella P., 2005, ApJ, 633, 424 Barker A. J., Ogilvie G. I., 2010, MNRAS, 404, 1849 Bechter E. B., et al., 2014, ApJ, 788, 2 Bedding T. R., et al., 2010, ApJ, 713, 935 Bergfors C., et al., 2013, MNRAS, 428, 182 Bohm-Vitense E., 2007, ApJ, 657, 486 Bond H. E., et al., 2015, ApJ, 813, 106 Bonomo A. S., et al., 2017, A&A, 602, A107 Choi J., Dotter A., Conroy C., Cantiello M., Paxton B., Johnson B. D., 2016, ApJ, 823, 102 Collins K. A., Kielkopf J. F., Stassun K. G., 2017, AJ, 153, 78 Fossati L., et al., 2010a, ApJ, 714, L222 Fossati L., et al., 2010b, ApJ, 720, 872 Gaia Collaboration et al., 2016, A&A, 595, A1 Gaia Collaboration Brown A. G. A., Vallenari A., Prusti T., de Bruijne J. H. J., Babusiaux C., Bailer-Jones C. A. L., 2018, preprint, (arXiv:1804.09365) Goldreich P., Nicholson P. D., 1977, Icarus, 30, 301 Goodman J., Dickson E. S., 1998, ApJ, 507, 938 Green G. M., et al., 2018, MNRAS, 478, 651 Hebb L., et al., 2009, ApJ, 693, 1920 Husnoo N., et al., 2011, MNRAS, 413, 2500 Husnoo N., Pont F., Mazeh T., Fabrycky D., H´ebrard G., Bouchy F., Shporer A., 2012, MNRAS, 422, 3151 Jerzykiewicz M., Molenda-Zakowicz J., 2000, Acta Astron., 50, 369 Knutson H. A., et al., 2014, ApJ, 785, 126 Kushnir D., Zaldarriaga M., Kollmeier J. A., Waldman R., 2017, MNRAS, 467, 2146 Lai D., Helling C., van den Heuvel E. P. J., 2010, ApJ, 721, 923 Levrard B., Winisdoerffer C., Chabrier G., 2009, ApJ, 692, L9 Lillo-Box J., Barrado D., Figueira P., Leleu A., Santos N. C., Correia A. C. M., Robutel P., Faria J. P., 2018, A&A, 609, A96 Lorimer D. R., 2008, Living Reviews in Relativity, 11 Maciejewski G., et al., 2013, A&A, 551, A108 Maciejewski G., et al., 2016, A&A, 588, L6 Mortier A., Santos N. C., Sousa S. G., Fernandes J. M., Adibekyan V. Z., Delgado Mena E., Montalto M., Israelian G., 2013, A&A, 558, A106 Murray C. D., Dermott S. F., 2000, Solar System Dynamics. Cam- bridge University Press Nielsen M. B., Gizon L., Schunker H., Karoff C., 2013, A&A, 557, L10 Patra K. C., Winn J. N., Holman M. J., Yu L., Deming D., Dai F., 2017, preprint, (arXiv:1703.06582) Paxton B., Bildsten L., Dotter A., Herwig F., Lesaffre P., Timmes F., 2011, ApJS, 192, 3 Paxton B., et al., 2013, ApJS, 208, 4 Paxton B., et al., 2015, ApJS, 220, 15 Paxton B., et al., 2018, ApJS, 234, 34 Penev K., Sasselov D., Robinson F., Demarque P., 2007, ApJ, 655, 1166 Penev K., et al., 2016, AJ, 152, 127 Phinney E. S., 1992, Philosophical Transactions of the Royal So- ciety of London Series A, 341, 39 Remus F., Mathis S., Zahn J.-P., 2012, A&A, 544, A132 Saar S. H., Brandenburg A., 1999, ApJ, 524, 295 Southworth J., 2012, MNRAS, 426, 1291 Stassun K. G., Collins K. A., Gaudi B. S., 2017, AJ, 153, 136 Terquem C., Papaloizou J. C. B., Nelson R. P., Lin D. N. C., 1998, ApJ, 502, 788 D. L., 2002, ApJ, 567, 544 Torres G., Fischer D. A., Sozzetti A., Buchhave L. A., Winn J. N., Antia H. M., Chitre S. M., Gough D. O., 2008, A&A, 477, 657 Applegate J. H., 1992, ApJ, 385, 621 Holman M. J., Carter J. A., 2012, ApJ, 757, 161 Vidotto A. A., Jardine M., Helling C., 2010, ApJ, 722, L168 MNRAS 000, 1 -- 12 (2017) 12 A. Bailey and J. Goodman Watson C. A., Marsh T. R., 2010, MNRAS, 405, 2037 Weinberg N. N., Sun M., Arras P., Essick R., 2017, ApJ, 849, L11 Zahn J. P., 1966, Annales d'Astrophysique, 29, 489 Zahn J.-P., 1977, A&A, 57, 383 Zahn J.-P., 2013, in Souchay J., Mathis S., Tokieda T., eds, Lecture Notes in Physics, Berlin Springer Verlag Vol. 861, Lecture Notes in Physics, Berlin Springer Verlag. p. 301, doi:10.1007/978-3-642-32961-6 8 van Saders J. L., Pinsonneault M. H., 2012, ApJ, 746, 16 This paper has been typeset from a TEX/LATEX file prepared by the author. MNRAS 000, 1 -- 12 (2017)
1903.12182
1
1903
2019-03-28T18:00:01
TESS Photometric Mapping of a Terrestrial Planet in the Habitable Zone: Detection of Clouds, Oceans, and Continents
[ "astro-ph.EP", "astro-ph.IM" ]
To date, a handful of exoplanets have been photometrically mapped using phase-modulated reflection or emission from their surfaces, but the small amplitudes of such signals have limited previous maps almost exclusively to coarse dipolar features on hot giant planets. In this work, we uncover a signal using recently released data from the Transiting Exoplanet Survey Satellite (TESS), which we show corresponds to time-variable reflection from a terrestrial planet with a rotation period of 0.9972696 days. Using a spherical harmonic-based reflection model developed as an extension of the STARRY package, we are able to reconstruct the surface features of this rocky world. We recover a time-variable albedo map of the planet including persistent regions which we interpret as oceans and cloud banks indicative of continental features. We argue that this planet represents the most promising detection of a habitable world to date, although the potential intelligence of any life on it is yet to be determined.
astro-ph.EP
astro-ph
Draft version April 1, 2019 Typeset using LATEX modern style in AASTeX62 TESS Photometric Mapping of a Terrestrial Planet in the Habitable Zone: Detection of Clouds, Oceans, and Continents Rodrigo Luger,1 Megan Bedell,1 Roland Vanderspek,2 and Christopher J. Burke2 1Center for Computational Astrophysics, Flatiron Institute, New York, NY 2Kavli Institute for Astrophysics and Space Research, Massachusetts Institute of Technology, Cambridge, MA ABSTRACT To date, a handful of exoplanets have been photometrically mapped using phase- modulated reflection or emission from their surfaces, but the small amplitudes of such signals have limited previous maps almost exclusively to coarse dipolar features on hot giant planets. In this work, we uncover a signal using recently released data from the Transiting Exoplanet Survey Satellite (TESS), which we show corresponds to time- variable reflection from a terrestrial planet with a rotation period of 0.9972696 days. Using a spherical harmonic-based reflection model developed as an extension of the starry package, we are able to reconstruct the surface features of this rocky world. We recover a time-variable albedo map of the planet including persistent regions which we interpret as oceans and cloud banks indicative of continental features. We argue that this planet represents the most promising detection of a habitable world to date, although the potential intelligence of any life on it is yet to be determined. (cid:135) Keywords: methods: data analysis, techniques: photometric, planets and satellites: oceans, planets and satellites: surfaces, planets and satellites: terrestrial planets 1. INTRODUCTION Of the many challenges and opportunities facing the field of exoplanets in the future, the prospect of mapping the surface of a potentially habitable planet is one of the most exciting -- and one of the most difficult. Present-day missions including the Transiting Exoplanet Survey Satellite (TESS; Ricker et al. 2015) aim to discover small terrestrial planets around bright, nearby stars. Once these targets are found, more detailed observational characterization must follow to evaluate their suitability [email protected] 2 for harboring life. The discovery of surface features such as continents and oceans would be a critical step toward understanding the habitability of an alien world. The most straightforward method of observing features on a terrestrial exoplanet would be to directly image the planet at sufficient resolution. However, such an obser- vation would require an effective telescope aperture many orders of magnitude larger and equipped with far more advanced coronagraphic optics than what we have avail- able today. Fortunately, surface mapping can also be done indirectly by observing rotational phase-dependent variations of the planetary signal in precise photometric time series. This is a complex data analysis problem: strong degeneracies are in- volved when transforming a 1-dimensional time series into a 2-dimensional surface map (Cowan et al. 2013). Interpretation is also challenging in the face of many po- tential explanations for a poorly resolved surface feature. Nevertheless, this technique has already proven successful in the mapping of large dipolar features on the surfaces of hot giant planets, most notably the hot Jupiter HD189733b (Knutson et al. 2007; Majeau et al. 2012; de Wit et al. 2012). More recently, this technique has also been applied to the potentially terrestrial planet 55 Cancri e, suggesting a large longitudi- nal offset in its peak surface emission due to either lava flows on the surface or strong atmospheric recirculation (Demory et al. 2016a,b; Hammond & Pierrehumbert 2017). However, obtaining a map of a temperate terrestrial exoplanet is still beyond the capabilities of current facilities, as both the thermal and reflected light signals from these planets are orders of magnitude below current measurement capabilities. Nevertheless, much work has been done on the theoretical front on mapping terrestrial planets in the habitable zone (e.g., Kawahara & Fujii 2010; Fujii & Kawahara 2012; Berdyugina & Kuhn 2017; Haggard & Cowan 2018; Lustig-Yaeger et al. 2018); for a review, see Cowan & Fujii (2018). In this work, we show how we can harness photometric data collected by the TESS mission to construct a map of Sol d, a rocky planet with a radius of 6.4×106 m orbiting near the inner edge of the habitable zone of its star, a nearby G2 main sequence star (Sagan et al. 1993). Although the TESS mission promised to deliver a handful of terrestrial planets in the habitable zone for follow-up with other missions, it came as a great surprise that data from the spacecraft could be used not only to detect Sol d but to characterize it in detail. This is due primarily to the strong nature of the reflected light signal originating from the planet, which because of complex internal optics illuminates the entirety of the TESS detector during most of Sector 1 and part of Sector 2. Previous attempts to map Sol d from the exoplanet community have suggested the presence of localized surface features, but their resolving power has been limited by the cadence and/or duration of observations (Cowan et al. 2009; Jiang et al. 2018). TESS offers precise 2-minute cadence data spanning a wide range of illumination phases, enabling a level of spatial and temporal resolving power that is unprecedented in the history of exoplanet phase mapping. Moreover, recent advances on the modeling side 3 have made the reconstruction of detailed maps more feasible than ever. Luger et al. (2019) developed an analytic algorithm for generating -- and inverting -- light curves of stars and planets whose surface emission is described in terms of spherical harmonics, dubbed starry. The analyticity of this method makes the computation of the model both fast and precise. In turn, Haggard & Cowan (2018) derived analytic expressions for the analogous case in reflected light, packaged into the EARL code. In this work, we derive similar expressions within the starry framework to perform fast, analytical inversion of the light curve of Sol d and thus construct a map of the planet in reflected light. The paper is organized as follows: we give an overview of the TESS data used in §2, and in §3 we describe the methods employed to infer a map from these data, including the adaptation of the starry algorithm to reflection mapping, the modeling of spacecraft-related systematics, and the likelihood and priors used. We present results in §4 and make a comparison of these findings to other state-of-the-art maps of Sol d in §5. Finally, we conclude with a summary of our major findings in §6. The code used to generate all the figures in this paper is open source and available on GitHub1. Clickable icons ƭ next to each figure caption link to the source code used to generate them. 2. DATA The TESS pipeline produces processed light curves for all short-cadence (2 minute) targets. Part of this processing is the determination of the localized background flux within each postage stamp, which is provided as an ancillary time series. As with all TESS time series data products, the time is given in units of TESS Julian Date (TJD), defined as BJD− 2457000, and the data are delivered in 27-day parcels corresponding to the varying pointing sectors of TESS. We use these background flux measurements from Sectors 1 and 2 as our primary data source in this work. For each of the two sectors considered, we downloaded a random subsample of 1,000 short-cadence light curve files corresponding to postage stamps located across the full TESS field of view and extracted the background flux time series (SAP_BKG) and corresponding uncertainties (SAP_BKG_ERR) from each. We then interpolated all light curves in a given sector onto the same time grid, masked all cadences with nonzero QUALITY flags, and masked cadences during which Sol d was below the edge of the sunshade. Not all postage stamps are equally informative, since the desired signal from Sol d is inhomogeneously spread across the TESS detectors and manifests itself differently in different regions (see §3.2). We down-selected our pool of 1,000 targets to 75 maximally useful background light curves from Sector 1 and 20 from Sector 2. This down-selection was done with several criteria in mind: each light curve should contain a significant signal from Sol d; they should be representative of the general background 1 https://github.com/rodluger/earthshine 4 Figure 1. Normalized background flux in each of ten postage stamps for each of Sectors 1 and 2. Cadences where Sol d was below or very close to the edge of the sunshade were masked. While the planet is in view for the majority of both orbits in Sector 1, it is only visible in the data at the very end of each orbit in Sector 2. ƭ flux modulation seen across the TESS cameras, rather than containing any truly localized signals from astrophysical background sources; and the sample as a whole should contain the full range of typical behaviors of background across the entire TESS field of view. To accomplish this, we applied a categorization and outlier rejection scheme to the ensemble of light curves. During this selection step only, we normalized the light curves to the same min-max range. We then computed the median flux at each cadence along with an estimate of the variance based on the median absolute deviation (MAD) across all light curves. For each light curve, we calculated a score χ2 equal to the sum of the squares of the deviation from the median at each of the N cadences, normalized by the variance. We removed all light curves with χ2 < N from the pool, placing them in a separate group. We repeated these steps several times, classifying light curves into four groups for Sector 1 and two groups for Sector 2. We found that this procedure effectively grouped together light curves with similar features, increasing in variation within each group from the first group (most homogeneous) to the last group (most heterogeneous). In practice, we found light curves in all but the last group to be dominated by the signal of Sol d, while light curves in the last group were dominated by signals other than the periodic modulation of the planet. To obtain our final set of light curves, we sorted the light curves in each group by the amplitude of the signal and kept the first 25 in each of the first three groups in Sector 1 and the first 20 in the first group in Sector 2 for a total of 75 and 20 light curves, respectively. We then computed an estimate of the non-Sol d background flux from the cadences during which the planet was below the edge of the sunshade and removed this baseline from the flux. Finally, we median-normalized all light curves (as opposed to the min-max normalization used for the classification). Our final dataset consisted of 912,605 data points in total. A representative sample of these signals is shown in Figure 1. 133013401350136013701380Time [TJD]0.00.51.01.52.02.5Normalized FluxS1.1S1.2S2.1S2.2 3. METHODS 3.1. Orbital Modeling 5 We downloaded2 the Solar System and TESS ephemeris data for Sectors 1 and 2 and used spiceypy (Acton 1996; Acton et al. 2017; Annex 2017) to convert the TJD timestamps to ephemeris time (ET). We then used the spkezr utility of spiceypy to compute the relative positions of Sol, Sol d, and TESS for every cadence in our dataset in the J2000 equatorial reference frame. In this right-handed coordinate frame, Sol d is centered at the origin, the y axis points along the planet's spin axis, and the x axis points toward the vernal equinox. We do not apply any light travel correction, as the expected delay is on the order of a few seconds. At any given cadence, we rotate Sol d about its spin axis to the correct phase, assuming a planetary rotation period of 0.9972696 days and an obliquity of 23.437◦. We determine the initial rotation phase of Sol d by computing the Sol d Rotation Angle (ERA), given by (Urban & Seidelmann 2013) ERA = 360◦(0.7790572732640 + 1.00273781191135448 tU ) mod 360◦ (1) where tU = t − 2451545.0. At t = t0 = 2458325.5, the first timestamp in our dataset, we find ERA = 303.4◦, meaning the vernal equinox will be aligned with the prime meridian 360◦ − 303.4◦ = 56.6◦ = 3.77 hours past t0. Finally, we rotate the frame to align TESS with the +z axis, with the spin vector of Sol d along the y − z plane. We verified our rotations by comparing our results on several days to the ephemerides provided by the JPL HORIZONS Web interface3. Figure 2 shows the view TESS has of Sol d on each of the dates in our dataset; these images were produced by applying the sequences of rotations described above. The planet is illuminated by Sol assuming it is a point source, and the night side is shaded black. The map of Sol d shown in this figure is based on current models of the actual distribution of continents on the planet. Note that although the images are computed an integer number of rotation periods apart, the planet appears to slowly rotate over the course of the observation. This is due to the orbit of TESS, which changes its vantage point relative to Sol d over the course of its 13 day orbit. 3.2. Systematics Modeling While the astrophysical signal that we wish to analyze is present in all extracted light curves, its strength is modulated by time-variable systematics that are correlated but not identical across all postage stamps, since the illumination of the CCD by Sol d is strongly inhomogeneous. A physical model of the reflections and scattering that gives rise to the signal of Sol d on the TESS detector is beyond the scope of this work. Instead, we treat this as a de-trending problem, where the signal we wish to fit is shared by all postage stamps, but multiplied by a systematic signal that is 2 https://archive.stsci.edu/missions/tess/models/ 3 See https://ssd.jpl.nasa.gov/horizons.cgi and this Jupyter notebook. 6 Synthetic images showing TESS's view of Sol d on every date in Sectors 1 Figure 2. and 2 that was included in the regression. Each image is labeled with the corresponding timestamp in TJD. The dayside features shown here correspond to a hypothetical model of Sol d, not to the map inferred in this study. ƭ different for each target. This problem is precisely what the technique of pixel-level decorrelation (PLD; Deming et al. 2015; Luger et al. 2016, 2018) seeks to address. In PLD, one computes the sum over all N light curves at each point in time, and uses this quantity to normalize each of the individual light curves. Since (by assumption) each light curve is the product of the astrophysical signal and the systematics signal, this procedure divides out the astrophysical signal, producing a basis of N vectors that contain only functions of the systematics. This now serves as a "clean" basis one can then use to fit the systematics component of the data, with minimal risk of fitting out astrophysical signals. We use PLD to determine a systematics basis set for each TESS sector, and from this basis we construct a separate design matrix B for each sector. To increase the flexibility of the systematics model, we append to B a basis of fifth order orthogonal polynomials in time for each of the two orbits in each sector. The systematics model p for the nth signal is thus pn = Bwn (2) where wn are the weights of each regressor. Since we solve for different weights for each of the 75 light curves in Sector 1 and 20 light curves in Sector 2, our systematics model has 75 × (75 + 2 × 5) + 20 × (20 + 2 × 5) = 6, 975 free parameters. The total number of data points is 912, 605, so overfitting is not particularly concerning; nevertheless, we impose an L2 prior on the weights to minimize this risk (see §3.4). 1325.01326.01327.01328.01329.01330.01331.01332.01333.01334.01340.01341.01342.01343.01344.01345.01346.01347.01348.01349.01363.01364.01365.01366.01379.01380.01381.01382.0 7 Figure 3. Geometry of the starry model in reflected light. In this frame, the y axis points up, the x axis points to the right, and the z axis points out of the page. The planet has unit radius and sits at the origin, while the illumination source is along the y − z plane, with the sub-stellar point is marked by a dot. The semi-major axis of the terminator is unity, and the semi-minor axis of the terminator is denoted b. The night side is colored black. ƭ 3.3. Reflected Light Mapping with starry We model the astrophysical component of the signal -- the rotational modulation of Sol d in reflected light -- using the starry code package (Luger et al. 2019). starry computes phase curves and occultation light curves for bodies whose surfaces can be expressed as a sum of spherical harmonics. The algorithm works by first transforming from spherical harmonic coefficients to coefficients in the polynomial basis p, whose terms are of the form xiyjzk for (non-negative) integer i and j and k = 0 or 1. Luger et al. (2019) showed how to compute the visible flux by integrating each of the terms in p across the visible surface of the projected disk. For unocculted spheres observed in emitted light, the area of integration is the full disk, which makes the integration trivial: the flux contribution from each term in p is a constant easily computed via recursion relations involving factorials. The algorithm is thus extremely fast and very precise. However, in the present work, we are interested in reflected light, so we must account for the illumination by the star. This illumination varies smoothly across the dayside of the planet, but its derivative discontinuously changes at the terminator, beyond which the illumination is zero everywhere. This requires two modifications to the starry algorithm: we must (1) weight all integrands by the illumination function on the dayside, and (2) modify the integration limits to truncate the integrals at the day/night terminator. Let us consider a right-handed coordinate frame in which the planet has unit radius and is located at the origin, the illumination source is along the y − z plane, and the sub-stellar point is on the +y axis at x = 0 (see Figure 3). In this frame, the terminator is a segment of an ellipse, with semi-major axis a = 1 aligned with the x axis and semi-minor axis b. It is straightforward to show that the sub-stellar point is b1b2 8 Figure 4. Normalized surface reflectance versus time for the maximum likelihood starry model in each of the four orbits of TESS in Sectors 1 and 2. ƭ located at y = √ (cid:40)√ 1 − b2 and the illumination is given by y ≥ −b otherwise 1 − b2y − bz I(b; x, y) = √ 1 − x2 , k = 1 (3) where z = (cid:112)1 − x2 − y2. Since the illumination function is just a polynomial in 0 , x, y, and z, weighting the terms in p by this function keeps all terms within the polynomial basis. Provided we always rotate the problem such that the sub-stellar point lies along the +y axis as above, the limits of integration are −1 ≤ x ≤ 1 and √ 1 − x2 ≤ y ≤ √ 1 − x2. Computing the total reflected flux for any surface map is b (cid:90) 1 (cid:90) √ therefore a matter of solving integrals of the form (cid:1) Γ(cid:0) j+1 Γ(cid:0) i+1 2Γ(cid:0) i+j+k+4 (cid:1) 0 1 − bj+1 √ (cid:18) j + 4 xiyjzkdy dx Jijk(b) = (cid:19) (cid:18) (cid:19) 1−x2 j + 1 j + 3 (cid:1) √ b 1−x2 i odd k = 0 −1 2 2 = 2 − bj+1Γ π 2 F1 2 2 −1 2 , ; 2 2 ; b2  (4) F1 is the regularized hypergeometric function. For all values of i, j, and k in where 2 p, 2 F1 reduces to simple trigonometric functions involving b. Moreover, it is straight- forward to derive recursion relations for Jijk(b), making its evaluation extremely fast. We implemented the algorithm for computing Jijk(b) in the development version of starry4; the details will be discussed in more detail in upcoming work. As with the systematics model (§3.2), the model for the astrophysical signal is lin- ear. The signal s is a linear operation on the vector of spherical harmonic coefficients 4 https://github.com/rodluger/starry/tree/linear 133013401350136013701380Time [TJD]0.00.10.20.30.40.50.60.7Normalized ReflectanceS1.1S1.2S2.1S2.2 y that describes the body's surface: s = Ay 9 (5) where A is the design matrix and y is the vector of spherical harmonic coefficients. Note that in Luger et al. (2019), y describes the emissivity of the surface, but here we will use it to represent the albedo of the planet. We chose a maximum spherical harmonic degree lmax = 10 for our fits. Increasing lmax above this value leads to a sharp rise in the amount of "ringing" without substan- tially increasing the quality of the fit. Note that l = 10 corresponds to a maximum surface resolution on the order of 180◦/10 ≈ 18◦. Finally, for reference, in Figure 4 we plot the starry model for our maximum like- lihood fit (§4). The sharp rise at the beginning of each orbit in Sector 1 is due to the changing vantage point of TESS as Sol d transitions from a thin crescent to full phase. 3.4. Full Model and Likelihood Function Our full model for each light curve is simply the product of the systematics model pn = Bwn (different for each target) and the starry model s = Ay (shared by all targets). Since we happen to know the exact distance r between TESS and Sol d at all times (§3.1), we also include the inverse square of r as a multiplicative term to scale the luminosity of Sol d to the flux observed by TESS. The model for the flux time series in the nth postage stamp is therefore mn = (r−2) ◦ (Bwn) ◦ (Ay) where ◦ denotes the element-wise product of two vectors. (6) We solve Equation (6) for the weights wn and y by maximizing the negative log likelihood function in two separate steps. In the first step, we take advantage of the linearity of the problem to perform a fast, semi-analytical optimization to obtain a starting guess for the weights. In the second step, we apply additional constraints to prevent overfitting and ensure an albedo in the range [0, 1] everywhere on the surface. In the first step, we initialize the starry model to unity and iteratively solve for wn and y by solving the L2-regularized least squares problem (see, e.g., §2.1 in Luger et al. 2018). We assume zero-mean Gaussian priors for the weights: y ∼ N (0, σy(l)2) , wn ∼ N (0, σ2 w) (7) with σw = 0.05 and σy(l) = (cid:40)1.75 × 10−5 l 3 2 1.40 × 10−4 l− 3 2 where l is the spherical harmonic degree. The latter prior was chosen by trial-and- error and was adjusted to keep the albedo positive everywhere on the surface. Note, l < 2 l ≥ 2 , (8) 10 importantly, that we do not fit for the coefficient of the constant Y0,0 spherical har- monic, but instead fix it to unity. Since the solutions obtained in this step are merely used as a starting point for the second step (see below), the choice of prior at this stage does not significantly impact on our results. In general, we find that the iterative scheme converges within about 50 iterations, taking about two minutes on a laptop computer. However, since it is strongly regular- ized, the model significantly underfits the data. In the second step, we remove our L2 prior on the spherical harmonic coefficients and instead regularize the actual surface albedo, which we compute on a uniform spherical grid with 50,000 points. We enforce a uniform prior in the range [0, 1], with a small amount of Gaussian smoothing on ei- ther side. We additionally impose a constraint on the power spectrum of the inferred map, requiring that it be drawn from a distribution whose mean is a decaying power law in the spherical harmonic degree l in order to suppress ringing and spurious high order features. Our full negative log likelihood function in this step is therefore − log L = 1 2 1 2 1 2 1 2 1 2 w w + (f − m)(cid:62)Σ−1(f − m) + w(cid:62)Λ−1 (a− + 1 − a+)(cid:62)Λ−1 (ρ − l−γ)(cid:62)Λ−1 (γ − γ0)2 ρ (ρ − l−γ) + a (a− + 1 − a+) + , σ2 γ (9) where f is the measured flux across all targets, m is the full model (Equation 6), Σ is the (diagonal) data covariance, w are the systematics model weights for all targets, Λw is the prior covariance of the weights (Equation 7), a− and a+ are vectors containing the values of the albedo that are below zero and above one, respectively, Λa is the covariance of the Gaussian used to smooth the edges of the top-hat prior on a, ρ is the power spectrum of the inferred spherical harmonic map (equal to the sum of the squares of the coefficients at each degree l), γ is the power law index of the power spectrum prior, Λρ is the covariance of that prior, γ0 = 1.5 is the mean of γ = 10−2 is the corresponding variance. For simplicity, all our the prior on γ and σ2 prior covariances are diagonal and homoscedastic with variances equal to 2.5 × 10−3 for w, 10−8 for a, and 10−1 for ρ. Unfortunately, the constraints outlined above break the linearity of the problem, and we must turn to a non-linear optimizer. We use the AdamOptimizer method in the TensorFlow package (Abadi et al. 2015) to maximize Equation (9). Starting with the values of the weights obtained in the first step, we run the AdamOptimizer for 1000 iterations, well past convergence. Since TensorFlow automatically computes and propagates gradient information, this step is fast, finishing in under one minute on a laptop. 11 Figure 5. Maximum likelihood map of Sol d obtained during the first stage of our optimization procedure, in which an iterative analytic approach was used to solve the bilinear model, subject to very conservative regularization. Despite the low resolution of this map, the bright features appear to track land masses which for specificity we will refer to as (from left to right) "South America," "Africa," and "Asia." The two dark features are associated with the planet's two purported oceans, which we call the "Pacific" and the "Atlantic." Black contours correspond to supposed coastlines derived from current state-of-the-art models of the surface geography of Sol d. ƭ 4. RESULTS Figure 5 shows the maximum likelihood map inferred during the first stage of our optimization procedure. Since the map coefficients were heavily regularized, the dynamic range of the albedo variations was less than 10%; we scaled it to span the range [0, 1] for better visualization. Furthermore, although the maximum resolution of the spherical harmonic model is on the order of 18◦, features typically span ∼45◦; this is also an artefact of our regularization. However, broad features are already visible in the map: three bright regions and two dark regions (one of which wraps around the antimeridian). Overplotted on the inferred map are coastal outlines corresponding to our current best understanding of the geography of Sol d. One can see that the three bright regions roughly track, from left to right, continents we shall refer to as "South America," "Africa," and "Asia," in keeping with the literature. The dark regions are associated with the two major purported oceans on the planet, the "Pacific" (centered at longitude ±180◦) and "Atlantic" (centered at longitude 45◦). As expected, the second, non-linear step in our optimization procedure yields a much better fit to the data. The top panel of Figure 6 shows the median of the max- imum likelihood model in this second step across all 95 targets (blue) overplotted on the median flux (black). The median residuals are shown in the bottom panel; these have standard deviation of less than one percent. While there is significant correlated structure in the residuals, their power spectrum displays a broad peak extending be- tween about 0.1 and 10 days, with a significant dip at 1.00 day. This suggests our -180-135-90-4504590135180Longitude-90-4504590LatitudeInferred0.00.20.40.60.81.0Normalized Albedo 12 Top: Median flux across all postage stamps (black) and the median maxi- Figure 6. mum likelihood model (blue) across Sectors 1 and 2. Bottom: Median residuals about the maximum likelihood fit. ƭ model is fitting the rotational variability of Sol d quite well, but additional tempo- ral variations -- most likely due to cloud movement -- cause aperiodic changes in the reflectivity of the surface. We revisit this point in §5. In the top panel of Figure 7 we show the final, maximum likelihood albedo map inferred for Sol d weighted by a visibility function. The visibility function is computed as the average of the product of the cosine of the illumination angle and the cosine of the angle between the observer vector and the vector normal to the surface of Sol d. This normalization has the effect of downweighting regions that are either seldom illuminated (such as regions near the southern pole of Sol d, as the observations were taken during southern winter) or seldom in view (such as regions near the northern pole of the planet, which during Sectors 1 and 2 were mostly on the opposite side of Sol d from TESS). Because there is little data from these regions, their albedo is prior- dominated and tends to fluctuate substantially with small changes to the choices we make when fitting the data. On the other hand, we find that the large scale features close to the equator of Sol d are mostly insensitive to our choices of prior. In the Figure, we use dashed lines to delineate the regions whose average visibility is larger than 0.5. The bottom panel of Figure 7 shows an approximate average cloud cover map of Sol d based on imagery taken by the Visible Infrared Imaging Radiometer Suite (VIIRS) instrument aboard the Suomi National Polar-orbiting Partnership (S-NPP) satellite5. The image was produced by analyzing the corrected true color reflectance in the I1 (red), M4 (blue), and M3 (green) bands and masking all pixels that were either dark or had significant variance among the three bands (i.e., inconsistent with a grey spectrum). We then averaged the images averaged over the 28 days of data 5 Data obtained via this wget script. 0.00.51.01.5Normalized FluxS1.1S1.2S2.1S2.2Median FluxMedian Model133013401350136013701380Time [TJD]-0.0250.0000.025Residuals 13 Figure 7. Top: The maximum likelihood model for the surface albedo of Sol d, weighted by the visibility across Sectors 1 and 2. The dashed white lines indicate latitudes above/below which the visibility drops below 50%. As in Figure 5, coastlines based on current models from the literature are overplotted as black contours. Bottom: Approximate cloud coverage map based on corrected surface reflectance obtained by the VIIRS imager, averaged over the same date range and weighted by the same visibility function. ƭ -180-135-90-4504590135180Longitude-90-4504590LatitudeLow visibilityLow visibilityInferred-180-135-90-4504590135180Longitude-90-4504590LatitudeLow visibilityLow visibilityTrue0.00.20.40.60.81.0Visibility-Weighted Albedo0.00.20.40.60.81.0Visibility-Weighted Albedo 14 analyzed in this work, smoothed it with a Gaussian filter, and weighted by the TESS visibility as in the top panel. While the two panels display fundamentally different quantities -- an albedo and a cloudiness factor -- we expect that the signal of Sol d in the TESS bandpass is domi- nated by cloud reflectivity, as water clouds have an albedo higher than soil, vegetation, or ocean in the optical/near infrared (e.g., Jedlovec 2009). We therefore expect large, coherent cloud structures to show up as the brightest regions in our inferred map, and in fact we find that this is generally the case. The dominant feature in our in- ferred map is a bright region spanning longitudes of +50◦ to +100◦ just northward of the equator. This is also the dominant feature in the VIIRS map, corresponding to the persistent weather system associated with summertime monsoons in the south of Asia. Secondary features common to both maps are a permanent cloudy region over central Africa with a westward spur over the Atlantic Ocean, a veneer of clouds over the north Pacific, and a pile up of clouds on the western coast of northern South America. The dark regions in the inferred map rather loosely track the Pacific, Atlantic, and Indian Oceans, although the correspondence is not exact. 5. DISCUSSION 5.1. An Ill-Posed Problem The problem of inferring a two-dimensional map from a one-dimensional time series is famously ill-posed given its large null space and many complex degeneracies (e.g., Cowan et al. 2013). In fact, for a uniformly illuminated body, its rotational light curve encodes at most 2lmax modes (one sine term and one cosine term per frequency), where lmax is the highest spherical harmonic degree of features on the surface of the body. The number of modes on the surface of the body, however, increases quadratically as (lmax + 1)2. This means that for a body whose surface is perfectly described by a spherical harmonic map of degree lmax = 10, which is described by 121 coefficients, only 20 linear combinations of those coefficients are actually projected onto the light curve. The problem is fundamentally non-invertible, since most of the information one wishes to extract never makes it into the light curve in the first place. Fortunately, there are two aspects of the problem at hand that facilitate this (seem- ingly impossible) inversion. First, the surface of Sol d is not uniformly illuminated; instead, different parts of the surface are weighted by different amounts as the planet rotates and as the angle of incident starlight moves over the course of the planet's year. This weighting is sufficient to break many of the degeneracies, which are rooted in the anti-symmetry of many of the spherical harmonic modes. Moreover, the discon- tinuous day/night terminator further breaks these degeneracies, as only modes that are anti-symmetric about the visible lune of the planet strictly remain in the null space. Second, the changing vantage point of TESS relative to Sol d (and our exact knowledge of it) breaks degeneracies that are viewpoint-dependent. In particular, the 15 motion of TESS changes the position and shape of the terminator on the projected disk of Sol d, enhancing the effect discussed above. In fact, because of these points, there should actually be no null space in this particular problem. In practice, however, it is evident from Figure 7 that degeneracies still abound. While we infer a few of the dominant features of the cloud map of Sol d, several of them are skewed, shifted, or entirely out of place. This is the case, for instance, with the cloud bank west of South America, which extends too far westward, and the northern Atlantic Ocean, which appears to extend into the east coast of North America. We identify four broad reasons for this. First, many of the issues arise because of the finite signal-to-noise of our measure- ments and various biases in our adopted model. While the measurement errors on our data points are quite small -- since the signal of Sol d is both quite strong and independently measured at 95 different positions on the detector -- our lack of under- standing of the complex optics (and our attempt to naïvely model it with pixel-level decorrelation) likely introduces significant correlated noise when we de-trend. We discuss this and other sources of bias in more detail in §5.3 below. Second, while we attempted to adopt agnostic priors on the map coefficients -- i.e., we fit for a spectral slope rather than impose one -- our inferred map is still heavily prior-dependent. Our choice of regularization strength for the systematics was based on trial-and-error and what "looked good," as was the choice for the hyperprior on the spectral slope and the maximum spherical harmonic degree of the map. Although the main features of the map (the monsoon over south Asia, the clouds over Africa and South America, and dark regions associated with the three major oceans) persist regardless of the choice of prior, their exact shapes and contrast vary significantly under different assumptions. Third, inspection of Figure 2 shows that most of the time, Sol d is close to full phase as seen by TESS. With the terminator close to the limb of the planet, its ability to break degeneracies is considerably lessened. We are therefore somewhat limited in our ability to infer in particular the latitudinal position of features. Observations of Sol d in upcoming sectors may contain longer segments in which Sol d is seen in crescent, which should help mitigate these degeneracies. Fourth, and most importantly, we have not performed actual inference: that is, we simply found the maximum likelihood solution (subject to our constraints), with no es- timate of the uncertainty anywhere on our map. A much better approach than simply maximizing the likelihood function is to marginalize over the nuisance parameters -- such as the systematic model weights and the spectral slope. This is more costly, but certainly not intractable, and we will pursue this in future work. 5.2. Confirmation Bias It is difficult to avoid being disingenuous in one's solution when the answer (or the expected answer) to a problem is known ahead of time. Since we had access to 16 "ground truth" from the beginning, many of the de-trending/processing/fitting choices we made were (however inadvertently) driven by the desire to come as close as possible to the correct answer. While many of these choices correspond to numbers we have explicitly quoted here, such as the prior variances on our model terms, several of them are implicit decisions on how we performed the analysis. For instance, in an early version of the project we used as regressors actual TESS housekeeping variables -- such as the angle and elevation of the Earth and the Moon relative to each of the four camera boresights -- instead of the PLD basis. However, the systematics model constructed from these quantities proved to be too inflexible, judging in part from the poor fidelity of the recovered maps. The results in this work must therefore be taken with a grain of salt. That said, we have attempted to offset the possibility of confirmation bias by making our model (Equation 6) as simple as possible in the end (a multiplicative baseline and a linear reflectance model) and thus limiting the number of knobs we could turn. We also opted to forego any use of real imaging data in our model; for instance, we fit for the spectral slope of the map as a free parameter rather than using the true power spectrum of Sol d. Moreover, the broad features in the strongly regularized map (Figure 5) -- and in particular the monsoon clouds over southeast Asia -- are insensitive to most of our model choices. 5.3. Other Kinds of Bias One of the largest sources of bias is the fact that we assumed our systematics model is strictly multiplicative. Although we subtracted an estimate of the baseline non-Sol d background flux from each target (measured from the mid-sector portions of the light curve where the planet was below the edge of the sunshade), we implicitly assumed that this constant baseline was the only source of additive noise. We know this to be strictly untrue, as Sol d is known to have a companion (Sol d I), a moon about one-third the diameter of Sol d, whose reflected light should cause periodic modulations in the background flux of TESS pixels. Because of its small size and very low (∼ 0.1) albedo, we chose to ignore its effect, but it is likely to contaminate our signal at certain points in the orbit where TESS is closer to Sol d I than to Sol d. While this will likely bias the starry model at those times, the effect is particularly concerning because of its impact on the PLD basis. While PLD can be an extremely powerful de-trending technique, it is notorious for grossly overfitting signals in the presence of uncorrected background modulations (Luger et al. 2016). This happens because PLD strictly assumes the noise is multiplicative, which is not the case for signals with varying (additive) background levels. In principle, one could model the reflectance variability of Sol d I in the same way as we did here, but since the optics are not well understood, this may be difficult in practice. In our analysis, we also neglected the effects of limb darkening. Because of atmo- spheric attenuation, features near the limb of the planet should in reality contribute less to the flux than our model predicts. This could bias our results particularly in 17 cases where the planet is seen at half phase (for instance, during days 1326 and 1341 in Figure 2). In the absence of limb darkening, the brightest regions on the planet are those close to the sub-stellar point, which is on the limb. Limb darkening should shift the regions of peak brightness closer to the center of the planet, adding an east-west bias in the location of the features we identify. We also ignored the finite angular size of Sol at Sol d (about 0.5◦), treating it instead as a point source. Regions immediately nightward of the terminator should therefore contribute a small amount of flux, which we neglect. Similarly, we used the timestamps reported in the TESS light curve FITS files, which are corrected to the barycenter of the Sol system. Since our signal originates from a source that is closer to TESS than the barycenter of the system, the timestamps may be wrong by up to 8 minutes, the light travel time from TESS to Sol. Over the course of 8 minutes, Sol d rotates by about 2◦, so there should be a blurring of our inferred map at this scale. However, these two effects are well below the resolution of our map, so we expect them not to introduce significant bias. Moreover, we implicitly assumed an isotropic prior on the spherical harmonic co- efficients by imposing a power spectrum in only the degree l. This has the effect of lessening the extent to which features can be confined to certain latitudes. In par- ticular, our prior implicitly disfavors banded cloud structure, which one may a priori expect on a rapidly rotating planet such as Sol d. The bottom panel of Figure 7 shows such banded structure at the equator and at mid-latitudes due to the combined effects of north-south circulation and deflection by the Coriolis force. Finally, and perhaps more importantly, our model assumed a static map for the surface of Sol d. We also know this to be incorrect, since clouds form and move on timescales of hours to days. We discuss this in detail in the following section. 5.4. Temporal Evolution As we mentioned in §4, the residuals about our maximum likelihood fit show correlated structure with power at a large range of periods ranging from 0.1 up to 10 days. This is likely due to the fact that rotational variability is not the sole source of the signal of Sol d: additionally, there is temporal variation in the albedo of certain parts of the map, most probably due to cloud movement. In order to test this, we used starry to construct a time-variable spherical harmonic map, defined by a vector of spherical harmonic coefficients that varies slowly over time: N(cid:88) y(t) = yncn(t) (10) n=0 where the yn are the N spherical harmonic coefficient vectors that describe the tem- poral variability and the cn are the components of the basis in time. Since this is a linear operation on the spherical harmonic coefficients, the starry model is still linear, and may be solved in the same way as before. 18 We choose the cn vectors to be an orthogonal polynomial basis in time of degree 6 spanning all of Sector 1. Under this basis, one can interpret y0 as the static component of the map, y1 as the component that varies (quasi-)linearly in time, y2 as the component that varies (quasi-)quadratically, and so forth. In all, we fit for six of these map components. We neglect the data taken in Sector 2, focusing instead on the portion of the data for which we have good temporal coverage. In order to mitigate overfitting due to the larger number of variables we are solving for, we add a term to our likelihood function that forces the coefficients of the n > 0 maps to remain small. We further evaluate the albedo of the map on a grid at six equally spaced points in time, enforcing our albedo and power spectrum priors separately at each of these points. The mean of the temporal solution is shown in Figure 8, and a link to an animated version is provided in the caption. No significant differences are evident, although the residuals improved by ∼30% and contain significantly less power on timescales longer than 1 day. We conclude that while this is evidence for cloud variability on these timescales, a more detailed investigation is warranted to say anything meaningful about the dynamics of clouds on Sol d. 5.5. Implication for Future Observations As we discussed above, light curve inversion is a hard problem, and we have shown that this is true even at high signal-to-noise and when the solution is known ahead of time. Presented with the top panel of Figure 7 (without the continent outlines), an inhabitant of Sol d may be hard-pressed to identify that as a map of their home planet. This is in part due to the biases in our optimization, which we discussed above, but also due to the simple fact that the reflectance of Sol d is dominated -- by far -- by time-variable clouds. While stationary cloud features on Sol d are typically associated with continents (the most pronounced cloud features being over South America, Africa, and south Asia), persistent cloud cover can occur over open ocean, as seen in the north Pacific in Figure 7. Given a baseline of just a few weeks, it is not possible to tell the difference between highly reflective land features (such as ice or vegetation) and stationary clouds. We should not expect other exoplanets in the habitable zone to be any different. Future missions such as LUVOIR or HabEx may enable us to collect data on the phase curves of such planets, but that data will be of significantly lower quality and poorer time resolution, and we will not know the planet's rotation period or obliquity ahead of time. The work presented here can therefore be seen as an upper limit on what we may be able to infer about exoplanets in the habitable zone with next-generation facilities. That said, the time baseline considered in this work was fairly limited -- three weeks in Sector 1 and one week in Sector 2, spanning two months in total. Observations taken throughout an entire year on Sol d -- during which time cloud patterns should 19 Figure 8. The inferred mean map when allowing for slight temporal variations in the spherical harmonic coefficients. Compare to Figure 7. An animated version of this figure is available on GitHub. ƭ -180-135-90-4504590135180Longitude-90-4504590LatitudeLow visibilityLow visibilityInferred-180-135-90-4504590135180Longitude-90-4504590LatitudeLow visibilityLow visibilityTrue0.00.20.40.60.81.0Weighted Albedo0.00.20.40.60.81.0Weighted Albedo 20 Figure 9. Artist's conception of Sol d. Credit: Ian Peng-Sue, age 7. change with the seasons and the angle between the day/night terminator and the axis of rotation should precess -- may help to disambiguate between clouds, oceans, and continents. Although the signal of Sol d mostly disappears from the TESS detector after Sector 2, it is expected to return in Sectors 14 and 15 when the planet will again be above the edge of the sunshade. We intend to perform this analysis in upcoming work once data from future TESS sectors becomes available. Additionally, time-resolved observations in multiple bands could also help with this problem. This was discussed in Cowan et al. (2009), who showed that multi-band photometry of the Earth enables one to disentangle the contributions from oceans, continents, and cloud cover. A robust detection of surface features on an exoplanet in the habitable zone will likely necessitate both approaches alongside a circumspect analysis of the uncertainties and degeneracies at play. 6. CONCLUSIONS We have used TESS photometry to produce a global map of Sol d, a terrestrial planet in the habitable zone of a G2 main-sequence star. We adapted the starry code to generate (and invert) light curves in reflected light, coupling it to a variation of pixel level decorrelation to model the systematics in the TESS detector. We detect three large scale bright regions near the equator of the planet, associated with persistent cloud structures at longitudes of −60◦, +30◦, and +90◦, which likely track continents. Pronounced dark features are identified at longitudes of ±180◦, −30◦, and +60◦ with significant latitudinal extent, likely corresponding to oceans. Because of the vantage point of TESS during Sectors 1 and 2, we are not sensitive to high latitudes on the planet, although this may change with upcoming sectors. We fit Sol d with both static and time-variable albedo maps. We find tentative evidence for variability in Sol d's cloud coverage over the course of Sector 1 observa- tions. Upcoming observations with TESS will yield further information on the time variability of surface features on Sol d on a variety of timescales. 21 To summarize, we have robustly detected surface features from TESS light curves of Sol d. While the resolution of the map we obtain is limited, it is clear that surface features including low-albedo oceanic regions and high-albedo cloud banks or land masses are present on this terrestrial world. An artist's impression of the planet is shown in Figure 9. While somewhat fanciful and rather far-fetched, this artistic interpretation is nevertheless broadly consistent with the data. The likely prevalence of continents, oceans, and complex weather patterns on Sol d make it an extremely promising site for the evolution of life. We strongly advocate for future missions to further investigate this possibility. In particular, Sol d may be a good choice of target for a lander mission. One prime landing site is located at the planetary (latitude, longitude) coordinates (33.3943◦ N, 104.5230◦ W).6 The work presented here also serves as a relevant case study for future efforts to characterize terrestrial exoplanets. While the circumstances of TESS's observations of Sol d will not be replicated exactly for other planets, similar rotationally-modulated phase curves of reflected light may be feasible in the future. We show that a fast, analytic linear model can be successfully applied to such data. We also underscore the importance of high cadence and long duration for these observations, and we advocate for multi-wavelength coverage to help disentangle the photometric signatures of continents and persistent cloud features. In conclusion, we have demonstrated that TESS is an effective addition to NASA's fleet of weather satellites. Our heartfelt thanks go to Ian Peng-Sue and his agent Stephanie Tonnesen for Fig- ure 9. We also gratefully acknowledge Jonathan Fraine, Dan Foreman-Mackey, David W. Hogg, and Ben Pope for useful conversations. Crucial parts of this work were car- ried out at the TESS Data Workshop, hosted by Space Telescope Science Institute, and the Building Early Science with TESS Workshop, hosted by the University of Chicago. Facility: TESS Software: Astropy, Matplotlib, Numpy, starry (Luger et al. 2019), TensorFlow (Abadi et al. 2015), spiceypy (Annex et al. 2019) 6 We considered the alternative landing site of southern Oceania, but rumors of murderous native fauna including the fearsome "drop bear" (Andrew R. Casey, personal communication) led us to conclude that this would not be a suitable location. 22 REFERENCES Abadi, M., et al. 2015, TensorFlow: Large-Scale Machine Learning on Heterogeneous Systems, v1.13.1, tensorflow.org. http://tensorflow.org/ Acton, C. H. 1996, Planet. Space Sci., 44, Haggard, H. M., & Cowan, N. B. 2018, MNRAS, 478, 371 Hammond, M., & Pierrehumbert, R. T. 2017, ApJ, 849, 152 Jedlovec, G. 2009, Automated Detection of Clouds in Satellite Imagery. https: //cdn.intechopen.com/pdfs/9541.pdf Jiang, J. H., et al. 2018, AJ, 156, 26 Kawahara, H., & Fujii, Y. 2010, ApJ, 720, 1333 Knutson, H. A., et al. 2007, Nature, 447, 183 Luger, R., et al. 2019, AJ, 157, 64 -- . 2016, AJ, 152, 100 -- . 2018, AJ, 156, 99 Lustig-Yaeger, J., et al. 2018, AJ, 156, 301 Majeau, C., et al. 2012, ApJL, 747, L20 Ricker, G. R., et al. 2015, Journal of Astronomical Telescopes, Instruments, and Systems, 1, 014003 Sagan, C., et al. 1993, Nature, 365, 715 Urban, S. E., & Seidelmann, P. K. 2013, Explanatory Supplement to the Astronomical Almanac, 3rd ed. (University Science Books) 65 Acton, C. H., et al. 2017, Planet. Space Sci., 150, 9 Annex, A. 2017, in Third Planetary Data Workshop and The Planetary Geologic Mappers Annual Meeting, Vol. 1986, 7081 Annex, A., et al. 2019, AndrewAnnex/SpiceyPy, v2.2.0, Zenodo, doi:10.5281/zenodo.2576445 Berdyugina, S. V., & Kuhn, J. R. 2017, arXiv e-prints, arXiv:1711.00185 Cowan, N. B., et al. 2013, MNRAS, 434, 2465 Cowan, N. B., & Fujii, Y. 2018, Mapping Exoplanets, 147 Cowan, N. B., et al. 2009, ApJ, 700, 915 de Wit, J., et al. 2012, A&A, 548, A128 Deming, D., et al. 2015, ApJ, 805, 132 Demory, B.-O., et al. 2016a, MNRAS, -- . 2016b, Nature, 532, 207 Fujii, Y., & Kawahara, H. 2012, ApJ, 755, 455, 2018 101
1810.06580
1
1810
2018-10-15T18:00:16
Science with an ngVLA: Resolving the Radio Complexity of EXor and FUor-type Systems with the ngVLA
[ "astro-ph.EP", "astro-ph.SR" ]
Episodic accretion may be a common occurrence in the evolution of young pre-main sequence stars and has important implications for our understanding of star and planet formation. Many fundamental aspects of what drives the accretion physics, however, are still unknown. The ngVLA will be a key tool in understanding the nature of these events. The high spatial resolution, broad spectral coverage, and unprecedented sensitivity will allow for the detailed analysis of outburst systems. The proposed frequency range of the ngVLA allows for observations of the gas, dust, and non-thermal emission from the star and disk.
astro-ph.EP
astro-ph
Resolving the Radio Complexity of EXor and FUor-type Systems with the ngVLA Jacob Aaron White1, Marc Audard2, Péter Ábrahám1, Lucas Cieza3, Fernando Cruz-Sáenz de Miera1, Michael M. Dunham4, Joel D. Green5, Manuel Güdel6, Nicolas Grosso7, Antonio Hales8, Lee Hartmann9, Kundan Kadam1, Joel H. Kastner10, Ágnes Kóspál1, Sebastian Perez11, Andreas Postel2, Dary Ruiz-Rodriguez10, Christian Rab12, Eduard I. Vorobyov13,14, & Zhaohuan Zhu15 1Konkoly Observatory, Hungarian Academy of Sciences, Budapest, Hungary 2Department of Astronomy, University of Geneva, Versoix, Switzerland 3Facultad de Ingeniería y Ciencias, Univ. Diego Portales, Santiago, Chile 4Dept. of Physics, State University of New York at Fredonia, Fredonia, USA 5Space Telescope Science Institute, Baltimore, USA 6Department of Astrophysics, University of Vienna, Vienna, Austria 7Aix Marseille Univ, CNRS, CNES, LAM, Marseille, France 8Joint ALMA Observatory, Santiago, Chile 9Dept. of Astronomy, University of Michigan, Ann Arbor, USA 10Center for Imaging Science, Rochester Inst. of Technology, Rochester, USA 11Departamento de Astronomía, Universidad de Chile, Santiago, Chile 12Kapteyn Astronomical Inst., Univ. of Groningen, Groningen, NL 13University of Vienna, Department of Astrophysics, Vienna, Austria 14Research Inst. of Physics, Southern Federal Univ., Roston-on-Don, Russia 15 Dept. of Physics and Astronomy, University of Nevada, Las Vegas, USA Abstract. Episodic accretion may be a common occurrence in the evolution of young pre- main sequence stars and has important implications for our understanding of star and planet formation. Many fundamental aspects of what drives the accretion physics, how- ever, are still unknown. The ngVLA will be a key tool in understanding the nature of these events. The high spatial resolution, broad spectral coverage, and unprecedented sensitivity will allow for detailed analysis of outburst systems. The proposed frequency range of the ngVLA allows for observations of the gas, dust, and non-thermal emission from the star and disk. 1 2 White et al. 1. Young Outburst Systems Understanding the underlying mechanisms that cause the observed outbursts of young stellar objects (YSOs) is important not only for building a complete picture of stel- lar formation, but also for the potential implications on the planet formation process. These outbursts offer a potential solution to the luminosity problem, i.e., the discrep- ancy between the mass accretion rates inferred from protostellar luminosities (10−8 M(cid:12) yr−1) and the accretion rates predicted by cloud core collapse models (10−6 M(cid:12) yr−1). The observed outbursts may be due to episodic accretion of disk material onto the star, which consists of a sudden increase in the accretion rate from 10−8 M(cid:12) yr−1 in quies- cence up to as high as 10−4 M(cid:12) yr−1 during an outburst. The relatively short ∼ 1 − 100 yr cumulative duration of the outbursts, compared to the quiescent phase, provides a potential solution to the luminosity problem (e.g., Dunham & Vorobyov 2012). The study of these outburst phenomena began with FU Orionis (Herbig 1966), which in 1937 showed an increase of 5 magnitudes in the V band and has since been slowly fading. Outbursts are now broadly categorized into two groups: FUors and EXors, named after their prototypes FU Orionis and EX Lupi, respectively (see Audard et al. 2014, for a recent review of these types of objects). FUors undergo large outbursts which can last 100s of years, while EXors exhibit smaller and shorter lived outbursts (∼ 1 yr). Indeed, there are objects that do not fit clearly into either classification (FUor- like objects) as they can show similar spectroscopic characteristics as a FUor without showing other evidence of an outburst. This leads to the possibility that EXors and FUor represent different stages of evolution of the same object. Cieza et al. (2018) showed that EXors, for example, have lower disk masses than FUors, indicating they may represent a later stage of disk evolution. Although over the past 50 years EX/FUors have been observed with broad spectral coverage and modeled with increasing high levels of detail, there are still many open questions on their formation, frequency of occurrence, and impact on circumstellar environments. What are the physical processes that cause these types of events? Despite being thoroughly studied, the triggering mechanisms of these outbursts are still not known. It is also not clear if the same mechanisms are responsible for both EXor and FUor-type outbursts. The ngVLA will be instrumental in addressing this issue by: 1. Accurately measuring disk masses. Knowing disk masses is crucial for under- standing the FUor phenomenon, as gravitational instability may play an impor- tant role in triggering of the outbursts (e.g., Vorobyov & Basu 2005; Bae et al. 2014). The dust disks are optically thick and have a small angular size, which complicates mass estimates (e.g. Liu et al. 2017; Hales et al. 2018). However, the unprecedented angular resolution of the ngVLA, in combination with its spectral coverage at longer wavelengths where the dust can be optically thin, will enable more precise dust mass estimates. In addition, higher resolution observations of cold gas will provide further constraints on the total disk mass. 2. Dynamically constraining stellar masses. As the luminosity of EX/FUor systems can be dominated by the disk, it can be difficult to accurately determine the mass of the protostar (e.g., V883 Ori has an expected photospheric luminosity of ∼ 6 L(cid:12) and a bolometric luminosity of ∼ 400 L(cid:12); Cieza et al. 2016). Moreover, the inferred disk inclination affects the inferred disk luminosity and accretion rate. High spatial resolution spectroscopy with the ngVLA will allow for dynamical ngVLA Observations of EX/FUors 3 constraints on both the stellar mass and inclination through the analysis of the Keplerian rotation of the gas disk (e.g., Czekala et al. 2015). 3. Determining the strength of turbulence in the inner disk. High disk turbulence has been suggested as the reason for the loss of angular momentum needed to trigger an enhanced accretion event. Flaherty et al. (2017) analyzed the DCO+ in HD 163296's outer disk and found little turbulence, going against the theoretical predictions. This could be due to HD 163296's more evolved state or a dimin- ished role of turbulence-driven accretion. The ngVLA will help in understanding the role of turbulence in the innermost parts of disks, where the burst is expected to be triggered, and how it can cause eruptive events. 4. Observing planet-disk-star interactions. The mass exchange between a young protoplanet and the star in the innermost parts of the disk may trigger accretion outbursts (Nayakshin & Lodato 2012). Ricci et al. (2018) estimated that the ngVLA will be able to characterize gaps and asymmetries caused by a super- Earth planet at close radii (<5 au). At those distances, it will be possible to see changes in the disk material distribution in a few years, thus making important mass transfer rate predictions in the innermost regions of the disk. 5. Assessing the impact of stellar encounters and multiplicity. Close stellar en- counters either with an external star or a companion may trigger accretion out- bursts (e.g., Bonnell & Bastien 1992; Pfalzner 2008). The spatial arrangement of the outburst star and the inferred intruder is sometimes not consistent with the timescales of the outburst (Liu et al. 2017), suggesting another companion in the inner unresolved disk may be present. The ngVLA will search for close-orbit companions and assess the role of companions/intruders as outburst triggers. What are the implications of these outburst events on the origin of our Solar System and for other planetary systems? The location of snow-lines (or ice-lines) in a planet forming disk is important as an evolving snow-line can greatly impact the abundance of solid material throughout the disk and affect grain growth. ALMA obser- vations of V883 Ori show the water snow-line at a radial distance of 42 au, in contrast with the expected distance of <5 au for a solar-like star (Cieza et al. 2016). Therefore, if these eruptive events are a common occurrence in the evolution of protoplanetary disks then their affects on the relocation of snow-lines, and how this can spur grain growth, must be considered in planet formation models. In this chapter we explore how the proposed next generation Very Large Array (ngVLA) will be an invaluable tool for studying FUor and EXor type protostars. In Sec. 2, we highlight key areas that still remain largely unexplored or unconstrained with current capabilities. In Sec. 3, we show the uniqueness of the ngVLA's capabilities to study EX/FUors. In Sec. 4, we highlight synergies at other wavelengths. 2. Exploring Accretion Mechanisms with the ngVLA Current millimeter and radio observatories do provide valuable insight into EX/FUor systems (e.g., Kóspál et al. 2017; Liu et al. 2017; Cieza et al. 2018; Hales et al. 2018), but the major limitation in these studies is typically the resolving capabilities of the observatory (e.g., ∼ 20 mas or ∼ 8 au at the Orion Nebula). When the circumstellar 4 White et al. disk around a given EX/FUor system is unresolved, or only marginally resolved, it can be difficult (if not impossible) to test many of the competing models for accretion-driven outbursts. The ngVLA will provide a major improvement in the resolving capabilities of current facilities that operate in the 1 − 116 GHz frequency range, where the inner disk may be transitioning to an optically thin regime. There will indeed be many properties of EX/FUors, and circumstellar environ- ments in general, that the ngVLA will be able to study. Some of the key areas are highlighted below. 2.1. Disk Morphology The proposed 1000 km maximum baselines of the ngVLA will yield a resolution of 0.7 mas at 93 GHz (Selina et al. 2018). This resolution corresponds to spatial scales of 0.25 and 0.10 au for the two prototypical outburst systems, FU Orionis and EX Lupi, respectively. For EX Lupi, this spatial scale is slightly smaller than the ∼ 0.12 au corotation region (Sicilia-Aguilar et al. 2015) and smaller than the ∼ 0.4 au inner dust- free zone (Ábrahám et al. 2009). The proposed largest recoverable scales of >20 arcsec (Selina et al. 2018) will correspond to spatial scales of 1000s of au for FU Orionis and EX Lupi, allowing for all material with sufficient surface brightness to be detectable in a single pointing. The high level of spatial resolution will enable a detailed analysis of the morphological structure of the material in the disks. Many of the proposed drivers of accretion in EX/FUor-type objects will have dis- tinct morphological features that can be distinguished with the ngVLA. These features are caused by mechanisms such as gravitational instabilities, disk fragmentation, cir- cumbinary disk interactions, and funnel flows (e.g., Armitage et al. 2001; Zhu et al. 2009; Hartmann et al. 2016). These features may be present both in the inner optically thick and the outer optically thin parts of the disks. However, current resolving capabil- ities make it difficult to confirm or reject the presence of any substructure in EX Lupi and FU Orionis (Liu et al. 2017; Hales et al. 2018). The mm-cm spectral index of an EX/FUor-type disk will yield insight into the opacity and size distribution of the grains. An optically thick inner disk may be com- mon, but the dust opacity may transition to an optically thin regime at the lower fre- quency range of the ngVLA. Indeed, it is likely that non-thermal emission will be sig- nificant at lower frequencies, but the high resolution capabilities should enable the ther- mal and non-thermal emission components to be spatially separated. Constraints on the size distribution (q) of solids can be made when a disk is optically thin. By measuring q, the grain growth, total dust mass, and mass distribution can be much more accurately constrained than is possible with current observations. The resolution and sensitivity of the ngVLA will allow for the first tests of time- dependent evolution of disk structure. There are many mechanisms that could cause significant changes to the morphological structure of the disk on timescales that can be probed with the ngVLA. For example, if a close-in companion were present, it could open a gap in the disk or cause significant spiral structure to form (e.g. Paardekooper & Mellem 2004; Dong et al. 2015; Ricci et al. 2018). This disturbance to the disk could cause an infall of material which could further drive accretion. The movement of this material could potentially be traced over a few years. Changes over much shorter timescales would also be observable. Many disk models (e.g., the model discussed in Sec. 3) have disk dead zones at separations of ∼ 0.5 au from the star. The orbital period of material at this distance would be on the order of months, and multiple observations ngVLA Observations of EX/FUors 5 over an ngVLA observing semester could probe morphological changes in the disk over the timescale of a single orbit. Low levels of variability in EXor/FUor-type stars have been observed at optical wavelengths (i.e., significantly lower than the outbursts themselves). If this variability is correlated with changes in the disk structure, then it could be potentially observed over the course of a single observation with the ngVLA. 2.2. Disk Chemistry Winds and outflows are fundamental ingredients of star formation and protoplanetary disk evolution (Williams & Cieza 2011). Thus, it is not surprising that YSOs undergo- ing episodic accretion, such as FUors, also show high levels of outflow activity. These outflows are responsible for carrying away some of the angular momentum, allowing circumstellar matter to accrete onto the forming star. Moreover, previous studies sug- gest that outflows could be an effective way to disrupt the disk and envelope at early evolutionary stages (Ruíz-Rodríguez et al. 2017a,b). Accordingly, it is imperative to in- vestigate the relation between accretion and outflows in young stellar systems, so as to disentangle the physical mechanisms that dictate the low mass star formation efficiency in turbulent clouds (Quillen et al. 2005; Krumholz et al. 2012) and efficient transport of angular momentum that permits accretion onto the star. In addition, strong X-ray emission can be produced during FUor eruptions and rapid YSO accretion episodes more generally (Audard et al. 2014), either by accretion shocks (Grosso et al. 2010; Teets et al. 2012), magnetic reconnection (Kastner et al. 2004), or both. It has been known that X-ray radiation is a critical ionizing source in the surface disk layers at distances of ∼ 10−100 au (Glassgold et al. 2012) and plays an important role in the dispersal of disks at pre-main sequence ages >1 Myr (Gorti et al. 2015). This is due to the large penetration depth of X-rays in the disk-envelope system, where they can ionize H2 molecules as well as abundant atoms, such as He, C, N, and O. Hence, X-ray irradiation plays a role in the formation and position of molecular disk snow lines and dead zones, thereby ultimately determining the semimajor orbits, masses, and compositions of any resulting planets (Öberg & Bergin 2016; Cridland et al. 2017). The resulting enhanced molecular ionization leads to rich chemistry that can be probed by tracers such as HCN and HCO+ (Kastner et al. 2008, and refs. therein). Spatially resolved observations of outflow/disk systems with the ngVLA in the wide range of chemical tracers accessible to its Band 6 (70-116 GHz) will allow us to distinguish between chemistry induced by outflow energetics (shocks) and radiation (X-rays). Specifically, sub-arcsecond resolution mapping of lines of SO, SiO, CH3OH, and H2CO will provide diagnostics of shock-generated chemistry within FUor outflows. Observations of the same systems in HCO+ will then serve to probe the presence and extent of molecular ionization, via spatial analysis of the HCO+/13CO surface bright- ness ratio. In addition, using 12CO, 13CO, C18O line data, and adopting standard meth- ods for correcting optical depth effects (e.g., Dunham et al. 2014; Ruíz-Rodríguez et al. 2017a,b), provides estimates of the mass and kinematic properties of the revealed outflows. In parallel, maps of HCN and CS emission will trace the dense gas within the disk and disk-outflow interface regions. 2.3. Magnetic Fields & Polarization Some of the most promising outburst theories for EXors rely on the role of the stellar magnetic field. Armitage (2016) predicted that the outbursts could be explained by changes in the polarity and strength of the stellar magnetic fields at the kG-level. In 6 White et al. Simulated observations with a 10× 10 au disk model (described in Sec- Figure 1. tion 3), 0.7 mas 100 GHz ngVLA observations, and 40 mas 100 GHz ALMA obser- vations. A 2.5 au scale is given in the bottom right of each panel. The synthesized ngVLA beam is given in the bottom left of the middle panel. The simulated ngVLA observations have a 1 hr on-source σRMS = 0.48 µJy beam−1 and 1000 km baselines. a competing scenario, D'Angelo & Spruit (2012) proposed an instability which can lead to quasi-periodic oscillations in the inner disk and associated recurrent outbursts. This instability can occur when the accretion disk is truncated close to the corotation radius by the strong magnetic field of the star. A confirmation of the magnetic field strength would give strong support to these models, and may constrain the cause of outbursts to magnetic phenomena in young stars, a novel result. In principle, if kG-level stellar magnetic fields were present in a given system then there could be significant emission above the photospheric flux. For example, strong magnetic fields could induce synchrotron or free-free emission within a few stellar radii. The flux from this could be easily spatially separated from the inner edge of the disk and could provide constraints on the strength of the stellar magnetic fields. The effects of magnetic fields within the disk itelf can be traced through polariza- tion. For example, small dust grains can align with magnetic fields and cause polar- ization (e.g., Hull et al. 2014), allowing for another probe of the grain size distribution throughout the disk. Bright molecular lines can also be linearly polarized by magnetic fields within the disk (e.g., the Goldreich-Kylafis effect; Forbrich et al. 2008). The large outflow jets discussed in Sec. 2.2 can generate polarized synchrotron emission (see e.g., Galván-Madrid et al. science chapter, for a detailed overview of jets in disks). These jets have been observed in both low and high mass protostellar systems and may provide another tool in studying how accretion events can vary as a function of stellar mass. The synchrotron emission associated with these jets can lead to polarization at the lower frequency range of the ngVLA. As these jets can be much more extended from the circumstellar disk, it should be straightforward to spatially separate their emission from any disk and other polarized or non-thermal emission. 3. Uniqueness to ngVLA Capabilities The ngVLA will be able to spatially resolve a very high level of detail in the inner regions of disks. To illustrate this, Fig. 1 shows simulated observations of a detailed disk model with both the ngVLA and ALMA. The model shown is a global hydrodynamic simulation of a protoplanetary disk in the thin-disk limit, starting with the collapse of ngVLA Observations of EX/FUors 7 molecular cloud (Vorobyov & Basu 2015). The model used canonical values of layered disk parameters (e.g., Tcrit = 1100K, Σcrit = 100g/cm2) and reproduced eruptions very similar to the observations of FUor outbursts. To simulate this disk model with the ngVLA, we assume the disk emission is consistent with 100 GHz thermal continuum (i.e., it does not include any non-thermal effects, disk winds, or ionized emission), the system is at a distance of 150 pc, and the disk flux is normalized to 35 mJy. The highest resolution antenna configuration of 0.7 mas was used and the background noise is set to 0.48 µJy beam−1 (achievable in 1 hr on-source in ngVLA Band 6; Selina et al. 2018). For comparison to ALMA, a simulated observation of the same disk is included with a sensitivity equivalent to 4 hours on-source (the most extended ALMA baseline of ∼ 16 km corresponds to a resolution of ∼ 0.04 arcsec at 100 GHz). The 10×10 au region of the disk model shows the complexity of the inner accretion disk which can be resolved by ngVLA, but not with current capabilities of ALMA. Should significant optically thick disk winds or non-thermal emission be present in a system at 100 GHz, then it could indeed make resolving the disk structure more complicated. The model in Fig. 1 predicts that this inner region is very dynamic and will show detectable changes in features on the timescale of a few years. The perturbations in the disk during the outburst propagate outwards, producing spirals or concentric rings also observed in young stellar disks. ALMA is the current state-of-the-art facility in regards to high sensitivity and res- olution observations of EX/FUor-types stars. As ALMA is located in the Southern hemisphere, there are potentially many targets in the Northern hemisphere that are not easily detectable or resolved with the current Northern hemisphere observing facilities (e.g., SMA and NOEMA). Therefore, the ngVLA is uniquely positioned to provide detailed observations of a large fraction of the sky. 4. Synergies at Other Wavelengths The frequency in which EX/FUor-types of outbursts occur is not well constrained due to the relatively small number of outburst events. Determining the frequency of these events requires large scale surveys, which is more practical with optical, near-IR, and IR telescopes. Future optical/IR survey facilities, such as LSST and WFIRST, will likely identify new outbursts and the ngVLA will be a key tool in follow-up observations and an accurate characterization of the systems. The SKA may be able to detect non-thermal emission in EX/FUors and will provide complimentary longer wavelength data. Higher-energy observations, e.g., in the X-ray regime, can provide important com- plementary information on plasma heating in magnetic fields (through corona-type emission like in active stars) and the presence of neutral gas components in the line of sight, potentially indicating the presence of neutral winds or accretion flows (Lieb- hart et al. 2014). Clarifying the role of X-rays generated during massive outbursts in the evolution and dispersal of planet-forming disks and envelopes is essential to progress in connecting the first stages of planet formation to mature planetary systems. While many of the morphological details cannot be revealed with current facili- ties, mm telescopes such as ALMA and NOEMA can still determine many of the bulk properties of a given EX/FUor. For example, ALMA can provide the total ∼mm flux which will be useful for spectral index constraints in optically thin regimes. There are also many aspects of disk chemistry that can be measured with ALMA (e.g., CO(3-2) 8 White et al. at 345 GHz) and future IR facilities such as JWST, that will be complimentary to the disk chemistry information obtained with the ngVLA. The ngVLA will provide unprecedented spatial resolution and sensitivity in the 1 − 116 GHz frequency and contribute to the broad spectral coverage that is imperative to accurately models of EX/FUors and young outbursting systems. References Ábrahám, P., Juhász, A., Dullemond, C.P., et al., 2009, Nature, 459(7244), p.224 Armitage, P.J., Livio, M. & Pringle, J.E., 2001, MNRAS, 324(3), pp.705-711 Armitage, P.J., 2016, ApJL, 833(2), p.L15 Audard, M., Ábrahám, P., Dunham, M.M., et al. 2014, PPVI 387, ed H. Beuther et al. Bae J., Hartmann L., Zhu Z., Nelson R. P., 2014, ApJ, 795, 61 Bonnell, I. and Bastien, P., 1992, ApJ, 401, pp.L31-L34 Cieza, L.A., Casassus, S., Tobin, J., et al., 2016, Nature, 535(7611), p.258 Cieza, L.A., Ruíz-Rodríguez, D., Perez, S., et al. 2018, MNRAS, 474, 4347 Cridland, A.J., Pudritz, R.E., Birnstiel, T., et al., 2017, MNRAS, 469(4), pp.3910-3927 Czekala, I., Andrews, S.M., Jensen, E.L., et al., 2015, ApJ, 806(2), p.154 D'Angelo, C.R. & Spruit, H.C., 2012, MNRAS, 420(1), pp.416-429 Dong, R., Zhu, Z., & Whitney, B., 2015, ApJ, 809(1), p.93. Dunham, M.M. & Vorobyov, E.I., 2012, ApJ, 747(1), p.52. Dunham, M.M., Arce, H.G., Mardones, D., et al.,2014, ApJ, 783(1), p.29 Flaherty, K. M., Hughes, A. M., Rose, S. C., et al. 2017, ApJ, 843, 150 Forbrich, J., Wiesemeyer, H., Thum, C., et al., 2008, A&A, 492(3), pp.757-766. Gammie C.F., 1996, ApJ, 457, 355 Glassgold, A.E., Galli, D., & Padovani, M., 2012, ApJ, 756(2), p.157 Gorti, U., Hollenbach, D., & Dullemond, C.P., 2015, ApJ, 804(1), p.29. Grosso, N., Hamaguchi, K., Kastner, J.H., et al., 2010, A&A, 522, p.A56 Hales, A.S., Pérez, S., Saito, M., et al., 2018, ApJ, 859(2), p.111 Hartmann, L., Herczeg, G., & Calvet, N., 2016, ARA&A, 54, pp.135-180 Herbig, G.H., 1966, VA, 8, pp.109-125 Hull, C.L., Plambeck, R.L., Kwon, W., et al., 2014, ApJS, 213(1), p.13 Kastner, J.H., Richmond, M., Grosso, N., et al., 2004, Nat, 430(6998), p.429 Kastner, J.H., Zuckerman, B., Hily-Blant, P., et al., 2008, A&A, 492(2), pp.469-473. Kóspál, Á., Ábrahám, P., Csengeri, T., 2017, ApJ, 843(1), p.45 Krumholz, M.R., Klein, R.I., & McKee, C.F., 2012, ApJ, 754(1), p.71 Liebhart, A., Güdel, M., Skinner, S.L., Green, J. 2014, A&A, 570, id. L11 Liu, H.B., Vorobyov, E.I., Dong, R., et al., 2017, A&A, 602, p.A19 Nayakshin, S. & Lodato, G., 2012, MNRAS, 426(1), pp.70-90. Öberg, K.I. & Bergin, E.A., 2016, ApJL, 831(2), p.L19 Paardekooper, S.J. & Mellema, G., 2004, A&A 425(1), pp.L9-L12 Pfalzner, S., 2008, A&A 492(3), pp.735-741 Quillen, A.C., Thorndike, S.L., Cunningham, A., et al., 2005, ApJ, 632(2), p.941 Ricci, L., Liu, S.-F., Isella, A., & Li, H. 2018, ApJ, 853, 110 Ruíz-Rodríguez, D., Cieza, L.A., Williams, J.P., et al., 2017, MNRAS, 466(3), pp.3519-3532 Ruíz-Rodríguez, D., Cieza, L.A., Williams, J.P., et al., 2017, MNRAS, 468(3), pp.3266-3276 Selina, R.J., Murphy, E.J., McKinnon, M., et al., 2018, 2018 SPIE Conf., arxiv:1806.08405 Sicilia-Aguilar, A., Fang, M., Roccatagliata, V., et al., 2015, A&A, 580, p.A82 Teets, W.K., Weintraub, D.A., Kastner, J.H., et al., 2012, ApJ, 760, p.89 Vorobyov, E.I. & Basu, S., 2005, ApJL, 633(2), p.L137 Vorobyov, E.I., Basu, S., 2015, ApJ, 805, 115 Williams, J.P. & Cieza, L.A., 2011, ARA&A, 49, pp.67-117 Zhu, Z., Hartmann, L., & Gammie, C., 2009, ApJ, 694(2), p.1045
1910.05439
1
1910
2019-10-11T23:22:08
The Effect of Land Fraction and Host Star Spectral Energy Distribution on the Planetary Albedo of Terrestrial Worlds
[ "astro-ph.EP", "astro-ph.SR" ]
The energy balance and climate of planets can be affected by the reflective properties of their land, ocean, and frozen surfaces. Here we investigate the effect of host star spectral energy distribution (SED) on the albedo of these surfaces using a one-dimensional (1-D) energy balance model (EBM). Incorporating spectra of M-, K-, G- and F-dwarf stars, we determined the effect of varying fractional and latitudinal distribution of land and ocean surfaces as a function of host star SED on the overall planetary albedo, climate, and ice-albedo feedback response. While noting that the spatial distribution of land masses on a given planet will have an effect on the overall planetary energy balance, we find that terrestrial planets with higher average land/ocean fractions are relatively cooler and have higher albedo regardless of star type. For Earth-like planets orbiting M-dwarf stars the increased absorption of water ice in the near-infrared (NIR), where M-dwarf stars emit much of their energy, resulted in warmer global mean surface temperatures, ice lines at higher latitudes, and increased climate stability as the ice-albedo feedback became negative at high land fractions. Conversely, planets covered largely by ocean, and especially those orbiting bright stars, had a considerably different energy balance due to the contrast between the reflective land and the absorptive ocean surface, which in turn resulted in warmer average surface temperatures than land-covered planets and a stronger potential ice-albedo feedback. While dependent on the properties of individual planetary systems, our results place so constraints on a range of climate states of terrestrial exoplanets based on albedo and incident flux.
astro-ph.EP
astro-ph
Draft version October 15, 2019 Typeset using LATEX default style in AASTeX62 The Effect of Land Fraction and Host Star Spectral Energy Distribution on the Planetary Albedo of Terrestrial Worlds Andrew J. Rushby,1 Aomawa L. Shields,1 and Manoj Joshi2 1Department of Physics & Astronomy, University of California, Irvine 4129 Frederick Reines Hall, Irvine, CA. 92697-4575 2School of Environmental Sciences, University of East Anglia Norwich, U.K., NR4 7TJ (Received 21 March 2019; Revised 8 October 2019; Accepted 11 October 2019) Submitted to ApJ ABSTRACT The energy balance and climate of planets can be affected by the reflective properties of their land, ocean, and frozen surfaces. Here we investigate the effect of host star spectral energy distribution (SED) on the albedo of these surfaces using a one-dimensional (1-D) energy balance model (EBM). Incorporating spectra of M-, K-, G- and F-dwarf stars, we determined the effect of varying fractional and latitudinal distribution of land and ocean surfaces as a function of host star SED on the overall planetary albedo, climate, and ice-albedo feedback response. While noting that the spatial distribution of land masses on a given planet will have an effect on the overall planetary energy balance, we find that terrestrial planets with higher average land/ocean fractions are relatively cooler and have higher albedo regardless of star type. For Earth-like planets orbiting M-dwarf stars the increased absorption of water ice in the near-infrared (NIR), where M-dwarf stars emit much of their energy, resulted in warmer global mean surface temperatures, ice lines at higher latitudes, and increased climate stability as the ice-albedo feedback became negative at high land fractions. Conversely, planets covered largely by ocean, and especially those orbiting bright stars, had a considerably different energy balance due to the contrast between the reflective land and the absorptive ocean surface, which in turn resulted in warmer average surface temperatures than land-covered planets and a stronger potential ice-albedo feedback. While dependent on the properties of individual planetary systems, our results place so constraints on a range of climate states of terrestrial exoplanets based on albedo and incident flux. Keywords: atmospheres -- climate -- exoplanets -- ice 1. INTRODUCTION The potential habitability of a given planetary environment -- that is the ability of that environment to support the activity of at least one organism (Cockell et al., 2016) -- is strongly dependent on the amount of energy available in that system. To first order, planetary surface temperature provides the strongest control on the extent and distribution of habitable conditions, and is, in most cases, controlled by three main factors: the amount of incoming stellar radiation; the albedo or reflectivity of the surface on which that radiation is incident; and any potential greenhouse effect that may be caused by the absorption and remission of outgoing radiation by atmospheric gases such as carbon dioxide (CO2) and water vapor (H2O). Shields et al. (2013), as well as other studies (Abe et al., 2011; Joshi & Haberle, 2012; Von Paris et al., 2013) have demonstrated that bond albedo exhibits significant dependence on the stellar spectral type, particularly in the context of icy or frozen planetary surfaces. In this paper, unless otherwise stated, 'albedo' refers to the Bond albedo of a Corresponding author: Andrew J. Rushby [email protected] 2 Rushby et al. planet, which describes the total proportion or fraction of incident stellar flux, across all wavelengths, that is reflected back to space. This is in contrast to the geometric, spherical, or V-band albedo, which is wavelength and phase angle dependent. The relationship between albedo, temperature, and global ice cover represents a positive feedback within the planetary system, known as the ice-albedo feedback, that operates as ice cover advances equator-ward from the poles (due to a reduction in temperature) and planetary broadband albedo increases, thereby compounding the reduction in temperature and the further growth of icy surfaces of higher albedo (Lian & Cess, 1977; von Paris et al., 2013). Frozen surface features comprised of the same overlying material on planets around stars of different spectral types have considerably different Bond albedos, due to their spectral energy distributions (SED), which differ significantly. For example, smaller, redder stars, have peak output in the ∼0.8 to 1.2 µm range, where water ice and snow are particularly absorptive (Shields et al., 2013; Joshi & Haberle, 2012; von Paris et al., 2013). The geometric albedo of ice and snow begins to decrease at wavelengths greater than ∼0.5 µm (see figure 1), and therefore the albedo of snow and ice covered surfaces on planets orbiting red dwarfs would be proportionally lower than that of the same surface on Earth (or any Earth-like planet in orbit around a F-, G-, or K-type star) (Joshi & Haberle, 2012; Shields et al., 2013). This unique radiative environment results in the Bond albedo of ice, for example, varying by ∼40% between planets orbiting F-dwarf and M-dwarf stars, an effect which serves to dampen the ice-albedo feedback on M-dwarf planets relative to planets orbiting stars with greater output in the visible and near-UV (Shields et al., 2013). The Bond albedos of land surfaces vary due to the SED of the host star, and also as a result of variations in their composition. Given the effect of albedo on the energy-balance of planets (Budyko, 1969; Sellers, 1969), the spectral dependence of surface albedo has a significant effect on the climate and long-term planetary habitability of terrestrial planets with land and ocean surfaces, depending on the spectral energy distribution of their host stars. Spectral type also has some control over the boundaries of the habitable zone (HZ), the effect of which is primarily evident at the inner, moist- greenhouse limit of the HZ, while the outer boundaries, at which sensitivity to the ice-albedo feedback is muted, seem to be relatively unaffected (Kasting et al., 1993; Pierrehumbert, 2010; Shields et al., 2013). However, the interaction between land and ocean fractional coverage, host star SED, and albedo in the context of planetary climate stability and ice-albedo feedback dynamics has not been previously explored. We are now aware of >4000 confirmed exoplanets, a small number of which are similar in size to the Earth (0.5 > R⊕ < 1.6), potentially rocky, and orbiting M-dwarf stars. Many of the surveys that have discovered exoplanets in the past (such as Kepler ), those that are currently under way (e.g. the Transiting Exoplanet Survey Satellite (TESS ) (Ricker, 2014)), as well as soon-to-launch (e.g. James Webb Space Telescope (JWST ) (Gardner et al., 2006)) and future planned missions (e.g. the Large UV Optical Infrared Surveyor (LUVIOR) (Kopparapu et al., 2018)), focus on M-dwarf stars. Therefore, understanding the complex interactions between these planets and their host stars is crucial for determining and understanding their broader potential for long-term habitability, and for comparative planetology between similar terrestrial worlds in the orbit of stars of different spectral classes. However, we note that detailed assessments of the ability of a particular planet to support life are highly sensitive to the unique properties of the planetary and astrophysical environment (Cockell et al., 2016). In this work, we use a seasonally-varying, one-dimensional (1-D) (across latitude) energy balance model to investigate the relationship between ocean fractional coverage, broadband planetary albedo, and climate, as a function of host star SED. In subsequent sections, we will outline our methods and provide a description of the 1-D energy balance model that was adapted for use in this study. Our results are presented in §3. §4 contains a discussion of our findings, particularly in the context of planetary habitability, and outlines limitations and potential areas for future work. 2. METHODS AND MODELS Here, we describe our method of using a range of land, ocean, and icy/frozen surface types with empirically-derived albedos as input to a one-dimensional (1-D) across latitude, seasonally-varying energy-balance model (EBM). With the EBM, we explore the effect of land fraction on broadband planetary albedo as a function of the incident stellar SED. After balancing incoming and radiated energy and accounting for area, the surface temperature of a planet can be approximated as: (1) Where σ denotes the Stefan-Boltzmann constant (5.67×10−8 W m−2), T is the surface temperature, ¯α denotes the planetary Bond albedo and L(cid:12) is the incident flux from the star, normalized to the present-day Sun (1367 W m−2) σT 4 = (1 − ¯α)L(cid:12) 1 4 Effect of Land Fraction & SED on Planetary Albedo 3 (Pierrehumbert, 2010). In this simple formulation, albedo takes the form of a single, zonally averaged value (Bond albedo) accounting for all wavelengths and phase angles. This final value is a normalized, weighted average of the mean 'land albedo' ( ¯αl) and the mean 'ocean albedo' ( ¯αo), which are in turn computed as a function of SED, land fraction, latitude (φ) and temperature (T ). The ice albedo ( ¯αI ) describes a representative surface comprised of a 1:1 mixture of course-grained ice and fine- grained snow, overlying either land or ocean surfaces, computed by Shields et al. (2013), displayed in table 2 and figure 1. The initial land and ocean Bond albedos ( ¯αl0, ¯αo0) are also taken from that source and vary by host star type. The stars used to compute SED are F-dwarf HD128167, K-dwarf HD22049 (Segura et al., 2003), the G-dwarf the Sun (Chance & Kurucz, 2010), and M-dwarf AD Leo1. These values were computed using a 1-D, multistream, multilevel, line-by-line, multiple-scattering radiative transfer model (Spectral Mapping Atmospheric Radiative Transfer Model (SMART)) (Meadows and Crisp, 1996) assuming Earth-like atmospheric conditions, including 64% cloud coverage and Rayleigh scattering (Shields et al., 2013). Instellation is spectrally integrated, and varies with both latitude and time of year in response to prescribed time-independent orbital parameters. The final albedo value for the land and ocean surfaces are then scaled proportionately to their fractional coverage, FL, when used to compute surface temperature and average climate in the 1-D EBM. Table 1. Model Parameters used in this work Parameter Incident Flux (L(cid:12)) Eccentricity (e) Obliquity Rotation Rate Land Fraction (FL) OLR when T = 273 K (A) OLR Temperature Sensitivity (b) Heat Capacity, Land (CL) Heat Capacity, Water (CW ) Diffusivity for Heat Transport (D) Cloud fraction Simulation Length Value/Range (units) 1367 W m−2 0.0167 0 ◦ 1 day(s) 0.01 -- 0.99 203.3 W m−2 2.09 W m−2 K−1 0.45 W yr m−2 K−1 9.8 W yr m−2 K−1 0.44 W m−2 K−1 64% 30 yr 2.1. 1-D EBM Energy-balance models describe a suite of models of varying complexity that are used to determine the energy- balance of a planetary body, and compute a possible climate solution. In this paper, we employ a modified version of the seasonally-varying 1-D EBM first described in North & Coakley (1979) to describe the climate of the Earth, which calculates energy-balance at each latitude as the sum of absorbed shortwave radiation, outgoing longwave radiation (OLR), and the convergence of horizontal heat transport. This latter approximation is equated with vertical column heat capacity and mediated by a heat diffusion coefficient, which is set as proportional to the local meridional temperature gradient to produce an accurate representation of meridional heat transport. Roe and Baker (2010) have investigated the physical nature of the ice-albedo feedback parameterized in models of this sort, and note that the strength of the feedback depends linearly on the albedo contrast between the ice-covered and ice-free surfaces, while also being proportional to the efficiency of the heat distribution parameter (D). For this work, rotation rate, obliquity, and eccentricity are assumed to be Earth-like, and incident flux for these simulations was held at the normalized present day value (1 S⊕ = 1367 W m−2) (see Table 1). Land fraction (FL) was varied between 0.01 and 0.99 distributed uniformly within each model latitude to represent a range of continent/ocean configurations to first-order. This method results in an average planetary land fraction that mirrors the uniform land fraction at each model latitude. Sea ice is allowed to 'form' when the temperature at a grid cell is below 271 K, which in turn alters the albedo to represent the overlying frozen surface. Ice accretion and ablation is determined by the relative difference between the top-of-the-ice heat flux 1 http://vpl.astro.washington.edu/spectra/stellar/mstar.htm 4 Rushby et al. and the ocean-ice heat flux, scaled by the ice density and remains independent of atmospheric water inventory (which cannot be set in the model), and therefore ice depth increases secularly with land fraction but we do apply a linear scaling with land fraction to the ice growth model in order to limit the thickness of ice sheets at high land fraction when atmospheric water vapor levels would be expected to be low. The ocean-ice heat flux is parameterized so that it is ∼4 W m−2 at the lowest ice coverage (one grid cell), and then declines linearly to zero as ice covers the globe. Any heat supplied via the ocean-ice heat flux is subsequently 'removed' from the ice-free ocean grid cells to ensure energy conservation. The ice temperature is maintained at 271 K by accretion or ablation based on energy conservation. Therefore, the variations in ice depth between planets orbiting different stars should be considered less as absolute depths of ice and more as relative values for comparison within the context of this model to identify areas/parameter space in which ice coverage is accumulating or receding. To compute average planetary climate, additional radiative forcing from radiatively active gases such as CO2 and H2O need also to be taken into account, represented here in the linearized form A + bT. The coefficients A and b are taken from fits to satellite data from Earth (we use 203.3 W m−2 and 2.09 W m−2 K−1, respectively, from North and Coakley (1979), but see also Cess (1976), North et al. (1981) and Budyko (1969) for details on the derivation of these values) thereby constraining the applicability of this model to Earth-like atmospheres dominated by CO2 and H2O greenhouse gases. Table 2. Bond albedo values for various surface compositions used in these simulations. F-dwarf G-dwarf K-dwarf M-dwarf Land ( ¯αl) Ocean ( ¯αo) Ice ( ¯αi) 0.414 0.329 0.537 0.415 0.319 0.514 0.401 0.302 0.477 0.331 0.233 0.315 Note -- These values were taken from table 1 in Shields et al., (2013), and computed by a multiple-scattering radiative transfer code under conditions of an Earth-like atmospheric composition, cloud coverage, and Rayleigh scattering. 'Ice' here refers to a 50% mix of large grained, blue marine ice and smaller grained snow particles that forms over both frozen land and frozen ocean cells. Table 2 lists the Bond albedos of the different surface compositions that are used in this paper, and Figure 1 their broadband reflectance spectra. The model uses the spectrally-independent albedo for the composition of the surface feature, surface temperature and latitude in order to recompute a final albedo and temperature. Convergence criteria were met when average surface temperature remained stable for 10 model years under no external forcings. Unfrozen land and water take the value of ¯αl and ¯αo, respectively, and when frozen (i.e. T < Ti) both take the value of ¯αi, which in this paper refers to a 1:1 mix of fine-grained snow and course grained blue marine ice. The spectrum used for bare land in this work is that of the silicate clay mineral kaolinite taken from the USGS spectral library2. We use kaolinite as a single representative land surface composite due to its relative commonality on Earth, and note that this spectrum has also been used in concert with this model in Shields et al. (2013). Other workers, including Robinson et al. (2011), who developed a whole-disk spectral model for Earth from EPOXI spacecraft observations, also use kaolinite as a representative composite, as does Meadows et al. (2018) in their comparative planetology study of Proxima Centauri b, in which they assume a 5.5% kaolinite coverage on the modern Earth, or 23.1% for early-Earth. 2.2. Model Validation We validated the use of the EBM by reproducing Earth's bond albedo and average surface temperature at its present- day eccentricity, obliquity, approximate land/ocean ratio by latitudinal distribution (FL = 0.3), and incident stellar flux. Initial albedo values for the land, ocean, and ice-covered surfaces used in this validation simulation were taken from Shields et al. (2013) using output from SMART for a Earth-like planet orbiting a G-dwarf star, and are outlined in more detail in the preceding paragraph. We also explored the effect of differing compositions of icy surfaces, ranging from the extremely reflective, fine-grained endmember 'snow' to darker, course grained 'blue marine' ice, as well as intermediate mixtures (25%, 50%, 75%) (see figure 1). Global surface temperatures varied by 0.7 K between the pure snow and blue ice cases, as the radiative effects of the change in albedos of these surfaces are confined to the high latitudes where instellation is low. For consistency with Shields et al. (2013), we use the 50% ice/snow mixture for our 2 https://speclab.cr.usgs.gov/spectral-lib.html Effect of Land Fraction & SED on Planetary Albedo 5 Figure 1. Top: the SEDs for the F-, G-, K-, and M-dwarf stars used in this work. Below: The spectral distribution of the albedo of fine-grained snow, blue marine ice, and 25%, 50%, and 75% mixtures of the two end-members. Ocean and land albedo spectral distributions are also shown (adapted from Shields et al., 2013). frozen land and ocean surfaces for the remainder of this work. The surface temperature returned by the model (284.1 K) was within 4 K (1.4%) of the observed value (288 K). Albedo varies by planetary surface type, the distribution of which was set to represent an idealized case of the land and ocean distribution on the present Earth, and averaged 0.34 over the entire planet. This can be compared with the current broadband albedo for the Earth given by Stephens (2015) as 0.29. 2.3. Limitations and Future Work The approach undertaken here is limited by the necessary parameterization of clouds, which are implicitly represented in albedo, and the climatic effect of a combined CO2 and water vapor greenhouse, which is linearized, thereby restricting the applicability of our results to Earth-like atmospheric compositions. As we note in the previous section, ice growth in this model is temperature-dependent and occurs as a function of temperature contrast between the ocean cells and 6 Rushby et al. the ice-covered ocean cells, and not that of a dynamic water inventory. Sensitivity studies in which land and ice albedo was modified to simulate a lower rate of ice growth due to the lack of water at high land fractions revealed that planets around F- and G-dwarfs no longer experienced a global glaciations due to the lowered ice-land-ocean albedo contrast that increases climate stability and buffers the positive feedback response. We also expect these results to be affected by different land surface compositions, and expect that future modelling efforts to expand the scope of this work beyond Earth-like conditions using a 3D GCM. However, using a model of greater complexity has the drawback of increased computational burden and processor time. The approach we have used here as allowed us to explore a wide parameter space in a computationally efficient manner, while maintaining reasonable agreement with observations and other models. EBM simulations were carried out to investigate the effect of varying land fraction and star type on broadband planetary albedo, and mean global surface temperature. The results of these simulations are presented in figures 2 -- 4. 3. RESULTS Figure 2. Mean global surface temperature ( ¯T ) (left) and Bond albedo ( ¯α) (right) by average planetary land fraction (FL) for terrestrial planets orbiting FGK and M-dwarf stars. The dashed lines labeled αiX denote the albedos of the 50% ice-snow mixture on planets orbiting X-type star; the gray shaded area depicts a range of FL over which ¯α exceeds αiM ; the blue shaded region illustrates temperatures below 271 K. Markers indicate some individual land fraction and latitudinal distribution configurations tested in the model: present-day 'Earth' (E), a scenario in which a narrow band of equatorial ocean exists on an otherwise land-covered planet (Eq. O), an 'oceanplanet' (O) configuration with some land at the poles, an equatorial land belt on an otherwise oceanic planet (Eq. L.), and a 'landplanet' (L) distribution with some open ocean at the poles. See text for discussion. Effect of Land Fraction & SED on Planetary Albedo 7 Figure 3. Contours of albedo (dimensionless) as a function of land fractional coverage (FL) and latitude (φ) for a planet orbiting an M-dwarf (top left), K-dwarf (top right), G-dwarf (bottom left), and F-dwarf (bottom right) stars. Ocean, land, and icy or frozen surfaces on planets orbiting M-dwarf stars have considerably lower Bond albedos than their analogues in the orbits of brighter stars, resulting in proportionately higher average global surface temperatures. The model output displayed in figures 2 and 3 illustrates surface albedo as a function of land fraction (FL), latitude (φ), and star type. Surfaces on planets orbiting F-dwarf stars have the highest albedos, and in general the albedo of a given planet increases with increasing land fraction. Figure 2 also illustrates a region (shaded) in FL space (FL > ∼0.72) at which the planetary Bond albedo is higher than the albedo of ice on M-dwarf hosted Earth-like planets (¯α > ¯αiM ) and the ice-albedo feedback becomes a negative, or stabilizing, feedback. We tested particular land distribution scenarios (i.e. configurations in which a non-uniform (by latitude) land fraction is used as input), including present-day Earth, and equatorial ocean and land belt configurations, that revealed some variation in surface temperature (∼5 K) and albedo (∼0.02), depending on the distribution of land in the model (see labelled markers in figure 2 ). In particular, ocean-dominated planets and land planets with oceanic equatorial belts (with otherwise identical ortibal parameters) have higher global mean surface temperatures when compared to a uniform land fraction distribution by latitude. Distinct zones of high and low albedo surfaces are shown in the simulations pertaining to planets orbiting F-, G-, and K-dwarfs (figure 3 ), which represent frozen (yellow) and unfrozen surfaces, respectively. Note that these frozen surfaces exhibit different Bond albedos depending on SED, despite being comprised of the same 50% snow/ice mixture. Transitional boundaries between these zones are also evident, representing the maximum equator-ward extent of glacial conditions, i.e. a nominal 'ice line' in latitude and FL parameter space, from which we can approximate the minimum land fraction required to force mid-latitude glaciations,assuming default model conditions: 0.46, 0.55, and 0.68 FL for planets around F-, G-, K-dwarfs respectively. M-dwarf planets do not exhibit mid-latitude glaciations in any of these simulations, although albedos for high land fraction planets in orbit around these stars are proportionally higher in the mid-latitudes and equatorial regions than at the frozen polar regions. A general trend exists in that planetary broadband albedo increases with increasing FL regardless of host star type, except in the case of the frozen polar regions where albedo remains constant in most 8 Rushby et al. cases due to the persistent frozen conditions at these latitudes. Equatorial or low-latitude glaciations, analogous to a 'snowball' or 'slushball'-type event, are induced on Earth-like planets hosted by F-dwarf and G-dwarf stars when FL > 0.8. For planets orbiting smaller, less luminous stars, equatorial glaciations are not induced by high land fractions, assuming an Earth-like atmospheric water vapor content, composition of the land surface, and the presence of variable clouds (see discussion). Figure 4. Contours of approximate ice thickness (>1m) as a function of land fraction (FL) and latitude (φ) for planets orbiting M (top left), K (top right), G (bottom left), and F-type (bottom right) stars. Host star SED and land fraction also influences the maximum equatorward extent of persistent frozen surface conditions, taken here to refer to areas in which ice of at least 1 meter in thickness persists for the 30 model-year simulation (Figure 4 ). This figure displays contours of approximate ice thickness by land fraction and latitude, and demonstrates that planets orbiting F- and G- dwarf stars are especially prone to persistent and substantial ice formation extending to the mid-latitudes and equatorial regions beyond the FL glaciation limits discussed earlier. Additionally, these planets exhibit frozen surface conditions between 15 and 20 degrees further towards the equator, albeit with uniformly distributed land fraction over latitude, than in the case of an M-dwarf planet. This comparison suggests a lower sensitivity towards the ice-albedo feedback mechanism and more stable long-term climate states on M-dwarf land planets, other conditions remaining the same. As ice growth in this model is driven by temperature, we note a peak in ice depth at high land fractions (∼0.9) when temperatures are lowest for planets orbiting G- and F-stars. This is followed by a reduction in thickness as FL approaches unity due to the scaling parameterization discussed earlier in this work that prevents unrealistically thick ice sheets forming on dry, land-dominated worlds. Planets with a greater ocean fraction, under otherwise identical conditions, absorb considerably more incident energy and are on average warmer as a result (see Figures 2 -4 ). When considering a planet orbiting a G-dwarf host (with otherwise 'Earth-like' properties), increasing the fraction of the surface covered by continent reduces the global average surface temperature by 10 - 15% between simulations using 10% land coverage ( ¯Ts = 290 K) and those using 90% Effect of Land Fraction & SED on Planetary Albedo 9 land surface coverage ( ¯Ts = 258.3 K) (see Figure 2 ). While this relationship holds for all star types considered in this work, Earth-like planets orbiting M-dwarf stars remain above the threshold for freezing conditions even when covered predominately by land, maintaining an average surface temperature of ∼303.9 K at 10% and 282.6 K at 90% land fractional coverage, respectively. An average a M-dwarf 'aquaplanet' (FL < 0.15) remains ∼15 K warmer than the same type of planet around an F-dwarf when both are receiving the equivalent stellar flux as the present Earth. 4. DISCUSSION The results presented in this paper suggest that planets dominated by land are, on average, cooler and more highly reflective than those with higher fractional ocean coverage. We note a sharp transition between partially and fully glaciated planets, forced by the ice-albedo feedback, that represents the land fraction threshold at which ice growth extends from the poles to the latitudinal 'tipping point' for global glaciation. Once ice growth extends beyond this limit, the ice-albedo feedback enters a runaway state and the transition to globally glaciated conditions is sharp and rapid; this threshold varies with star type as the surface albedo contrast is wavelength dependent. The magnitude of the runaway response is sensitive to several parameters, including heat distribution efficiency and land/ice/ocean albedo contrast, which is controlled by land fraction, as well as stellar flux and SED. Given the lower albedo and increase in absorbed incident flux, coupled with the findings of previous work in this area (Shields et al., 2013; Joshi & Haberle, 2012), we would expect a significant climatological effect due stellar spectral energy distribution affecting the albedo of equivalent surfaces on planets orbiting M-dwarf stars compared to those around larger K-, G-, and F-dwarf stars. We find that, due to the reduction in the reflectance of water ice at wavelengths longer than ∼1µm, planets orbiting M-type stars are warmer and represent the only case in this model formulation where non-glacial conditions can exist on a largely land-covered planet. This effect is also pronounced at higher latitudes, where surfaces that would freeze on planets around more luminous stars remain unfrozen on M-dwarf planets, resulting in overall warmer global mean surface temperatures. Due to the interaction between the unique spectral energy distribution of M-dwarf stars and the optical properties of frozen surfaces on orbiting planets, the reversal of the effect of the ice-albedo feedback results in greater shortwave absorption and ultimately a stabilization of the climate in response to further cooling. The relative equator-ward extent of the ice line is also less evident. This effect is especially pronounced at low land fractions, and suggests that above ∼0.85 FL, the surfaces of M-dwarf planets may remain unfrozen even at polar latitudes at Earth-equivalent incident flux. Seasonal variations in temperature are also dampened for planets dominated by ocean, as land has a lower specific heat capacity that modulates the seasonal cycle of temperature. Idealized M-dwarf 'aquaplanets' are also on average ∼15 K warmer than an equivalent ocean planet orbiting a G- dwarf. While ocean albedo remains relatively similar across the visible and infrared wavelengths, the strong reflectance of marine and terrestrial icy surfaces in the shorter wavelengths at which the F-, G-, and K- stars experience their peak output (0.4 - 0.7 µm) dominates the radiative balance of these planets. This serves to reduce the amount of absorbed stellar flux at the surface and results in the formation of highly reflective frozen areas that are resilient to changes in radiative forcing by reducing both the amount and pole-ward extent of darker, unfrozen ocean areas. Our results vary somewhat from Abe et al. (2011), who used a 3D-GCM with a dynamic hydrological cycle to investigate changes to the boundaries of the habitable zone for dry and moist planets orbiting G-dwarfs. Abe et al. (2011) found land planets to be more difficult to freeze than their aquaplanet counterparts, due to unsaturated atmospheric conditions above the tropics facilitating greater longwave emission, lower rates of hydrogen escape, and lower thermal inertia. However, this work focussed on G-dwarf host stars in their simulations and a dependence of snow and ice deposition on atmospheric moisture, the albedo of which was not wavelength-dependent and scaled linearly with temperature between 0.5 and 0.75. In contrast, Turbet et al. (2016) found that when using a GCM to explore a range of potential volatile inventories for the terrestrial exoplanet orbiting the nearby mid-type M-dwarf star Proxima Centauri, a drier atmosphere and planet surface resulted in cooler climates. This difference is due to a reduction in the water vapor greenhouse effect on dry, land-covered planets. While the effect was not taken into account in this work, the increased IR emission from M-dwarfs relative to G-dwarf stars would likely result greater greenhouse warming from a water vapour atmosphere at equivalent solar flux distances thereby further accentuating the effect of the ice-land albedo contrast seen here (Kasting et al., 1993; Shields et al., 2013). At low land fractions (aquaplanets) the effect would be considerable, while land-dominated planets around M-dwarfs would remain warmer with higher latitude ice lines. Therefore, it is likely that further investigations of this effect over a range of land fractions and distributions and star types will be necessary to better determine the complex response of the climate system under these conditions. 10 Rushby et al. Our model assumed an atmospheric water content (in the context of a greenhouse forcing) parameterized for an Earth-like land/ocean fraction and distribution. We performed sensitivity studies by scaling the OLR temperature sensitivity parameter with land fraction (assuming a G-dwarf planet with the otherwise default conditions outlined in table 1), and found the results to vary little from our original results at low to moderate land fractions. At high land fractions (FL > 0.7) this scaling results in planets that are considerably cooler (-20 K) than we found with our default parameterization, but this simplified approach is likely pushing the OLR linearization beyond its intended range of applicability. We also tested OLR linearization parameterizations from Budyko (1969), Cess (1976), and Lindzen and Farrell (1977). While all of these parameterization sees a marked decline in global mean temperatures as land fractions approach unity, the primary variation between these schemes occurs at high land fractions where the latter two fits diverge to be very much cooler than our default case. Increases in b result in a greater climate sensitivity (i.e. a larger temperature response to a given perturbation). Our results therefore constitute a conservative lower limit on the possible differences in climates between planets with high and low land fractions. A more complex treatment of atmospheric water vapor requires a dynamic hydrological cycle and a coupled ocean-atmosphere, which is beyond the scope of this work. We also assumed that the average land fraction used here is representative of a planet with uniformly distributed land by latitude, but recognize that heterogeneous continental distributions (for example, the 'aquaplanet', 'equatorial ocean', 'equatorial land', or a 'landplanet' configurations shown in Figure 2 ) that exhibit the same average land fraction will result in a somewhat different energy balance from these idealized cases depending on the specific configuration of land and ocean due to the difference in specific heat capacities and thermal diffusivity between these two surface types. Furthermore, other land surface composites and mixtures have been considered in the context of exoplanet studies (e.g. Hu et al., 2012), and we expect that using a composite with higher Bond albedo will proportionately increase the land-ocean-ice contrast and therefore lower the surface temperatures of land dominated worlds, while also reducing climatic stability; a lower albedo land composition will lower this contrast and thereby buffer climate stability to a greater degree than our default composite. F-dwarf planets remain consistently cooler and more reflective than analogues orbiting smaller stars, and figures 2 through 4 illustrate that planets around these stars could experience equatorial glaciation at high land fractions (>0.8) even when receiving the equivalent incident flux as present day Earth. The high output of stars in this class in the 0.4-0.5 µm wavelength range, where ices are highly reflective, contributes to the proportionately higher albedo of these surfaces. This increased surface reflectivity and cooling also affects the location of the 'ice line' (the minimum latitude at which glacial conditions are present), which extends relatively further towards the equator in the case of brighter stars; for an F-dwarf planet with land fraction coverage of 0.2 FL the ice line falls at approximately 52 degrees, whereas the same fractional coverage and incident flux around an M-dwarf sets the ice line closer to 70 degrees (see figure 4 ). The potential for lower latitude glaciation also affects the stability, and feedback strength, of the ice-albedo feedback mechanism itself. Joshi & Haberle (2012) note that changes to the ice albedo affect the 'ice-albedo response' term (I ), a component of the 'clear-sky feedback parameter' (along with the black body response (B ) and the water vapor feedback (W )). This approach can be used to approximate the temperature response (dT ) of a radiative forcing (dF ) by dT = dF / (C + B + W + I ), where C is the cloud feedback. These feedback parameters are in units of W m−2 K−1, and will be negative if the feedback is positive (e.g. I for Earth ≈ -0.3 W m−2 K−1). Decreasing albedo, in the case of icy surfaces around M-dwarfs, increases I thereby reducing the total climate response to a change in radiative forcing, whereas an increase in albedo associated with F-dwarf hosts would decrease I and increase the planetary climate response to a given perturbation. The results presented in this work suggest that I reverses sign and becomes positive (i.e. a negative or stabilizing feedback) when considering M-dwarf planets with high land surface coverage. However, the precise magnitude of this response is contingent on the properties of the individual planetary system under consideration, and would be sensitive to variations in atmospheric composition, planet size, incident flux, as well as long-term orbital effects on climate (such as eccentricity, obliquity, precession) (Deitrick et al., 2017, 2018). Additionally, the uniform distribution of land fraction with latitude, though idealized, allows for the normalization of topographic variability across planets for the purposes of comparative planetology. The modelled differences in the ice-albedo response on planets around main-sequence stars of different spectral types have considerable implications for habitability and observability. The increasing climatic stability exhibited by the 4.1. Habitability and Observability Effect of Land Fraction & SED on Planetary Albedo 11 modelled M-dwarf planets (assuming otherwise Earth-like atmospheric conditions) regardless of land fraction suggests a boon to the habitablity potential of these worlds, while the contrasting trend (lower climate stability, lower surface temperatures for a given incident flux, and lower latitude ice lines) for F-dwarf planets may make these worlds less amenable to life. However, we note that the effects a negative ice-albedo feedback on plnaetary habitability are complex and uncertain, given that Earth remains habitable while exhibiting a positive ice-albedo feedback, and further study of this phenomena is warranted. Previous work finds an anti-correlation between stellar mass and planet occurrence rate, noting that M-dwarfs host 3.5 times more planets in the 1 - 2.8 Earth radii (R⊕) range than main-sequence FGK stars, but half the number of planets larger than 2.8 R⊕ (Mulders et al., 2015; Shields et al., 2016). These systems are also more likely to host a transiting planet in the HZ (Charbonneau & Deming, 2007). Additionally, in terms of observational prospects, the near-term detection of Earth-sized and super-Earth planets (>2 R⊕) from TESS, as well as follow-up studies carried out by ground-based radial velocity instruments or JWST, are likely to focus on M-dwarf systems (Sullivan et al., 2015; Barclay et al., 2018). The direct measurement of planetary albedo, which has been the primary focus of this work, is possible in the case of transiting giant planets using secondary eclipse measurements, but this technique is unlikely to be applicable to smaller, rocky planets due to the high transit depth and signal-to-noise required (Seager & Deming, 2016). However, the transmission spectrum of the super-Earth planet GJ 1214b provided the first evidence of a cloudy or hazy terrestrial exoplanet atmosphere (Kriedberg et al., 2014), and future space-based observatories such as JWST, as well as large (>30 m) ground-based telescopes, may provide further spectroscopic characterization of major molecules in the atmospheres of M-dwarf super-Earths for climate and habitability assessments (Seager & Deming, 2016). Kriedberg & Loeb (2016), Turbet et al. (2016) and Wolf et al. (2019) find that detecting and characterizing the potential atmosphere of nearby, Earth-like exoplanet Proxima Centauri b could be achieved by JWST using thermal phase curve observations to detect variations in heat distribution of a planet with an atmosphere versus one comprised of bare-rock. The results presented here suggest a general trend between land fraction, albedo, and climate based on star type and incident flux, which is valuable for determining the reflectivity, emmissivity and overall energy-balance of the surface. However, future observations of terrestrial planets will be affected by the composition of their atmospheres, as well as the presence of clouds and/or hazes, and, specifically in the case of planets around M-dwarf planets, UV-driven water dissociation and hydrogen escape (Luger & Barnes, 2015). Given their relative commonality (comprising ∼70% of the main-sequence stars in the stellar neighborhood), high small planet occurrence rate, and future observational opportunities, considerable theoretical and modelling efforts remain focused on the climates and potential habitability of M-dwarf systems (Shields et al., 2016). The effects of land and ocean distribution on long-term climate stability presented here should be considered in habitab ility metrics of terrestrial planets discovered around low mass stars. 5. CONCLUSIONS Using a 1-D energy balance model incorporating wavelength-dependent albedos for a range of surface types (land, ocean, water ice and snow), we quantified the effect of land and ocean fractional surface coverage on the broadband albedo and climate of Earth-like planets orbiting M-, K-, G-, and F-dwarf stars. Planets with a higher fractional coverage of land are cooler and more reflective regardless of star type, thereby increasing the sensitivity and response of the ice-albedo feedback on heavily land-dominated planets. Coupled with the lower Bond albedo of ice, land, and open ocean on planets hosted by M-dwarf stars, these planets, uniquely, can exhibit relatively higher fractional land coverage and remain unglaciated as the ice-albedo feedback reverses sign and becomes stabilizing. Ices on planets hosted by large, hot stars have a higher Bond albedo (due to the high output of these stars at shorter wavelengths) and therefore exhibit a greater sensitivity to changes in radiative forcing, which includes variations in land coverage; high land fractions result in a strong ice-albedo feedback and global glaciations beyond FL > ∼0.8 and FL > ∼0.89 for Earth- like planets hosted by F- and G-dwarf stars, respectively. 'Aquaplanets' (those dominated by ocean) are consistently warmer on average than those with greater land fractions, and exhibit lower seasonal variation in temperature due to the higher heat capacity of water. This effect is particularly pronounced on planets orbiting M-dwarfs, which may be ∼15 K warmer than a similar world orbiting an F-dwarf star. Earth-like planets hosted by M-dwarf stars are able to maintain a global mean surface temperature of 283 K at 0.9 FL and 304 K at 10% land fraction. This material is based upon work supported by the National Science Foundation under Award No. 1753373, and by NASA under grant number NNH16ZDA001N, which is part of the NASA "Habitable Worlds" program. We thank 6. ACKNOWLEDGEMENTS 12 Rushby et al. Cecilia Bitz and Igor Palubski for helpful discussions regarding this work, and Eric Wolf for providing the spectrum for AD Leo for figure 2. REFERENCES Abe, Y. et al. (2011), Astrobio 11 (5) pp. 443-460 Budyko, M. I. (1969), Tellus (21) pp. 611-619 Barclay, T. et al. (2018), ApJ: SS 239 2 Cockell, C. S. et al. (2016), Astrobio 16 (1) pp. 89-117 Cess, R.D. (1976), JAS (33) pp. 1831-1843 Chance, K. & Kurucz (2010), J Quant Spectrosc Radiat Transf (111) pp. 1289-1295 Charbonneau, D., & Deming, L. D. (2007), submitted to Exoplanet Task Force (AAAC) (arXiv:0706.1047) Deitrick, R. et al. (2017), Astron. J. 155 60 Deitrick, R. et al. (2017), Astron. J. 155 266 Deming, L. D. & Seager, S. (2016), JGR: Planets 122 pp. 53 - 75 Mulders, G. D. et al. (2015). ApJ 798 112 North, G. R. & Coakley, J. A. (1979), JAS (36) pp. 1189-1204 North, G. R. et al. (1981) Reviews of Geophysics and Space Physics 19 (1) pp. 91-121 Pierrehumbert, R. (2010). Principles of Planetary Climate. Cambridge: Cambridge University Press. Ramirez, R. (2018), Geosci 8 (280) Ricker, G.R. et al. (2014), J. of Astron. Telescopes, Instruments, & Systems 1 (1) 014003 (2014). Robinson, T.D. et al. (2011), Astrobiology 11 (5) pp. 393-408 Gardner, J. P. et al. (2006), Space Sci. Revs. 123 4 Roe, G.H. & Baker, M.B (2010), J. Climate (23) pp. pp.485-606 4694-4703 Hall, S. (2016), Sci. Am., 314, pp.14-15 Hu, R., Elhmann, B. L., Seager, S. (2012), ApJ, 752, 15 pp. Heller, R. & Barnes, R. (2013), Astrobio 13 (1) pp. 18-46 Joshi, M. & Haberle, R. (2012), Astrobio 12 (1) pp. 3-8 Kasting, J. F. et al. (1993), Icarus 101 pp. 108 128 Kopparapu, R. et al. (2018), ApJ 856 122 Kopparapu, R. et al. (2013), ApJ 765 131 Kreidberg, L. et al. (2014), Nature 505 pp. 69 - 72 Kreidberg, L. & Loeb, A. (2016), ApJL 832 L12 Lian, M.S. & Cess, R.D. (1977), JAS 34 pp. 1058 - 1062 Meadows, V. S. & Crisp, D. (1996). JGR 110 E2 pp. 4595 - Sellers, W.D. et al. (1969), J. Applied Met. and Climatology 58 (2) pp. 392-400 Segura et al. (2003), Astrobiology 3 (4) pp. 689-708 Shields, A.L. et al. (2013), Astrobio 13 (8) pp. 715-739 Shields, A.L. et al. (2016), Physics Reports 663 (2016) pp. 1-38 Stephens, G.L. et al. (2015), Reviews of Geophysics 53 pp. 141163 Sullivan, G.L. et al. (2015), ApJ 837 99 Turbet, M. et al. (2016), A&A 596 A112 4622 Von Paris, P. et al. (2013), Astrobio 13 (10) pp. 899 909 Meadows, V. S. et al. (2018). Astrobiology 18 2 pp. 133 - Wolf, E. T. et al. (2019), ApJ 877 (1) 18 pp. 189
1707.02133
1
1707
2017-07-07T12:00:19
Rotation of Cometary Nuclei: New Lightcurves and an Update of the Ensemble Properties of Jupiter-Family Comets
[ "astro-ph.EP" ]
We report new lightcurves and phase functions for nine Jupiter-family comets (JFCs). They were observed in the period 2004-2015 with various ground telescopes as part of the Survey of Ensemble Physical Properties of Cometary Nuclei (SEPPCoN) as well as during devoted observing campaigns. We add to this a review of the properties of 35 JFCs with previously published rotation properties. The photometric time-series were obtained in Bessel R, Harris R and SDSS r' filters and were absolutely calibrated using stars from the Pan-STARRS survey. This specially-developed method allowed us to combine data sets taken at different epochs and instruments with absolute-calibration uncertainty down to 0.02 mag. We used the resulting time series to improve the rotation periods for comets 14P/Wolf, 47P/Ashbrook-Jackson, 94P/Russell, and 110P/Hartley 3 and to determine the rotation rates of comets 93P/Lovas and 162P/Siding-Spring for the first time. In addition to this, we determined the phase functions for seven of the examined comets and derived geometric albedos for eight of them. We confirm the known cut-off in bulk densities at $\sim$0.6 g $\mathrm{cm^{-3}}$ if JFCs are strengthless. Using the model of Davidsson (2001) for prolate ellipsoids with typical density and elongations, we conclude that none of the known JFCs require tensile strength larger than 10-25 Pa to remain stable against rotational instabilities. We find evidence for an increasing linear phase function coefficient with increasing geometric albedo. The median linear phase function coefficient for JFCs is 0.046 mag/deg and the median geometric albedo is 4.2 per cent.
astro-ph.EP
astro-ph
MNRAS 000, 1 -- 38 (2017) Preprint 15 November 2018 Compiled using MNRAS LATEX style file v3.0 Rotation of Cometary Nuclei: New Lightcurves and an Update of the Ensemble Properties of Jupiter-Family Comets R. Kokotanekova1,2(cid:63), C. Snodgrass2, P. Lacerda3, S. F. Green2, S. C. Lowry4, Y. R. Fern´andez5, C. Tubiana1, A. Fitzsimmons3 and H. H. Hsieh6,7 1Max Planck Institute for Solar System Research, Justus-von-Liebig-Weg 3, 37077, Gottingen, Germany 2Planetary and Space Sciences, School of Physical Sciences, The Open University, Milton Keynes, MK7 6AA, UK 3Astrophysics Research Centre, Queen's University Belfast, Belfast BT7 1NN, UK 4Centre for Astrophysics and Planetary Science, School of Physical Sciences (SEPnet), The University of Kent, Canterbury, CT2 7NH, UK 5Dept. of Physics, Univ. of Central Florida, 4000 Central Florida Blvd., Orlando, FL 32816-2385, USA 6Planetary Science Institute, 1700 E. Ft. Lowell Road, Suite 106, Tucson, AZ 85719, USA 7Institute of Astronomy and Astrophysics, Academia Sinica, P.O. Box 23-141, Taipei 10617, Taiwan Accepted XXX. Received YYY; in original form ZZZ ABSTRACT We report new lightcurves and phase functions for nine Jupiter-family comets (JFCs). They were observed in the period 2004-2015 with various ground telescopes as part of the Survey of Ensemble Physical Properties of Cometary Nuclei (SEPPCoN) as well as during devoted observing campaigns. We add to this a review of the properties of 35 JFCs with previously published rotation properties. The photometric time-series were obtained in Bessel R, Harris R and SDSS r' filters and were absolutely calibrated using stars from the Pan-STARRS survey. This specially-developed method allowed us to combine data sets taken at different epochs and instruments with absolute-calibration uncertainty down to 0.02 mag. We used the resulting time series to improve the rotation periods for comets 14P/Wolf, 47P/Ashbrook-Jackson, 94P/Russell, and 110P/Hartley 3 and to determine the rota- tion rates of comets 93P/Lovas and 162P/Siding-Spring for the first time. In addition to this, we determined the phase functions for seven of the examined comets and derived geometric albedos for eight of them. We confirm the known cut-off in bulk densities at ∼0.6 g cm−3 if JFCs are strength- less. Using the model of Davidsson (2001) for prolate ellipsoids with typical density and elongations, we conclude that none of the known JFCs require tensile strength larger than 10-25 Pa to remain stable against rotational instabilities. We find evidence for an increasing linear phase function coefficient with increasing geometric albedo. The median linear phase function coefficient for JFCs is 0.046 mag/deg and the median geometric albedo is 4.2 per cent. Key words: comets: general -- comets: individual 7 1 0 2 l u J 7 . ] P E h p - o r t s a [ 1 v 3 3 1 2 0 . 7 0 7 1 : v i X r a 1 INTRODUCTION Comets are believed to preserve pristine material from the epoch of planet formation. Therefore, their proper- ties have often been studied in the search for constraints on the conditions during solar system formation. A mile- stone in cometary exploration came from the European Space Agency's Rosetta mission which followed comet (cid:63) E-mail: [email protected] © 2017 The Authors 67P/Churyumov-Gerasimenko through its perihelion pas- sage in the period 2014-2016. The successful rendezvous of Rosetta with comet 67P/C-G has provided a unique ensem- ble of comprehensive observations which are set to fully char- acterise this comet. However, in order to interpret the de- tailed measurements from Rosetta, as well as those for other comets visited by spacecraft, we need to consider them in the context of other known comets. One fundamental technique to derive the physical prop- erties of comet nuclei is through their rotational lightcurves. 2 R. Kokotanekova et al. Lightcurves can be used to extract the individual objects' spin rates and axis ratios, which in turn can be used to constrain important properties of the comets, e.g. collisional history, density, tensile strength. Additionally, knowing the lightcurve brightness variation of JFCs can significantly im- prove the results of optical studies of JFC colour and size dis- tributions. Despite being such a rich source of information, just a small fraction of JFCs have well-studied lightcurves. There are two main techniques to derive rotational lightcurves from telescope observations: (1) photometric time-series of bare nuclei and (2) periodic variability of coma structures of active comets (for an overview see Samaras- inha et al. 2004). The former relies on the direct detection of the nucleus signal, and is expected to produce more pre- cise results (Samarasinha et al. 2004). In order to detect the nucleus brightness variation directly, the comets need to be observed at large heliocentric distances when they are in- active. Observing the comets when they are weakly active can also allow reliable lightcurve derivations, but only in the cases when the nucleus signal dominates over the coma con- tribution. It is also possible to derive the nucleus rotation rate of active comet nuclei, provided that they are observed with sufficient spatial resolution to distinguish the nucleus signal from that of the coma. Such observations have been performed with the Hubble Space Telescope (HST; see Lamy et al. 2004). Additionally, comet rotations can be studied during spacecraft flybys. Such missions have allowed the rotational properties of three comets to be studied in greater detail: 9P (Chesley et al. 2013, and references therein), 103P (Belton et al. 2013, and references therein), and 67P (Jorda et al. 2016). The ground- and space-based telescope techniques for period derivations usually do not account for Sun-comet- Earth geometry changes and therefore produce synodic ro- tation periods, while the spacecraft observations allow the sidereal spin periods to be derived (e.g. see Samarasinha et al. 2004). However, the difference between the synodic and sidereal rotation periods is usually small (∼0.001 hour) even for near-Earth asteroids (e.g. Pravec et al. 1996), so the synodic rotation periods are good approximations for the spin rates. The photometric observations used to study the rota- tional properties of comets can also be a valuable source of information about the comets' surface properties. The lightcurve-resolved photometry allows a precise determina- tion of the nucleus absolute magnitude. Combined with ther- mal infrared data, the photometric magnitude can be used to determine the geometric albedo (hereafter, albedo) of the nucleus. In some cases when the comets have been observed at multiple epochs, the photometric data can be used to derive the phase functions of the nuclei. Albedos and phase functions provide us with the opportunity to characterise the surfaces of comets (e.g Li et al. 2013; Fornasier et al. 2015; Ciarniello et al. 2015) and to compare the icy populations in the Solar system (e.g. Belskaya et al. 2008; Masoumzadeh et al. 2017). To distinguish between the various small body popula- tions in the Solar system, we use the Tisserand parameter with respect to Jupiter: (cid:115) TJ = aj a + 2 a(1 − e2) aJ cos(i), (1) where e, i, and a are the eccentricity, inclination and semi- major axis of the orbit of the object, and aJ is the semi- major axis of Jupiter's orbit (aJ is approximately 5.2 au). The Tisserand parameter is a useful characteristic of the orbits of minor planets since it remains approximately con- stant for any object even after perturbations by Jupiter. (Levison 1996). For the purposes of this paper, we consider JFCs to be objects with 2 ≤ TJ ≤ 3 and periodic comet des- ignations. According to the distinction in Levison (1996), all objects with TJ > 2 are classified as ecliptic comets. Thus, the class of ecliptic comets includes objects with TJ > 3 such as 2P/Encke, active asteroids and active Centaurs. Since active asteroids and Centaurs are believed to have different physical properties from JFCs, we focus the analysis only on objects with 2 ≤ TJ ≤ 3. We, however, include 2P/Encke at TJ=3.025 as it is possible for comets of JFC origin to achieve TJ of slightly above TJ = 3 following terrestrial planet inter- actions (e.g. Levison et al. 2006). Previous extensive overviews of the known JFCs sur- face and rotation properties were published by Lamy et al. (2004), Samarasinha et al. (2004), Snodgrass et al. (2006), and Lowry et al. (2008). In this work we study the lightcurves and the surface properties of nine JFCs and com- pare them with a broad sample of JFCs with known rota- tional properties. This updated sample contains a collection of 37 well-studied comets, and allows us to investigate the population properties of JFCs for the first time after NASA's Deep Impact and EPOXI and ESA's Rosetta missions. In Section 2 we review the studies of all comets, which to our knowledge have period determinations published af- ter the last review by Snodgrass et al. (2008b). This section includes Table 1 which contains the nucleus properties of the whole sample of JFCs used in this work. We describe the observations and the method for precise absolute cali- bration of multi-epoch time-series observations in Section 3. In Section 4 we present the results from the observations of nine JFCs. After adding our comets to the rest of the JFCs with known rotational properties, we study the cumulative properties of JFCs in Section 5. Finally, Section 6 contains a summary of our results. 2 OVERVIEW OF THE KNOWN JFC ROTATION PROPERTIES The aim of this study is to combine the newly obtained nuclei properties with those from previous works in order to analyse the bulk properties of the expanded sample of JFCs. Previously, the collective rotational properties of JFCs were studied by Lamy et al. (2004), Samarasinha et al. (2004) and Snodgrass et al. (2006). We expand their samples to include the cometary nuclei whose rotations were derived since then, and complement them with the newly obtained results from this work. Table 1 contains the properties of all considered comets together with the sources of all known parameters. However, the sections below focus in MNRAS 000, 1 -- 38 (2017) detail only on the comets with updates since the reviews in (Lamy et al. 2004) and Snodgrass et al. (2006), including the unpublished HST results quoted in Lamy et al. (2004) that were revised by Lamy et al. (2011). In addition to the rotational properties, we also review below the published size and shape estimates of the con- sidered comets. While photometric lightcurves can be used to determine nucleus shapes, they do not provide absolute sizes. For those comets visited by spacecraft, the dimensions of shape models are directly measured. Radar data also pro- vide absolute sizes for the comets with close approaches to the Earth. Combined thermal infrared and optical (reflected sunlight) data allow the albedo and cross-sectional area of the body (and hence an effective radius) to be determined. For those objects with only photometric data, the nucleus size can be estimated by assuming a geometric albedo of typically 4%. The most recent reviews of comet sizes from visible photometry and thermal IR Spitzer photometry are given by Snodgrass et al. (2011) and Fern´andez et al. (2013) respectively. 2.1 Jupiter-family comets with recently updated rotation rates 2.1.1 2P/Encke Comet 2P/Encke is among the comets with the shortest known orbital periods, 3.3 years, which has allowed differ- ent observers to study its properties over multiple appari- tions. Its relatively small heliocentric distance at aphelion of 4.1 au allows the comet to stay mildly active at almost all times, which has hindered the direct observation of the comet's nucleus. Nevertheless, 2P is one of the best-studied JFCs, having well-constrained spin rate, rotation changes, colour, albedo and phase function. All of the earlier works leading to today's relatively good understanding of 2P are thoroughly described in Lamy et al. (2004) and Lowry & Weissman (2007). Newer papers have added spectroscopy of the nucleus (Tubiana et al. 2015) and a study of the aphelion activity of this comet (Kelley et al., in prep.). Here, we pro- vide an outline of the most important results on the nucleus shape and rotation rate. The earliest attempts to determine the rotational lightcurve of 2P came from Jewitt & Meech (1987). Their time-series optical photometry suggested a most-likely pe- riod of 22.43 ± 0.08 hours. A later study by Luu & Jewitt (1990) led to a best-fit period of 15.08 ± 0.08 hours, although both studies note that alternative periods were also consis- tent with their data. Fern´andez (2000) used thermal infrared time series data to confirm the 15.08 hour period. A large data set of observations between July 2001 and September 2002 when 2P was close to perihelion was used by Fern´andez et al. (2005) to determine that the comet's synodic period was either 11.079 ± 0.009 hours or 22.158 ± 0.012 hours. Fern´andez et al. (2005) also discussed that these periods are not compatible with the spin rates found by Jewitt & Meech (1987) and Luu & Jewitt (1990). Belton et al. (2005) compiled the available optical and infrared photometry and reached the conclusion that the nucleus of 2P is in a complex or excited rotation state. Ac- cording to this analysis, the nucleus precesses about the to- MNRAS 000, 1 -- 38 (2017) Rotation of Cometary Nuclei 3 tal angular momentum vector with a period 11.8 hours and oscillates around the long axis with period 47.8 hours. Lowry & Weissman (2007) added new optical data sets collected in October 2002, just a few weeks apart from some of the observations in Fern´andez et al. (2005). This allowed Lowry & Weissman (2007) to combine data from the two studies and to derive an effective radius 3.95 ± 0.06 km, an axis ratio of 1.44 ± 0.06 and a rotation period of 11.083 ± 0.003 hours. 2P was later observed during the following aphelion, and the lightcurves obtained suggested that the spin pe- riod increases by ∼ 4 minutes per orbit (Mueller et al. 2008; Samarasinha & Mueller 2013). The early nucleus size estimates of ≤ 2.9 km (Campins 1988, we use effective radius to characterise the nucleus size hereafter) and 2.8 ≤ reff ≤ 6.4 km (Jewitt & Meech 1987; Luu & Jewitt 1990) were confirmed by the later estimate of 2.4 ± 0.3 km by Fern´andez (2000) . Comet 2P was also ob- served with radar during two apparitions (Kamoun et al. 1982; Harmon & Nolan 2005). The data from Harmon & Nolan (2005) confirmed a period of ∼ 11 hours and excluded the longer periods of ∼ 15 and ∼ 22 hours. Harmon & Nolan (2005) combined the radar data with previous infrared ob- servations and obtained a solution for 2P's shape with an effective radius of 2.42 km and an axis ratio of 2.6. Fern´andez (2000) also managed to obtain the phase function of 2P with phase coefficient 0.06 mag degree−1 (in the range between 0 and 106 degrees) as well as a relatively high visual geometric albedo of 5 ± 2 %. 2.1.2 9P/Tempel 1 9P/Tempel 1 was the target for two NASA missions: Deep Impact and Stardust-NExT. It was also extensively observed from ground during the supporting campaigns (Meech et al. 2005, 2011a). Multiple authors studied the size, shape and rotation rate of 9P before the Deep Impact flyby (e.g. Weissman et al. 1999; Lowry et al. 1999; Lowry & Fitzsimmons 2001; Lamy et al. 2001; Fern´andez et al. 2003). A detailed overview of their contributions can be found in Lamy et al. (2004). The two flybys provided sufficient information to deter- mine the size of the nucleus with good precision. The mean radius of the shape model after the Deep Impact flyby was estimated as 3.0 ± 0.1 km, with axes of 7.6 and 4.9 km, and an axis ratio a/b = 1.55 (A'Hearn et al. 2005). Thomas et al. (2013a) combined the data sets from the two spacecraft and obtained a radius of 2.83 ± 0.1 km. They reported a shape model with radii between 2.10 and 3.97 km, which gives an axis ratio a/b = 1.89. The two flybys combined with the ground observing campaigns gave an insight into the rotation of 9P. Belton et al. (2011) analysed multiple available data sets and de- termined that 9P had the following sidereal rotation periods: 41.335 ± 0.005 h before the 2000 perihelion passage; 41.055 ± 0.003 h between the perihelion passages in 2000 and 2005; 40.783 ± 0.006 h from the Deep Impact photometry slightly before the 2005 perihelion passage, and 40.827 ± 0.002 h in the period 2006-2010. Chesley et al. (2013) updated their work and concluded that 9P/Tempel 1 spun up by either 12 or 17 minutes during perihelion passage in 2000 and by 13.49 ± 0.01 minutes during the perihelion passage in 2005. 4 R. Kokotanekova et al. 2.1.3 10P/Tempel 2 10P/Tempel 2 is one of the largest known JFCs. It is also known to be only weakly active at perihelion. The combina- tion of these two factors has allowed its nucleus to be ob- served with very small coma contribution both at aphelion and perihelion, making 10P one of the best-studied comets. A series of works have determined that 10P has a spheroidal shape with dimensions a=8-8.15 km and b=c=4- 4.3 km (axis ratio of 1.9), albedo AR = 2.4 ± 0.5% and ro- tation period about 9 hours (Sekanina 1987; A'Hearn et al. 1989; Jewitt & Luu 1989). A detailed summary of the works which have estimated the size of the nucleus of 10P can be found in Lamy et al. (2004). Lamy et al. (2009) used HST photometry to determine a nucleus radius of 5.98 ± 0.04 km. 10P is one of the first comets observed to change its spin rate on orbital timescales. It is progressively slowing down by ∼ 16 s per perihelion passage (Mueller & Ferrin 1996; Knight et al. 2011, 2012). The most recent analysis by Schleicher et al. (2013) led to the conclusions that 10P has a prograde rotation with a period of 8.948 ± 0.001 hours, and that the rate of spin down has decreased over time, most likely in accordance with the known decrease in water production by the comet since 1988. 2.1.4 19P/Borrelly The nucleus of comet 19P/Borelly was studied using HST images by Lamy et al. (1998b). Their analysis suggested a rotation rate of 25.0 ± 0.5 hours and dimensions of 4.4 ± 0.3 km × 1.8 ± 0.15 km, assuming an albedo of 4%. The comet was observed during five nights in July/August 2000 at the CTIO-1.5 m telescope (Mueller & Samarasinha 2002). These data yielded a lightcurve with period 26.0 ± 1 hours and a large lightcurve variation - between 0.84 mag and 1.0 mag. On September 22, 2001, just eight days after 19P passed perihelion, the NASA-JPL Deep Space 1 Mission had a flyby of the comet (Soderblom et al. 2002). Using the encounter images, Buratti et al. (2004) determined that the nucleus has a radius of 2.5 ± 0.1 km and axes 4.0 ± 0.1 km and 1.58 ± 0.06 km. Dividing these two values yields an axis ratio a/b = 2.53 ± 0.12. HST/STIS observations were conducted in parallel to the Deep Space 1 encounter (Weaver et al. 2003). They could not be used to derive an independent measure of the nucleus rotation rate but were in agreement with the previous period measurement from Lamy et al. (1998b). Mueller & Samaras- inha (2002) collected all available ground-based data from 2000 and the HST data from 2001 and improved the period by one order of magnitude. They narrowed down the possi- ble periods to three values P = 1.088 ± 0.003 days, P = 1.108 ± 0.002 days, and P = 1.135 ± 0.003 days, which were con- sistent with the initial period of P = 1.08 ± 0.04 days from Mueller & Samarasinha (2002) (Mueller et al. 2010b). These authors continued studying the comet with observations from the SOAR telescope in Chile in September/October 2014 (Mueller & Samarasinha 2015). These new data were used in an attempt to choose between the three possible ro- tation periods as well as to look for activity-induced spin changes of the nucleus during the two apparitions since the last observations. The most likely period was 1.209 days (29.016 hours) but 1.187 days (28.488 hours) could not be excluded (Mueller & Samarasinha 2015). The newly derived period suggested that the rotation of 19P slows down by ap- proximately 20 minutes per orbit (Mueller & Samarasinha 2015). 2.1.5 61P/Shajn-Schaldach Lowry et al. (2003) used snapshot observations of the nucleus of 61P (in non-photometric conditions) to determine a radius of 0.92 ± 0.24 km. Lamy et al. (2011) observed the comet at heliocentric distance 2.96 au (inbound) and determined a mean nucleus radius of 0.61 ± 0.03 km and axis ratio a/b ≥ 1.3. Their partial rotational lightcurve suggested a few possible periods, but the shortest one of them, 4.9 ± 0.2 hours was considered as most likely (Lamy et al. 2011). 2.1.6 67P/Churyumov-Gerasimenko Comet 67P/Churyumov-Gerasimenko was selected as the backup target for the Rosetta mission after the 2003 launch of the mission had to be postponed due to a failure of the Ar- iane rocket (Glassmeier et al. 2007). The comet was observed in detail during only two apparitions before the rendezvous in August 2014. The rotation period of 67P was first constrained to ∼12 hours by Hubble Space Telescope observations in March 2003, soon after its perihelion passage in September 2002 (Lamy et al. 2006). After the comet moved to greater he- liocentric distances and its activity was quenched, it was possible to directly observe the nucleus from ground and to determine the spin rate with greater precision. Lowry et al. (2012) combined all available ground observations (Lowry et al. 2006; Tubiana et al. 2008, 2011) and de- termined the sidereal rotation period of the nucleus to be P = 12.76137 ± 0.00006 hours. Mottola et al. (2014) revised the period before the second perihelion passage in 2009, and set it to P = 12.76129 ± 0.00005 hours. The next period determination was done with measure- ments from the Rosetta camera OSIRIS in March 2014 (Mot- tola et al. 2014). The new period of the comet was deter- mined as P = 12.4043 ± 0.0007 hours and suggested that the nucleus had spun up by 1285 s (∼ 21 minutes; Mottola et al. 2014). OSIRIS continued monitoring the temporal evolution of the rotation rate of 67P throughout the extent of the mis- sion (Jorda et al. 2016). The perihelion measurements of the orientation of the comet's rotational axis determined an ex- cited rotational state with period of 11.5 ± 0.5 days and an amplitude of 0.15 ± 0.03◦ (Jorda et al. 2016). They deter- mined a rotation period of 12.4041 ± 0.0001 h, which stayed constant from early July 2014 until the end of October 2014. After that, the rotation rate slowly increased to 12.4304 h until 19 May 2015, when it started dropping to reach 12.305 h just before perihelion on August 10, 2015 (Jorda et al. 2016). According to the Rosetta measurements made available by ESA1, the rotation rate continued decreasing until Febru- ary 2016, and at the end of the mission, the sidereal period of 67P was 12.055 hours (ESA provided no uncertainty on 1 http://sci.esa.int/rosetta/58367-comet-rotation-period/ MNRAS 000, 1 -- 38 (2017) this value). These measurements imply that 67P spun up by 1257 s (∼ 21 minutes) during its latest perihelion passage (2014-2016). This period change is similar to the change of 1285 s measured by Mottola et al. (2014), which suggests that the comet spins up with a rate of approximately 21 minutes per orbit. The overall spin evolution of 67P is in very close agree- ment with the activity model of Keller et al. (2015). Accord- ing to their analysis, the sign of the rotation period change is determined by the nucleus shape, while the magnitude of the change is controlled by the activity of the comet. Rosetta measured the precise dimensions of the bilobate nucleus of 67P (Sierks et al. 2015). The overall dimensions along the principal axes are (4.34 ± 0.02) × (2.60 ± 0.02) × (2.12 ± 0.06) km, with the two lobes being 4.10 × 3.52 × 1.63 km and 2.50 × 2.14 × 1.64 km (Jorda et al. 2016). Using the longest and the shortest axes of the comet, we calculated an axis ratio a/b = 2.05 ± 0.06. The mean radius derived from the shape model of 67P is 1.743 ± 0.007 km. The area equivalent radius and the volume equivalent radius are 1.93 ± 0.05 km and 1.649 ± 0.007 km, respectively (Jorda et al. 2016). 2.1.7 73P/Schwassmann-Wachmann 3 Comet 73P/Schwassmann-Wachmann had a strong outburst in September 1995 (Crovisier et al. 1995) which was accom- panied by a split-up into at least four pieces (Bohnhardt et al. 1995; Scotti et al. 1996). The remnants of the 73P nu- cleus were detected during the subsequent apparitions. The largest one of them is fragment C, which was estimated to have a radius of 0.5 km (Toth et al. 2005; Toth & Lisse 2006; Nolan et al. 2006). In 2006, the comet approached Earth to less than 1 au and provided an excellent opportunity for different observers to study the lightcurve of fragment C. Toth et al. (2005) and Toth & Lisse (2006) used HST data to determine the dimensions of fragment C. Assuming an albedo of 0.04 and a linear phase coefficient of 0.04 mag deg−1 for the R-band, they obtained an effective radius of 0.41 ± 0.02 km. The de- rived lightcurve suggested an elongated body with axes 0.57 ± 0.08 km and 0.31 ± 0.02 km, which results in a minimum axis ratio a/b ≥ 1.8 ± 0.3 (Toth & Lisse 2006). Drahus et al. (2010) collected all of the reported lightcurves (Farnham 2001; Toth & Lisse 2006; Storm et al. 2006; Nolan et al. 2006), and added a further estimate of the spin rate using variations in the production rates of the HCN molecule from sub-mm observations. Their analysis showed that 73P-C had a stable rotation during the 21-day observ- ing campaign in May 2006 and narrowed down the possible periods to 3.392 h, 3.349 h, or 3.019 h. Since none of these values could be excluded, Drahus et al. (2010) concluded that the rotation period of 73P-C was between 3.0 and 3.4 hours during the duration of their observing campaign. This is the fastest known rotation period of a JFC and its stability against rotational splitting suggests that 73P-C has a bulk tensile strength of at least 14-45 Pa (Drahus et al. 2010), or that it has a higher than expected density (see Section 5.3). Given that 73P has previously split, and continues to fragment (Williams 2017), it is most likely at the very limit of stability. MNRAS 000, 1 -- 38 (2017) Rotation of Cometary Nuclei 5 2.1.8 76P/West-Kohoutek-Ikemura Tancredi et al. (2000) observed the nucleus of 76P and esti- mated a radius of 1.3 km. However, the authors note that the collected photometric measurements of the nucleus bright- ness had a large scatter which makes the radius value un- certain. Lamy et al. (2011) obtained a partial lightcurve of the comet with most likely period of 6.6 ± 1.0 hours and brightness variation of 0.56 mag which corresponds to an axis ratio a/b ≥ 1.45. They estimated the nucleus radius to be 0.31 ± 0.01 km (Lamy et al. 2011). 2.1.9 81P/Wild 2 Comet 81P/Wild 2 was the primary target of the sample- return mission Stardust. The observations of 81P before 2004 provided an estimate of its size (summarised in Lamy et al. 2004). During the Stardust flyby in January 2004, the instruments on board revealed the shape of the nucleus as well as great details from the surface. Duxbury et al. (2004) used the obtained images to model the nucleus as a triaxial ellipsoid with radii 1.65 × 2.00 × 2.75 km ± 0.05 km, while the model of Sekanina et al. (2004) provided an effective radius of 1.98 km. The rotation rate of the comet remained unknown until 81P was observed at perigee in March/April 2010 (Mueller et al. 2010a). Their narrow-band filter photometry revealed a periodic variation in the CN features of the coma with a period of 13.5 ± 0.1 hours. 2.1.10 82P/Gehrels 3 The radius of 82P was estimated to be Reff < 3.0 km (Li- candro et al. 2000) or Reff = 2.0 km (Tancredi et al. 2000). However, 82P shows signs of activity all along its orbit (e.g. Licandro et al. 2000), and these values are therefore most likely influenced by the presence of coma. Lamy et al. (2011) obtained a partial lightcurve with a rotation period P = 24 ± 5 hours. However, the lightcurve is poorly sampled and this result most likely corresponds to a lower limit of the comet's rotation period (Lamy et al. 2011). The authors used the same data set to derive a mean radius Reff = 0.59 ± 0.04 km and axis ratio a/b ≥ 1.59. 2.1.11 87P/Bus The attempts to determine the size of the nucleus of 87P resulted in the following upper limits: rn ≤ 0.8 km (Lowry & Fitzsimmons 2001), rn ≤ 0.6 km (Lowry et al. 2003) and rn < 3.14-3.42 (Meech et al. 2004). Lamy et al. (2011) analysed a partial HST lightcurve of 87P and determined a most likely period of 32 ± 9 hours, a mean radius of 0.26 ± 0.01 km and an axis ratio a/b ≥ 2.2. 2.1.12 103P/Hartley 2 103P/Hartley 2 was extensively studied during the EPOXI flyby on 4 November 2010, and has been the target of mul- tiple ground observations due to its favourable observing geometry during close approaches to Earth. The first de- terminations of its radius rn = 0.58 km came from Jorda et al. (2000) but was later revised to rn = 0.71 ± 0.13 km 6 R. Kokotanekova et al. (Groussin et al. 2004). This result was consistent with the upper limits set by Licandro et al. (2000), Lowry et al. (2003), Lowry & Fitzsimmons (2001) and Snodgrass et al. (2008b). In preparation for the EPOXI mission Lisse et al. (2009) used Spitzer to measure an effective radius of 0.57 ± 0.08 km. This value was practically the same as the mean radius of 0.580 ± 0.018 km measured with the in situ instru- ments of EPOXI (Thomas et al. 2013b). The shape model presented in Thomas et al. (2013b) results in an estimated diameter range for the nucleus of 0.69 - 2.33 km. We divided the two extreme diameter values to obtain an axis ratio a/b = 3.38. The rotation period of 103P was studied in detail using the EPOXI data as well as the extensive support observa- tions from ground. It was established that the nucleus is slowing down during the perihelion passage and that it is in a non-principal axis rotation (A'Hearn et al. 2011; Belton et al. 2013; Drahus et al. 2011; Harmon et al. 2011; Jehin et al. 2010; Knight et al. 2011, 2015; Meech et al. 2011b; Samarasinha et al. 2010, 2011, 2012). The EPOXI lightcurve suggested several periodicities ranging from 17 to 90 hours (A'Hearn et al. 2011; Belton et al. 2013), which were used to understand the complex rotation of the nucleus (A'Hearn et al. 2011; Belton et al. 2013; Samarasinha et al. 2012). The ground observations between April 2009 and December 2010 monitored the change in the strongest periodicity of ∼ 18 hours, which corresponds to the precession of the long axis of the nucleus around the angular momentum vector (Meech et al. 2011b). Over the period covered by the cam- paign, the rotation rate increased by ∼ 2 hours, from 16.4 ± 0.1 hours (Meech et al. 2009, 2011b) to 18.4 ± 0.3 or 19 hours (Jehin et al. 2010). 2.1.13 147P/Kushida-Muramatsu 147P is among the smallest known JFC nuclei. Regarding the orbit class of this comet, Ohtsuka et al. (2008) showed that 147P is a quasi-Hilda comet, which underwent a tem- porary satellite capture by Jupiter between 1949 and 1961. Tancredi et al. (2000) reported a nucleus radius of 2.3 km but noted that the measurement is uncertain. Lowry et al. (2003) reported rn ≤ 2.0 km after a non-detection at he- liocentric distance of 4.11 au. Lamy et al. (2011) derived a complete but poorly sampled lightcurve, which suggested that the rotation period of 147P was either 10.5 ± 1 hours or 4.8 ± 0.2 hours, where the former period is slightly favoured by the obtained periodogram. They estimate a radius of 0.21 ± 0.02 km and an axis ratio a/b ≥ 1.53. 2.1.14 169P/NEAT Comet 169P/NEAT was discovered as asteroid 2002 EX12 by the NEAT survey in 2002. Later it was designated as 169P/NEAT due to the detection of cometary activity (Warner & Fitzsimmons 2005). Due to its albedo of 0.03 ± 0.01 (DeMeo & Binzel 2008) and its weak activity level, 169P is considered to be a transition object on its way to becoming a dormant comet. Warner (2006) reported the first rotational lightcurve of 169P with a double-peaked period 8.369 ± 0.05 hours and peak-to-peak amplitude ∆m = 0.60 ± 0.02 mag. Later, Ka- suga et al. (2010) observed the comet with a much larger (1.85-m) telescope and separated the nucleus brightness from the slight coma contribution.Therefore their derived lightcurve period of 8.4096 ± 0.0012 hours, photometric range ∆m = 0.29 ± 0.02 mag and consequent effective ra- dius of 2.3 ± 0.4 km are more reliable measures of the nucleus properties. However, the presence of coma during the observations done by Warner (2006) would suppress the lightcurve amplitude. Therefore the higher amplitude mea- sured by Warner (2006) must instead be the result of a more elongated shape, measured at a different aspect than Ka- suga et al. (2010), unless the coma is highly variable on a timescale shorter than the spin period. However, due to the weak levels of activity present in this comet, this level of variability is unrealistic and we adopt the larger implied axis ratio limit from the Warner (2006) data. of 2.48+0.13 Fern´andez et al. (2013) determined an effective radius −0.14 km for 169P using Spitzer mid-infrared data. 2.1.15 209P/LINEAR Hergenrother (2014) observed 209P and found its rotation rate to be either 10.93 or 21.86 hours. In May 2014, the comet had an exceptionally close approach to Earth (0.6 AU) which provided an opportunity for detailed studies of its intrinsically faint nucleus. Howell et al. (2014) used the Arecibo and Goldstone planetary radar systems to directly measure the nucleus to be 3.9 × 2.7 × 2.6 km in size, and calculated an effective radius of ∼ 1.53 km. These observa- tions ruled out the longer period by Hergenrother (2014) since the measured rotational velocities were too fast for the longer period. Schleicher & Knight (2016) also observed 209P during its perigee in May 2014. They used images obtained mainly with the 4.3 m Discovery Channel Telescope to study the coma and the nucleus of the comet. They used a small aper- ture with fixed projected size of 312 km, minimising the coma contribution so that the estimated nucleus fraction of the obtained light was 52-69 percent (Schleicher & Knight 2016). Their lightcurve was consistent with the two periods from Hergenrother (2014). However, Schleicher & Knight (2016) preferred the shorter value, 10.93 hours, since it also agreed with the radar observations. Schleicher & Knight (2016) reported that their lightcurve had a different shape than the one in Hergenrother (2014). Additionally, they mea- sured variation of 0.6-0.7 mag, which is larger than the pre- diction of 0.4 mag based on the radar measurements. These differences can be explained by a possible interplay between shape and viewing geometry as well as albedo effects (Schle- icher & Knight 2016). Despite these discrepancies, all three investigations agree on the spin period of 10.93 hours. 2.1.16 260P/McNaught 260P was discovered in 2012, and the most reliable estimate of its effective radius to date is 1.54+0.09 −0.08 km (Fern´andez et al. 2013). Its rotational characteristics were studied by Manzini et al. (2014) with ground photometric observations while the comet was around perihelion in 2012 and 2013. Manzini et al. (2014) used coma structures to constrain the pole orientation of the comet, but they were unable to use the coma morphology to derive a rotational period. Instead, MNRAS 000, 1 -- 38 (2017) the comet's lightcurve was obtained by measuring the coma brightness with apertures larger than the seeing disc but small enough to include only contribution from the coma at a distance up to 2000 -- 2500 km from the surface (Manzini et al. 2014). The resulting lightcurve had a variation of 0.07 mag and could be phased with a few possible periods, best summarised as 8.16 ± 0.24 hours. While the method used in Manzini et al. (2014) has been used successfully to derive other rotations periods of comets with weak jet activity (e.g. Reyniers et al. 2009), we regard the results on 260P with caution. It is very likely that the coma contribution in the selected apertures dilutes the received nucleus signal and dampens the possible varia- tion caused by rotation. Therefore the limit on the nucleus elongation derived from the brightness variation is a weak constraint on the nucleus shape. 2.1.17 322P/SOHO 1 Comet 332P/SOHO 1 was discovered by SOHO as C/1999 R1, but after it was identified again in the SOHO fields during the following apparitions (Hoenig 2005), it became the first SOHO-discovered comet with conclusive orbital periodicity. The observations of 322P during four consec- utive apparitions displayed no clear signatures of a coma or tail and showed a nearly identical asymmetrical heliocentric lightcurve, implying repeated activity at similar levels each orbit (Lamy et al. 2013). Despite its comet-like orbit with Tisserand parameter with respect to Jupiter of 2.3, the unusual properties of 322P suggest that it has asteroidal rather than cometary origin (Knight et al. 2016). Their optical lightcurve indi- cates a fast rotation rate of 2.8 ± 0.3 hr and photometric range of >∼ 0.3 mag. These figures imply a density of > 1000 kg m−3, which strengthens the argument for asteroidal ori- gin (Knight et al. 2016). This density is significantly higher than the typical values of other known comets but is typi- cal for asteroids (see Section 5.3). Additionally, the colour of 322P is indicative of V- and Q-type asteroids, and its albedo (estimated to be between 0.09 and 0.42) is higher than the albedos measured for any other comet (Knight et al. 2016). These, together with the very low activity of the nucleus, indicate the possibility that 322P is an asteroid which be- comes active when very close to the Sun. However, since no other comet nucleus has been studied so close to the Sun, it is not excluded that it has a cometary origin, but proximity to the Sun has changed the properties of its surface (Knight et al. 2016). 2.2 Comets with new rotation rates derived in this work 2.2.1 14P/Wolf Rotation of Cometary Nuclei 7 et al. 2013). SEPPCoN used Spitzer infra-red photometry to measure sizes, and should be more reliable than visible photometry from earlier ground-based surveys. Snodgrass et al. (2005) obtained time-series of the bare nucleus of 14P on 20 and 21 January 2004 with the New Technology Telescope (NTT) in La Silla. The observations showed a clear brightness variation of the nucleus with a pe- riod of 7.53 ± 0.10 hours. The peak-to-peak variation of the lightcurve was 0.55 ± 0.05 mag, which corresponds to an axis ratio a/b ≥ 1.7 ± 0.1. The mean absolute magnitude of the time series was 22.281 ± 0.007, which suggested an effective radius 3.16 ± 0.01, assuming an albedo of 4% (Snodgrass et al. 2005). In Section 4.1 we provide the results from our lightcurve analysis. We combined the re-analysed data from 2004 with a SEPPCoN dataset from 2007 in order to improve the lightcurve of the comet and to derive its phase function. 2.2.2 47P/Ashbrook-Jackson The early estimates of the nucleus size of 47P from photo- metric observations close to aphelion determined an effec- tive radius Reff = 3.0 km Licandro et al. (2000) and Reff = 2.9 km (Tancredi et al. 2000). Snodgrass et al. (2006) and Snodgrass et al. (2008b) observed the nucleus in 2005 and 2006 at large heliocentric distance close to aphelion and es- timated Reff = 2.96 ± 0.05 km. However, their photometric comet profiles showed signatures of activity, and therefore this estimate was considered an upper limit of the nucleus size. Lamy et al. (2011) used HST observations of the active nucleus of 47P to determine a mean effective radius of 2.86 ± 0.08 km. The most recent effective radius measurement of 3.11+0.20 −0.21 km was obtained within the SEPPCoN survey (Fern´andez et al. 2013). Lamy et al. (2011) derived a partial lightcurve with mul- tiple possible periods. Analysing the periodogram, they sug- gested that the rotation period of the comet is ≥ 16 ± 8 hours. Both Snodgrass et al. (2008b) and Lamy et al. (2011) attempted to constrain the phase function of 47P by com- bining all mentioned photometric observations. While the analysis of Snodgrass et al. (2008b) clearly suggested a lin- ear phase function with a slope β = 0.083 mag/deg , Lamy et al. (2011) showed that a less steep phase function similar to that of 19P/Borelly (0.072 ± 0.020 %; Li et al. 2007b) is also possible. In Section 4.2, we show the result from our analysis of the data from Snodgrass et al. (2008b) complemented by a new data set obtained in 2015. We determined the lightcurve and the phase function of 47P, but the derived results need to be considered with caution since the comet was active during both observing runs. 2.2.3 93P/Lovas The first attempt to find the size of the nucleus of comet 14P/Wolf resulted in an effective radius of 1.3 km (Tancredi et al. 2000). However, the authors classified the estimate as poor due to the large scatter in the data points. Lowry et al. (2003) determined a radius of 2.3 km using snapshots of the comet at large heliocentric distance (3.98 au). The most recent value for the comet effective radius is 2.95 ± 0.19 km, obtained within the SEPPCoN survey (Fern´andez Comet 93P/Lovas was one of the targets of the SEPPCoN survey. Its effective radius Reff = 2.59± 0.26 km was derived from Spitzer thermal emission observations (Fern´andez et al. 2013). Our optical time-series observations are presented in Section 4.3. Despite the weak activity detected on the frames, we attempted to constrain the comet's rotation lightcurve. MNRAS 000, 1 -- 38 (2017) 8 R. Kokotanekova et al. 2.2.4 94P/Russell 4 Tancredi et al. (2000) tried to estimate the effective radius of 94P. However, at the time of the observations, the comet exhibited slight activity and the absolute magnitude mea- surements of the nucleus had large scatter. Therefore Tan- credi et al. (2000) considered their effective radius estimate of 1.9 km as uncertain and estimated the error bars of the measurement to be between ± 0.6 and ± 1 mag. Snodgrass et al. (2008b) observed the comet during four nights in July 2005 at heliocentric distance 4.14 au, out- bound. The analysis pointed to a nucleus with effective ra- dius of 2.62 ± 0.02 km and a lightcurve with period ∼ 33 hours (Snodgrass et al. 2008b). The peak-to-peak variation of the lightcurve was 1.2 ± 0.2 mag, implying axis ratio a/b ≥ 3.0 ± 0.5. Their nucleus size estimate Reff = 2.62 ± 0.02 km is in a good agreement with the SEPPCoN Spitzer data from Fern´andez et al. (2013), who reported an effective radius of 2.27+0.13 −0.15 km. In Section 4.4, we present two additional data sets from 2007 and 2009 with time-series photometry of 94P. They allowed us to determine the rotational lightcurve and the phase function of the comet. 2.2.5 110P/Hartley 3 110P/Hartley 3 was observed with HST on November 24 2000 at heliocentric distance of 2.58 au, inbound (Lamy et al. 2011). The data yielded an estimate of the effective radius of the nucleus Reff = 2.15 ± 0.04 km and a lightcurve with period 9.4 ± 1 hours. The peak-to-peak amplitude of the obtained lightcurve was 0.4 mag, which suggested an axis ratio a/b ≥ 1.30. In Section 4.5, we analyse a further data set from 2012 which our team had obtained in order to derive the comet's phase function. We used the data to derive a precise phase function of the comet as well as to constrain better the lightcurve of 110P. 2.2.6 123P/West-Hartley Tancredi et al. (2000) estimated a radius of 2.2 km for the nucleus of comet 123P/West-Hartley. However, the authors consider this result as very uncertain as the individual pho- tometric measurements of the comet nucleus displayed a large scatter. The SEPPCoN mid-infrared observations of 123P yielded an effective radius of 2.18 ± 0.23 km (Fern´an- dez et al. 2013). In Section 4.6 we present the results from our analy- sis of a SEPPCoN data set from three observing nights in 2007. The comet was very faint (mr = 23.3 ± 0.1 mag) and weakly active during the observations, which significantly obstructed the lightcurve analysis. of the still active nucleus of 137P at heliocentric distance 2.29 au. Tancredi et al. (2000) observed the comet at 5 au from the sun and estimated the effective nucleus radius to be 2.9 km. Finally, Fern´andez et al. (2013) targeted the comet as part of SEPPCoN and measured an effective radius of 4.04+0.31 −0.32 km. Snodgrass et al. (2006) obtained time-series photometry from one night on NTT/EMMI in La Silla. The data did not show brightness variation within the 3 hours of the observa- tions and could not be used to determine the rotation rate of the nucleus. However, Snodgrass et al. (2006) used these frames to estimate the nucleus radius as 3.58 ± 0.05 km. We added 2 further nights of time-series obtained within SEP- PCoN to the one night reported in Snodgrass et al. (2006) and we used the combined data set in an attempt to char- acterise the phase function and the rotational properties of the comet (Section 4.7). 2.2.8 149P/Mueller 4 149P/Mueller was among the SEPPCoN targets. The Spitzer observations revealed a nucleus with an effective ra- dius of 1.42+0.09 −0.10 km (Fern´andez et al. 2013). To our knowl- edge, no previous lightcurves of this comet are available. In Section 4.8, we present an analysis of the optical ob- servations taken as part of SEPPCoN. We use the data to derive the phase function of the comet and to place con- straints on its shape and albedo. 2.2.9 162P/Siding Spring Comet 162P was discovered as asteroid 2004 TU12 but was later identified as a comet since it shows weak intermittent activity (Campins et al. 2006, and references therein). Fernandez et al. (2006) analysed its thermal emission from NASA's Infrared Telescope Facility in December 2004 during the same apparition. Their measurements suggested a remarkably large nucleus with an effective radius of 6.0 ± 0.8 km (Fernandez et al. 2006). 162P was also observed within SEPPCoN. The Spitzer mid-infrared observations from 2007 provided a more precise estimate of the effective radius, Reff = 7.03+0.47 −0.48 km (Fern´andez et al. 2013). There are no published rotational lightcurves of the nu- cleus of 162P to our knowledge. However, there is a well- sampled lightcurve with period Prot ∼ 33 hours by the am- ateur observatory La Canada2. Those data were taken in November 2004, just a month after the discovery of the comet. In Section 4.9, we analyse two time-series data sets from 2007 and 2012. These data allow us to derive the phase func- tion of 162P and to estimate its rotation period at two dif- ferent epochs. 2.2.7 137P/Shoemaker-Levy 2 2.3 Other objects Licandro et al. (2000) observed 137P at heliocentric distance 4.24 AU and determined an effective radius of 4.2 km and a brightness variation of 0.4 mag. As described in Licandro et al. (2000), their observations suffered from different tech- nical problems, and therefore this result is uncertain. Lowry et al. (2003) obtained a radius ≤ 3.4 km from observations There are a number of objects which are not comets but have been observed as active during multiple orbits, and therefore have been given periodic-comet designations. These objects 2 http://www.lacanada.es/Docs/162P.htm MNRAS 000, 1 -- 38 (2017) are either Centaurs or active asteroids, and can be distin- guished from JFCs dynamically using the Tisserand param- eter with respect to Jupiter. While JFCs have 2 ≤ TJ ≤ 3, Centaurs have a Jovian Tisserand's parameter above 3.05 and semi-major axes between these of Jupiter and Neptune. The list of Centaurs with known activity includes 29P/Schwassmann- Wachmann 1, 39P/Oterma, 95P/Chiron, 165P/Linear, and 174P/Echeclus (see Jewitt 2009). JFCs are likely to have originally been Centaur objects as both are believed to have evolved from the scattered disk in the Kuipter belt inwards towards the inner Solar system (e.g. Duncan et al. 2004; Volk & Malhotra 2008). However, the known active Centaurs are larger than JFCs and show mass loss at heliocentric dis- tances larger than 5 au where water sublimation cannot be the major driving mechanism for the observed activity. This suggests that Centaurs are shaped by different processes and must be studied as a separate population. Active asteroids have semi-major axes a < aJ and TJ > 3.08 (see Jewitt et al. 2015). Despite showing evidence for mass loss, these objects have typical asteroid-like charac- teristics such as orbital dynamics, colours, and albedos (for a review, see Jewitt et al. 2015). Active asteroids must there- fore also be considered as a separate population from JFCs, and we do not include them when considering the ensemble properties of JFCs (in Section 5). 3 OBSERVATIONS AND DATA ANALYSIS 3.1 Data collection The main goal of this paper is to expand the sample of JFCs with known rotational properties, in an attempt to define better constraints on their bulk properties. Below we present the optical lightcurves of nine JFC nuclei which were observed in the period 2004-2015 (Table 2). Most of the data come from SEPPCoN (Survey of En- semble Physical Properties of Cometary Nuclei). SEPPCoN surveyed over 100 comets between 2006 and 2013 in order to determine distributions of the radius, geometric albedo, thermal inertia, colours, and axis ratio of the JFC nuclei (Fern´andez et al. 2013). The survey combined mid-infrared measurements from the Spitzer Space Telescope with quasi- simultaneous ground-based visible light observations from 2- 8m telescopes. For a small subset of the SEPPCoN targets, the optical data sets included long time-series observations aimed at detecting the rotational variation of bare nuclei around aphelion. Here, we present the lightcurves of eight of those comets. The remaining comets had time series which were not sufficient to measure reliable brightness variations. They will be included in a further publication which will focus on the sizes, albedos and phase curves of all observed comets. For some of the SEPPCoN comets presented below, we were also able to retrieve archival time-series from other programmes. For 14P and 94P, this included already pub- lished data from previous studies (Snodgrass et al. 2005, 2006). These archival data sets could be consolidated with the newly obtained data, since all observations were from the same aphelion passages. All observations were analysed with our newly developed method which ensured that the com- bined time series from all different epochs were consistent. MNRAS 000, 1 -- 38 (2017) Rotation of Cometary Nuclei 9 Combining all available data allowed us to derive more ac- curate lightcurves and phase functions for these two comets. Comet 47P was also part of SEPPCoN although it was at an unfavourable orbital configuration during the ground observing campaign. We managed to collect time series of the comet later, in 2015, when 47P was observed as a backup target of the ESO large program 194.C-0207. These data were combined with an archival data set from 2005 (Snod- grass et al. 2008b). Another major source of time-series data were the ESO observing programmes P87.C-107 and P89.C-0372. Those campaigns, led by our team, aimed to follow the same comets over an extended period in order to provide a good phase-function sampling. Despite having a different observ- ing strategy, those datasets were suitable for the extraction of rotational lightcurves. They provided short-time series of comets 110P and 162P over the course of a few months. Although the data came from different epochs and geome- tries, they could be linked together owing to our specially- developed procedure for absolute photometric calibration described in section 3.4. 3.2 Instruments The lightcurve data analysed in this paper were obtained from five different instruments on four telescopes (see Table 2). Comets 14P, 47P, 94P, 123P and 137P were observed us- ing the red arm of the EMMI instrument which was mounted at the f/11 Nasmyth-B focus of the 3.6m New Technology Telescope (NTT) at the European Southern Observatory's (ESO) La Silla site. The red arm of EMMI was equipped with a mosaic of two MIT/LL 2048 × 4096 CCDs. The ob- servations were done in 2 × 2 binning mode which gave a pixel scale of 0.332 arcsec pixel-1. The effective size of the field of view was 9.1 × 9.9 arcmin2. All images presented here were taken with the Bessel R filter. EFOSC2 replaced EMMI at the Nasmyth focus of the NTT in 2008 (Buzzoni et al. 1984; Snodgrass et al. 2008a). The effective field of view of EFOSC2 is 4.1 × 4.1 arcmin2. It contains a LORAL 2048 × 2048 CCD which was used in a 2 × 2 binning mode with an effective pixel scale of 0.24 arcsec pixel-1. The observations of comets 93P, 94P, 110P, 149P and 162P were taken through a Bessel R filter, while 47P was observed with an SDSS r' filter. Some of the data for the lightcurves of 93P, 110P, 149P and 162P were obtained with the visual and near-UV FOcal Reducer and low-dispersion Spectrograph (FORS2) instru- ment at ESO's 8.2 m Very Large Telescope (VLT) on Cerro Paranal, Chile (Appenzeller et al. 1998). The detector of FORS2 consists of a mosaic of two 2k × 4k MIT CCDs. The pixel scale at the default readout mode used (2 × 2 pixel binning) is 0.25 arcsec pixel-1. The field of view of the in- strument is 6.8 × 6.8 arcmin2. Comets 14P, 93P, 149P and 162P were observed with the 4.2m William Herschel Telescope (WHT) at the Roque de Los Muchachos observatory on the island of La Palma, Spain. The observations were done using the Prime Focus Imaging Platform (PFIP) which contains an optical mosaic of two EEV 2k × 4k CCDs. The total field of view of the instrument is 16.2 × 16.2 arcmin2 with a gap of 9 arcsec between the two chips. Both chips were used in an unbinned 10 R. Kokotanekova et al. Table 1. Summary of the properties of the comets with published rotation rates and the comets studied in this work Comet R eff (km) 3.95 ± 0.06 2.23+0.13 −0.15 2.64 ± 0.17 2.83 ± 0.1 5.98 ± 0.04 2.95 ± 0.19 1.62 ± 0.01 2.5 ± 0.1 1.0 2.15 ± 0.17 10.7 ± 0.7 1.65+0.11 −0.12 2.55 ± 0.01 0.56 ± 0.04 3.11+0.20 −0.21 2.97+0.19 −0.20 4.24 ± 0.2 0.61 ± 0.03 1.649 ± 0.007 0.41 ± 0.02 0.31 ± 0.01 1.98 ± 0.05 0.59 ± 0.04 0.26 ± 0.01 2.08 ± 0.01 2.59 ± 0.26 2.27+0.13 −0.15 0.58 ± 0.018 2.31 ± 0.03 3.87+0.26 −0.21 2.18 ± 0.23 4.04+0.31 −0.32 4.79+0.32 −0.33 0.21 ± 0.02 1.42+0.09 −0.10 7.03+0.47 −0.48 2.48+0.13 −0.14 ∼ 1.53 1.54+0.09 −0.08 0.150 - 0.320 2P 6P 7P 9P 10P 14P 17P 19P 21P 22P 28P 31P 36P 46P 47P 48P 49P 61P 67P 73P 76P 81P 82P 87P 92P 93P 94P 103P 110P 121P 123P 137P 143P 147P 149P 162P 169P 209P 260P 322P Ref. R eff (1) (2) (2) (5) (8) (2) (11) (12) (14) (2) (18) (2) (21) (22) (2) (2) (18,25,26) (27) (28) (30) (27) (32) (27) (27) (4) (2) (2) (35) (*) (2) (2) (2) (2) (27) (2) (2) (2) (40) (2) (43) ∆m 0.4 ± 0.04 0.082 ± 0.016 0.30 ± 0.05 0.6 ± 0.2 0.7 0.37 ± 0.05 0.30 ± 0.05 0.84-1.00 0.43 0.55 ± 0.07 0.45 ± 0.07 0.5 ± 0.1 0.7 ± 0.1 0.38 0.33 ± 0.06 0.32 ± 0.05 0.5 0.26 0.4 ± 0.07 - 0.56 - 0.58 0.94 0.6 ± 0.05 0.21 ± 0.05 1.11 ± 0.09 -- 0.20 ± 0.03 0.15 ± 0.03 0.5 ± 0.1 0.18 ± 0.05 0.45 ± 0.05 0.40 0.11 ± 0.04 0.59 ± 0.04 0.60 ± 0.02 0.4 - 0.7 0.07 ≥ 0.3 Ref. ∆m a/b Ref. a/b (1) (3) (4) (6) (9) (*) (11) (13) (15) (17) (19) (20) (21) (22) (*) (24) (25) (27) (29) - (27) - (27) (27) (4) (*) (*) -- (*) (21) (*) (*) (37) (27) (*) (*) (38) (40,41) (42) (43) ≥ 1.44 ± 0.06 ≥ 1.08 ≥ 1.3 ± 0.1 1.89b ≥ 1.9 ≥ 1.41 ± 0.06 ≥ 1.3 ± 0.1 2.53 ± 0.12b ≥ 1.5 ≥ 1.66 ± 0.11 ≥ 1.51 ± 0.07 ≥ 1.6 ± 0.15 ≥ 1.9 ± 0.1 ≥ 1.4 ± 0.1 ≥ 1.36 ± 0.07 ≥ 1.34 ± 0.06 ≥ 1.63 ± 0.07 ≥ 1.3 2.05 ± 0.06b ≥ 1.8 ± 0.3 ≥ 1.45 1.67 ± 0.04 ≥ 1.59 ≥ 2.2 ≥ 1.7 ± 0.1 ≥ 1.21 ± 0.06 ≥ 2.8 ± 0.2 3.38b ≥ 1.20 ± 0.03 ≥ 1.15 ± 0.03 1.6 ± 0.1 1.18 ± 0.05 ≥ 1.49 ± 0.05 ≥ 1.53 1.11 ± 0.04 ≥ 1.72 ± 0.06 ≥ 1.74 ± 0.03 ≥ 1.55 ≥ 1.07 ≥ 1.3 (1) -a (4) (5) (9) (*) (11) (12) (15) (17) (19) (20) (21) (22) (*) (24) (25) (27) (28) (30) (27) (33) (27) (27) (4) (*) (*) (35) (*) (21) (*) (*) (18) (27) (*) (*) -a (40) -a (43) P rot 11.0830 ± 0.0030 6.67 ± 0.03 7.9+1.6−1.1 41.335 ± 0.005c 8.948 ± 0.001 9.02 ± 0.01 7.2/8.6/10.3/12.8 26.0 ± 1.0 9.50 ± 0.2 12.30 ± 0.8 12.75 ± 0.03 5.58 ± 0.03 ∼ 40 6.00 ± 0.3 15.6 ± 0.1 29.00 ± 0.04 13.47 ± 0.017 4.9 ± 0.2 12.055 ± 0.001 3.0 - 3.4 6.6 ± 1.0 13.5 ± 0.1 ≥ 24 ± 5 32 ± 9 6.22 ± 0.05 18.2+1.5−15 20.70 ± 0.07 16.4 ± 0.1 10.153 ± 0.001 10+8−2 -- -- 17.21 ± 0.1 10.5 ± 1 / 4.8 ± 0.2 -- 32.853 ± 0.002 8.4096 ± 0.0012 10.93 ± 0.020 8.16 ± 0.24 2.8 ± 0.3 Ref. P rot (hr) (1) (3) (4) (7) (10) (*) (11) (13) (16) (17) (19) (20) (21) (23) (*) (24) (25) (27) ESA/Rosetta (31) (27) (34) (27) (27) (4) (*) (*) (36) (*) (21) -- (37) (27) -- (*) (39) (40,41) (42) (43) a Calculated with Eq. 5 using the brightness variation ∆m. b The exact shape model was derived by spacecraft observations in the cited paper. The provided axis ratio is obtained by dividing the highest shape model radius to the lowest one. c The comet is known to increase its period and this is the minimum known value measured with sufficient precision. * Results derived in this work. References: 1 Lowry & Weissman (2007); 2 Fern´andez et al. (2013); 3 Gutierrez et al. (2003); ; 4 Snodgrass et al. (2005); 5 Thomas et al. (2013a); 6 Fern´andez et al. (2003); 7 Belton et al. (2011); 8 Lamy et al. (2009); 9 Jewitt & Luu (1989); 10 Schleicher et al. (2013); 11 Snodgrass et al. (2006); 12 Buratti et al. (2004); 13 Mueller & Samarasinha (2002); 14 Tancredi et al. (2000); 15 Mueller (1992); 16 Leibowitz & Brosch (1986); 17 Lowry & Weissman (2003); 18 Lamy et al. (2004); 19 Delahodde et al. (2001); 20 Luu & Jewitt (1992); 21 Snodgrass et al. (2008b); 22 Boehnhardt et al. (2002); 23 Lamy et al. (1998a); 24 Jewitt & Sheppard (2004); 25 Millis et al. (1988); 26 Campins et al. (1995); 27 Lamy et al. (2011); 28 Jorda et al. (2016); 29 Tubiana et al. (2008); 30 Toth & Lisse (2006); 31 Drahus et al. (2010); 32 Sekanina et al. (2004); 33 Duxbury et al. (2004); 34 Mueller et al. (2010a); 35 Thomas et al. (2013b); 36 Meech et al. (2009); 37 Jewitt et al. (2003); 38 Warner (2006); 39 Kasuga et al. (2010); 40 Howell et al. (2014); 41 Schleicher & Knight (2016); 42 Manzini et al. (2014); 43 Knight et al. (2016) mode with a pixel scale of 0.24 arcsec pixel-1. All obser- vations were done using CCD2, as it has fewer bad pixels and defective columns than CCD1. The filter used for the observations was Harris R with a central wavelength 640.8 nm. Finally, the re-analysed dataset from Snodgrass et al. (2006), used to obtain the lightcurve of 94P, was taken using the 2.5m Isaac Newton Telescope (INT) at the Roque de Los Muchachos observatory. The Wide Field Camera (WFC), mounted at the primary focus of INT, was used for the ob- MNRAS 000, 1 -- 38 (2017) Table 2. Summary of all analysed observations. Comet UT date Rh [au]a ∆ [au] α [deg.] Filter Number Exposure time (s) Instrument Proposal ID Rotation of Cometary Nuclei 11 14P 47P 93P 94P 110P 123P 137P 149P 162P 2004-01-20 2004-01-21 2007-05-14 2007-05-18 2007-05-19 2005-03-05 2005-03-06 2006-06-01 2015-04-19 2015-04-21 2015-04-22 2015-04-23 2015-04-24 2009-01-21 2009-01-22 2009-01-24 2009-01-27 2009-01-28 2009-01-29 2005-07-04 2005-07-05 2005-07-06 2005-07-07 2007-07-17 2007-07-18 2007-07-19 2007-07-20 2009-01-22 2009-01-27 2009-01-28 2009-01-29 2012-06-17 2012-06-18 2012-06-22 2012-06-24 2012-07-12 2012-07-15 2012-07-26 2012-08-19 2007-07-17 2007-07-18 2007-07-20 2005-03-06 2007-05-13 2007-05-14 2009-01-21 2009-01-22 2009-01-23 2009-01-24 2009-01-27 2009-01-28 2009-01-29 2007-05-17 2007-05-18 2007-05-19 2012-04-23 2012-05-24 2012-06-14 2012-06-17 2012-06-23 5.51O 5.51O 4.36I 4.35I 4.34I 5.42I 5.42I 4.96I 4.55I 4.55I 4.55I 4.54I 4.54I 3.79O 3.80O 3.81O 3.83O 3.83O 3.84O 4.14O 4.14O 4.14O 4.15O 4.68I 4.68I 4.68I 4.68I 3.41I 3.39I 3.39I 3.39I 4.51I 4.51I 4.50I 4.50I 4.47I 4.47I 4.45I 4.41I 5.57O 5.57O 5.57O 6.95I 5.26I 5.25I 3.56I 3.56I 3.56I 3.55I 3.54I 3.54I 3.54I 4.86O 4.86O 4.86O 4.73O 4.77O 4.80O 4.80O 4.81O 4.96 4.95 3.43 3.41 3.41 4.47 4.47 4.23 3.64 3.62 3.61 3.60 3.60 3.25 3.24 3.22 3.20 3.19 3.19 3.19 3.18 3.18 3.18 4.38 4.36 4.35 4.33 3.12 3.18 3.19 3.21 3.73 3.72 3.67 3.65 3.50 3.48 3.44 3.47 4.77 4.76 4.74 6.17 4.25 4.24 2.69 2.69 2.69 2.69 2.70 2.70 2.70 4.03 4.04 4.05 3.79 4.12 4.44 4.49 4.59 8.96 8.87 6.05 5.79 5.75 3.49 3.30 8.87 5.77 5.40 5.22 5.04 4.86 13.40 13.30 13.00 12.50 12.30 12.20 5.62 5.37 5.13 4.88 12.30 12.30 12.20 12.20 16.60 16.80 16.90 16.90 9.22 9.06 8.37 8.01 4.23 3.54 1.28 5.49 6.92 6.79 6.53 5.36 0.83 0.62 8.41 8.57 8.73 8.90 9.42 9.61 9.79 7.51 7.69 7.86 4.68 10.02 11.84 11.97 12.14 R R R R R R R R* r' r' r' r' r' R R R R R R r' r' r' r' R R R R R R R R R R R* R* R* R* R* R* R R R R R R R R* R* R* R R R R R R R* R* R R R* 29 29 6 18 29 20 34 4 5 7 19 21 29 4 2 8 18 29 16 7 17 17 15 1 4 6 8 6 6 8 8 26 42 22 28 25 18 13 11 14 23 18 18 26 31 8 21 19 34 16 14 36 13 13 12 30 5 18 13 29 NTT-EMMI 220 NTT-EMMI 220 NTT-EMMI 60 WHT-PFIP 70 WHT-PFIP 70 NTT-EMMI 85 NTT-EMMI 85 VLT-FORS2 300 NTT-EFOSC2 100 NTT-EFOSC2 150 NTT-EFOSC2 17x80 , 2x100 NTT-EFOSC2 20x80 , 1x120 NTT-EFOSC2 26x80 , 3x120 WHT-PFIP 150 VLT-FORS2 250 VLT-FORS2 250 NTT-EFOSC2 120 NTT-EFOSC2 120 NTT-EFOSC2 120 INT-WFC 75 INT-WFC 75 INT-WFC 75 INT-WFC 75 NTT-EMMI 750 NTT-EMMI 340 NTT-EMMI 360 NTT-EMMI 400 WHT-PFIP 120 NTT-EFOSC2 100 NTT-EFOSC2 100 NTT-EFOSC2 100 NTT-EFOSC2 160 NTT-EFOSC2 10x250, 32x180 VLT-FORS2 21x70, 1x40 VLT-FORS2 70 VLT-FORS2 70 VLT-FORS2 70 VLT-FORS2 70 VLT-FORS2 70 NTT-EMMI 150 NTT-EMMI 110 NTT-EMMI 200 140 NTT-EMMI 1x14, 1x30, 24x75 NTT-EMMI NTT-EMMI 1x15, 30x75 WHT-PFIP 60 VLT-FORS2 3x130, 18x80 VLT-FORS2 4x110, 15x80 80 VLT-FORS2 NTT-EFOSC2 60 NTT-EFOSC2 60 NTT-EFOSC2 60 WHT-PFIP 90 3x90, 10x110 WHT-PFIP WHT-PFIP 90 VLT-FORS2 60 VLT-FORS2 60 NTT-EFOSC2 180 300 NTT-EFOSC2 VLT-FORS2 60 072.C-0233(A) 072.C-0233(A) 079.C-0297(A) W/2007A/20 W/2007A/20 074.C-0125(A) 074.C-0125(A) 077.C-0609(B) 194.C-0207(C) 194.C-0207(C) 194.C-0207(C) 194.C-0207(C) 194.C-0207(C) W/2008B/23 082.C-0517(B) 082.C-0517(B) 082.C-0517(A) 082.C-0517(A) 082.C-0517(A) I/2005A/11 I/2005A/11 I/2005A/11 I/2005A/11 079.C-0297(B) 079.C-0297(B) 079.C-0297(B) 079.C-0297(B) W/2008B/23 082.C-0517(A) 082.C-0517(A) 082.C-0517(A) 089.C-0372(A) 089.C-0372(A) 089.C-0372(B) 089.C-0372(B) 089.C-0372(B) 089.C-0372(B) 089.C-0372(B) 089.C-0372(B) 079.C-0297(B) 079.C-0297(B) 079.C-0297(B) 074.C-0125(A) 079.C-0297(A) 079.C-0297(A) W/2008B/23 082.C-0517(B) 082.C-0517(B) 082.C-0517(B) 082.C-0517(A) 082.C-0517(A) 082.C-0517(A) W/2007A/20 W/2007A/20 W/2007A/20 089.C-0372(B) 089.C-0372(B) 089.C-0372(A) 089.C-0372(A) 089.C-0372(B) a Superscripts I and O indicate whether the comet is inbound (pre-perihelion) or outbound (post-perihelion). * ESO R SPECIAL+76 filter with effective wavelength 655 nm and and FWHM 165.0 nm. MNRAS 000, 1 -- 38 (2017) 12 R. Kokotanekova et al. servations. The WFC is a mosaic of four thinned EEV 2048 × 4096 pixel CCDs. Only CCD3 was used for collecting the 94P time series. It has an effective field of view of 11.5 × 23 arcmin2 and the pixel scale of the instrument is 0.33 arc- sec pixel-1. All observations were done through an SDSS r' filter. 3.3 Data reduction To ensure compatibility, the same reduction routine was fol- lowed consistently for each individual dataset. We performed the data reduction using standard IRAF tasks (Tody 1986, 1993) implemented on PyRAF3. A master bias frame for each night was created by using 9-19 individual bias frames. The master bias frame was then subtracted from each frame. If at least five twilight sky flats for the corresponding night were taken, the normalised sky flats were median combined. Since all used instruments have demonstrated stable night- to-night flat fields, in some cases the same flat field was used for more than one night. This was done only when there were no sky flats available for some of the nights within the same run. In the cases when no sky flats were obtained within 2 nights of the observations, dome flats were used. All science images were flat-field corrected by division to the median- combined flat field of the corresponding night. The R-band images affected by fringing were corrected using the IRAF script provided by Snodgrass & Carry (2013). 3.4 Data analysis In an attempt to expand the sample of comets with known rotation rates, we had to analyse archival data sets taken during different observing runs which belong to different scientific programs. This posed the challenge of combining data from different instruments and different observing ge- ometries. In order to be able to reconcile all observations, we developed a robust method for absolute photometric cal- ibration which uses the Pan-STARRS1 (PS1) survey (Cham- bers et al. 2016). The main advantage of this method is that on each frame the comet is compared to numerous neigh- bouring stars with precisely measured PS1 magnitudes. This provides the opportunity to calibrate absolutely the comet's magnitude even in non-photometric conditions, and allows absolute photometric calibration with uncertainties as low as 0.02 mag. 3.4.1 Selecting comparison stars The first step of our photometric calibration procedure was to identify comparison stars on the science frames. For each observing night, the comet brightness variation was deter- mined with respect to a number of rigorously selected neigh- bouring stars. The selected stars had to be present on all comet frames for the corresponding night, so that we could measure the comet variation with respect to each compar- ison star throughout the night. We ensured that no stars located in bad sections of the CCDs were used. In order to avoid vignetting effects, all stars close to the edges of the frames were excluded, taking care that the specific limits of each instrument were respected. All stars used were taken from the Pan-STARRS PS1 Data Release 14 (DR1) archive which was publicly released on 16 December 2016 (Kaiser et al. 2002, 2010; Chambers et al. 2016, and references therein). PS1 used a 1.4 Gigapixel camera mounted on a 1.8 metre telescope to complete a 3π steradian survey of the sky in five broadband filters (gP1, rP1, iP1, zP1, yP1). The PS1 filter system is slightly different from SDSS, and the magnitudes from the two systems can be converted using the equations presented in Tonry et al. (2012). The catalogue stars were matched to the objects on our science frames after the WCS system of each frame was fixed using WCSTOOLS5. This was done in order to maximise the number of PS1 stars identified on the frames. The survey provides positions and magnitudes of both stars and extended objects. To distinguish between them, we followed the PS1 DR1 guidelines for star-galaxy separation. A careful comparison of the PSF of the selected PS1 ob- jects identified on FORS2 images confirmed that indeed the selected catalogue objects corresponded to objects with stel- lar profiles on the frames. This study of the 8.2m VLT tele- scope data allowed excellent identification of non-stellar pro- files and gave us confidence that very few galaxies should be contaminating our selected comparison stars. Even if some galaxies were left in the list of selected catalogue objects, their influence would become negligible due to the large to- tal number of comparison stars per frame (typically > 20). To ensure that the photometric calibration is dominated by good comparison stars, we applied two additional criteria for selecting PS1 stars. We removed PS1 entries with uncer- tainties in the rP1, magnitude larger than 0.08 mag and used stars with colours gP1-rP1 < 1.5 mag. 3.4.2 Photometry To measure the frame magnitudes of the comet and the selected comparison stars, we performed circular aperture photometry. All measurements were done using the IRAF packages DIGIPHOT and APPHOT (Davis 1999). The observations were taken with telescope tracking at sidereal rate. Exposure times were generally short enough so that the apparent motion of the comet would be less than 0.5"-0.6" and the comet would thus remain within the seeing disk. The few frames which did not fulfil this criterion were excluded from the analysis below. Having stellar profiles for both the comet and the background comparison stars guar- anteed that the adopted circular aperture photometry pro- cedures allowed direct comparison with the catalogue mag- nitudes of the stars. The aperture radius used to measure the brightness of the comet nucleus was set equal (within the nearest integer pixel) to the full width at half maximum (FWHM) of the stellar point spread function (PSF) for each frame. This ap- proach was previously found to be optimal for maximising the signal-to-noise ratio (S/N; e.g. Howell 1989). This was 3 http://www.stsci.edu/institute/software_hardware/pyraf 4 http://panstarrs.stsci.edu 5 http://tdc-www.harvard.edu/wcstools/ MNRAS 000, 1 -- 38 (2017) also beneficial for slightly more crowded sky fields, as it de- creased the probability that light from neighbouring stars influences the measured brightness. To find the FWHM of the stellar PSF on each frame, we used the IRAF routine PSFMEASURE. The value for each frame was determined using the median of the measured FWHM of the best fit Gaussian profile to each of the selected comparison stars. The motion of the comet on the sky over the course of the observing night can be non-linear. Therefore instead of using the position of the comet predicted from its ephemeris, we determined the centre of the comet on each frame inter- actively using the IRAF task IMEXAMINE. The main purpose of the analysis is to derive the bright- ness variation of the comet during the individual nights, and subsequently to combine all data points into a common lightcurve. This is best achieved by first deriving a differen- tial lightcurve of the comet with respect to the comparison stars for each night. Then, the lightcurves from the separate nights can be calibrated absolutely by shifting all points by a factor derived from the absolute calibration of just one refer- ence frame for each night. Taking the differential magnitude of the comet rather than absolutely calibrating each frame is a better approach since the brightness variation within each night is independent of the absolute calibration uncertainty. The differential photometry was implemented as fol- lows. First, once the magnitudes of the comet and the stars were determined, we calculated the differences between the comet magnitude and each star, i (∆mcomet,i = mcomet−mi). We also determined the difference in brightness between each star and the brightest non-saturated star (∆m∗,i = mi − m∗). The brightest star was selected because it had the highest S/N. Then, we scaled the difference of the comet and each star with ∆m∗,i (∆mframe,i = ∆mcomet,i − ∆m∗,i). Finally, the differential photometry magnitude of the comet with respect to the brightest star, mcomet,diff, was calculated as the me- dian of ∆mframe,i. Its uncertainty was estimated from the median absolute deviation of ∆mframe,i. 3.4.3 Absolute calibration A key aspect of our method is the absolute calibration of comet magnitudes using stars from the PS1 catalogue. This procedure allows us to combine data from different observing runs with smaller systematic uncertainties than traditional absolute calibration methods (e.g. using Landolt stars). In order to convert the relative magnitudes of the comet to standard magnitudes, we need to derive a correction fac- tor for each night. There are two main factors we need to take into account while deriving the conversion: 1) the colour term of the instrument set up (CCD chip and filter) with re- spect to the star catalogue (PS1), and 2) the zero point for each night. The colour term for each of the set ups was determined from comparison between the frame magnitudes and the PS1 magnitudes of 500-1500 stars in total. For each observing night, we chose the frame with the best seeing as a refer- ence frame. The frame magnitudes of the comparison stars on the reference frame (Rframe) were then compared to the corresponding PS1 rP1 and gP1 magnitudes. After PS1 stars with extreme colour indices (gP1 - rP1 > 1.5 mag) were ex- cluded, the differences Rframe - rP1 were plotted versus the MNRAS 000, 1 -- 38 (2017) Rotation of Cometary Nuclei 13 Figure 1. Colour term of the red arm of NTT-EMMI used with a Bessel R filter. The scaled difference of the measured R magni- tudes and the PS1 rP1 magnitudes of the comparison stars from all used datasets are plotted against their PS1 (gP1-rP1) colour indices. The orange line indicates the best linear fit to all points and its slope corresponds to the colour term. The colour term is used to correct the magnitude of the comparison stars before finding the zero point of each frame taken with that instrument configuration. Table 3. Derived colour terms for all instruments used in this work Instrument Filter ca NTT-EMMI R NTT-EFOSC R NTT-EFOSC r' VLT-FORS2 WHT-PFIP INT-WFC R** R r' -0.117 -0.158 -0.194 -0.071 -0.100 -0.007 σcb 0.005 0.012 0.005 0.006 0.008 0.002 (gP1- rP1) rangec 0.0 - 1.0 0.4 - 1.5* 0.0 - 1.5 0.0 - 1.0 0.0 - 1.0 0.0 - 1.5 a Colour term c derived from comparison with PS1 star magni- tudes in rP1 and gP1 b Uncertainty in the colour term c Range of the PS1 gP1- rP1 colour indices of the used stars. Colour indices < 1 for Johnson-Cousins R filters and < 1.5 for SDSS r fil- ters * This range was selected due to an insufficient number of stars with gP1- rP1< 0.4 in the observations used in this paper. ** ESO R SPECIAL+76 filter with effective wavelength 655 nm and and FWHM 165.0 nm. colour indices of the stars. All points were scaled so that the median of Rframe - rP1 was brought to 0 mag. After this was done for all observed fields, all points were combined into a common plot such as Fig. 1. The colour term of the instrument was determined by taking the slope of the best fitting linear function. The derived colour indices of each instrument and their uncertainties are presented in Table 3. In order to use the colour term, we need to know the comet's colour index. The surface colours of JFCs are rela- tively well constrained with average colour indices (V - R) = 0.50 ± 0.03 and (B - V) = 0.87 ± 0.05 (Lamy et al. 2009). The (V - R) colour index can be converted to SDSS filter system: (g' - r') = 0.67 ± 0.06 using the relations in Jester et al. (2005) and to (gP1- rP1) = 0.58 ± 0.06 in the PS1 sys- tem (Tonry et al. 2012). Since no further colour information was available for most comets, this colour index was used for the absolute calibration throughout the analysis. 0.00.20.40.60.81.0gP1 - rP1 [mag]0.40.30.20.10.00.10.20.30.4Scaled RFrame - rP1 [mag] 14 R. Kokotanekova et al. The next step was to find the zero point of the refer- ence frame from the difference between the colour-corrected frame magnitudes and the corresponding PS1 rP1 magni- tudes. With the colour term and the zero point of the refer- ence frame at hand, we converted the comet's magnitude to the PS1 system. Once the comet magnitude on the reference frame was converted to PS1 rP1 magnitudes, we shifted all the relative magnitudes to produce an absolutely-calibrated lightcurve of the comet for each night. 3.4.4 Observing geometry correction The absolutely calibrated lightcurves from each night had to be corrected for viewing geometry effects. Firstly, we cor- rected each time series for light-travel time, converting "ob- servation times" to "times when the light left the nucleus". The next step was to convert the absolutely calibrated frame magnitudes, mr, to absolute magnitudes, mr(1, 1, 0). The comet magnitude, mr, depends on the observing geom- etry of the comet. It is given by : mr = Hr + 5 log(Rh∆) + βα, (2) where Hr = mr(1, 1, 0) is the hypothetical absolute magnitude of the comet nucleus measured at an imaginary point at heliocentric distance Rh = 1 au; geocentric distance ∆ = 1 au and phase angle α = 0◦. This equation is valid for objects whose phase functions don't show an opposition surge and can be described by a linear fit with slope β. In the case of JFCs, a linear model with β of 0.035 mag/deg is generally accepted (e.g Lowry & Fitzsimmons 2001; Snodgrass et al. 2005). For most comets we have data from different epochs, which could be used to derive a phase function slope β in- dependently. In all other cases, where the observations cov- ered phase angle ranges smaller than 2◦, we used β of 0.04 mag/deg to find the absolute magnitude of the nucleus Hr. For such single-run observations, we used the frame magni- tude mr, rather than Hr, to derive the lightcurves. 3.4.5 Checking for activity To determine whether the comets were active at the time of the observations, we compared the average comet PSF profile to that of a star. We first median-combined all sky- subtracted images for the night to produce a deep image of the background stars without cosmic rays and the moving comet. We then scaled this image and subtracted it from each comet frame in order to remove the background stars. Next, we centred each difference frame on the comet and combined all frames using a median filter, removing all cos- mic rays. Finally, the measured comet profile on the com- bined frame was compared to the PSF of a bright star mea- sured on the combined star field image. In some cases described in detail below, the comet pro- file was noticeably different from that of the comparison stars (see. Sections 4.2 and 4.3). This was interpreted as a strong indication of activity around the time of the ob- servations. Nevertheless, we attempted to use these datasets to estimate the rotation rates and the properties of the nu- clei. However, the derived results need to be interpreted with caution. 3.4.6 Period search We used the Lomb-Scargle method (LS; Lomb 1976; Scar- gle 1982) to detect periodicities in the brightness variation of the observed nuclei. LS is among the most widely used meth- ods for finding periods in unevenly-sampled time series. We ran the python gatspy6 LombScargleFast implementation of LS (VanderPlas & Ivezic 2015) to look for periods be- tween 3 and 40 hours. In the cases where the nightly bright- ness variations suggested slower rotation, we extended the range to cover larger periods. Since we sampled a large range of possible periods, we computed the periodogram with the option of LombScargleFast to automatically determine the period grid. This guaranteed that the longer periods are as well-sampled as the shorter ones. We assumed that the brightness variation of the comets is a result of their shape rather than surface albedo varia- tions. As the lightcurves of elongated bodies have two min- ima and two maxima per rotation cycle, we focused our search on double-peaked lightcurves. Experience shows that Lomb-Scargle periodograms preferentially fit single-peaked lightcurves. Therefore, we interpreted the derived peaks in the periodograms as half the rotation period of comets. The LS periods were cross-checked using two other methods for detecting periods of unevenly spaced samples: phase dispersion minimization7 (PDM; Stellingwerf 1978) and string-length minimization (SLM; Dworetsky 1983). For all comets below, the three methods detected the same set of possible periods and showed general agreement. There- fore, for simplicity, we have chosen to show only the LS periodograms. 3.4.7 Nucleus size and shape and density estimates We used the lightcurves we derived to set constraints on the sizes, shapes and albedos of the observed nuclei. The mean apparent magnitude of the comet (mr) and the mean abso- lute magnitude (Hr) were calculated as the arithmetic mean of all magnitudes mr and Hr. The uncertainty we report cor- responds to the median of the uncertainties of all individual points. The mean absolute magnitude can be converted to an average radius for the nucleus in kilometres using: rN = (k /(cid:112)Ar) × 100.2(m(cid:12)−Hr) , (3) where k = 1.496 × 108 km is the conversion factor between au and km; Ar is the geometric albedo of the comet and m(cid:12) = −27.08 mag is the apparent magnitude of the Sun, both in PS1 rP1-band. We used the commonly assumed geometric albedo value for comets of Ar = 0.04. The reported uncertainties on the radii are based only on the photometric uncertainty. They do not account for the uncertainties introduced by the albedo and the phase func- tion slope. The albedos of JFCs are between 2-7 percent (see Table 5), which is within a factor of 2 of the commonly as- √ sumed value of 4 percent. Therefore, the radius estimate can 2 from the reported value. Since we vary with maximum observed all comets in a narrow phase angle range (typically < 10 deg), the influence of the phase function uncertainty 6 http://www.astroml.org/gatspy/ 7 https://github.com/sczesla/PyAstronomy MNRAS 000, 1 -- 38 (2017) is also small. In the worst case, if the phase function slope varies with up to 0.08 mag/deg, the absolute magnitude of the comet will vary with 0.8 mag, and the estimated radius will be within a factor of 1.5 from the estimated value. Eight of the comets have SEPPCoN thermal measure- ments of the radii. We can use our absolute magnitudes Hr and the SEPPCoN effective radii Reff to derive their geomet- ric albedos using: Ar = (k2 / R2 eff) × 100.4(m(cid:12)−Hr) . (4) The peak-to-peak variation ∆Hr can also be used to set a lower limit on the elongation of the comet nucleus. We determined ∆Hr by taking the observed range of magnitudes of the corresponding dataset. If the nucleus is modelled as a prolate ellipsoid with semi-axes a,b and c, where b = c and a > b, the axis ratio a/b can be determined by a b Since we do not know the orientation of the rotational axis of any of the considered nuclei, we can only measure the pro- jection of the axis ratio onto the plane of the sky. Therefore, Eq. 5 provides only a lower limit of the elongation. ≥ 100.4∆Hr . (5) We can also place a lower limit on the bulk density of the comets by combining the derived rotation periods (Prot) in hours and axis ratios (a/b). For a strengthless body, the nucleus density (DN) must be sufficient to prevent rotational break up due to centrifugal forces. In units of g cm−3 this constraint can be approximated to: DN ≥ 10.9 P2 rot a b, (6) where the period is given in hours (Pravec & Harris 2000). 3.4.8 Monte Carlo method Determining the uncertainty in the lightcurve period is a challenging and often neglected task. In this work, that prob- lem is often additionally complicated by the large time span between the different observations, which leads to aliases in the periodograms. Additionally, as is shown for the individ- ual comets below, sometimes more than one period seems to characterise the variation well, and it is not possible to decide on the most likely spin rate. In such cases, providing an uncertainty in the determined period based just on the information on the periodogram (e.g. FWHM of the highest peak) can be misleading. Moreover, it is not clear to what extent the detected periods are influenced by the intrinsic uncertainties of the comet magnitudes. Two main effects are at play when con- sidering what might dominate the uncertainties of the avail- able time series. Firstly, the data from the different nights are linked using absolute calibration. In some cases the sky area under consideration has few stars, which increases the absolute calibration uncertainty. Second, when we combine observations from two different observing runs, the applied phase angle correction determines the relative difference be- tween the comet magnitudes from the different epochs. This effect is hard to quantify, unless the influence of the different possible phase function correction parameters is explored. In an attempt to account for these effects, we adopt MNRAS 000, 1 -- 38 (2017) Rotation of Cometary Nuclei 15 a Monte Carlo method which allows us to retrieve better- validated values for the phase function coefficients and the rotation periods for the comets. The Monte Carlo method consists of the following steps: (i) Each magnitude from the time series of the comet is re- placed by another randomly selected value. The new magni- tude is selected from a normal distribution with mean equal to the original magnitude value and standard deviation equal to the uncertainty of the magnitude. The result is a clone i of the original time series, where the times and observing geometries are the same as the original time series, but the magnitudes were varied within the uncertainty space. (ii) The clone magnitudes are used to find the best fitting linear phase function coefficient βi. (iii) The clone data set is corrected for the phase function by converting from m(1, 1, α) to m(1, 1, 0) using the derived βi. (iv) The Lomb-Scargle period search routine is run on the clone magnitudes m(1, 1, 0) to determine the best-fitting period Pi. (v) This procedure is repeated for i = 1, 2, . . . , 5000. (vi) To determine the phase function coefficient, we plot the histogram of the determined βi and fit a gaussian prob- ability density function to it. In the final results, we report the best fit for the phase function coefficient to be the mean of the distribution, while its uncertainty is taken to be equal to the central 3σ range of the distribution. (vii) To determine the most likely rotation period, we plot the histogram of the derived Pi and fit a gaussian probability density function to it. As a final result we report the period of the comet as equal to the mean of the distribution, and an uncertainty equal to the central 3σ range of the fitted probability density function. In all cases the distribution of the derived βi can be described well by a normal distribution. However, for some comets the Pi distributions are more irregular. In the cases when the distribution is irregular, we take the highest peak as the most-likely period candidate, but we carefully explore the alternatives in the analysis. 4 TIME SERIES PHOTOMETRY RESULTS 4.1 14P/Wolf The lightcurve of comet 14P/Wolf was first determined from 2 observing nights close to aphelion in 2004 by Snodgrass et al. (2005). Our team observed 14P as part of SEPPCoN once more in 2007 during the same aphelion passage. We analysed both datasets with our method for absolute photo- metric calibration and combined them in order to constrain better the comet's rotational period. We used the procedure described in Section 3.4.5 to check whether 14P was active during the observations in 2004. The comet appears stellar in the co-added comet com- posite image and its surface brightness profile is indistin- guishable from that of the comparison star (Fig. 2). This confirms the conclusion of Snodgrass et al. (2005) that 14P was not active during the observations in 2004. Figure 3 shows the Lomb-Scargle periodogram for the 2004 observations of 14P. The highest peak is at Pfit = 4.46 hours, corresponding to a rotation period Prot = 8.93 hours 16 R. Kokotanekova et al. Figure 2. Surface brightness profile of 14P from the 2004 data set. The lower panel shows a 30 × 30 arcseconds composite image of 14P made up of 29 × 220 s exposures taken on 21 January 2004. The frames are added in a method which removes cosmic rays, the background sky and fixed objects. The comet appears stellar and no signatures of activity can be recognised. The surface brightness of the comet is plotted as a function of radius ρ from the centre of the comet. The profile matches the scaled stellar PSF (solid line), indicating that the comet appears as a point source and is therefore considered to be inactive. (Fig. 4). Using the Monte Carlo method without phase func- tion correction, we determined that the best-fitting rotation period is Prot = 8.93 ± 0.04 hours (Fig. 5). Using the same dataset, Snodgrass et al. (2005) iden- tified 7.53 ± 0.10 hours as the most likely rotation period of 14P. That period corresponds to the third highest peak in our periodogram and results in an unusual asymmetric lightcurve. The difference in the periods likely originates from the different methods for night-to-night calibration adopted in the two works. While Snodgrass et al. (2011) used Landolt star calibration, here we applied our newly devel- oped method for absolute calibration with PS1, which allows precise absolute calibration independent of the changing ob- serving conditions during the night. Thus, by re-analysing the data from 2004 with our method, we improved the pe- riod determination of 14P. The lightcurve of 14P in 2004 phased with Prot = 8.93 ± 0.04 hours has a peak-to-peak brightness variation of ∆mr = 0.36 ± 0.05 mag, which corresponds to axis ratio a/b ≥ 1.39 ± 0.06. From Eq. 6 we estimated a minimum nucleus density of 0.19 ± 0.04 g cm−3. Next, we analysed the observations from 2007. The comet appears stellar on the composite images and its sur- face brightness profile does not deviate from that of the com- parison star (Fig. 6). We can therefore assume that 14P was inactive at the time of the observations. The highest peak of the Lomb-Scargle periodogram for the 2007 observations is at Pfit = 4.51 hours corresponding to a rotation period Prot = 9.02 hours. (Fig. 7). We used the Monte Carlo approach without geometric corrections to determine a rotation rate Prot = 9.02 ± 0.04 hours (right panel on Fig. 5). The lightcurve phased with the identified period (Fig. 8) has a peak-to-peak variation ∆mr = 0.39 ± 0.05 mag corresponding to a/b ≥ 1.43 ± 0.07 and DN ≥ 0.19 ± 0.04 g cm−3. The periods from 2004 and 2007, around the same aphe- lion passage, are compatible within the uncertainties. Fur- Figure 3. Lomb-Scargle periodogram for 14P from the dataset collected in 2004. The plot shows the LS power versus period. The highest peak occurs at 4.46 hours, which corresponds to the most likely period Prot = 8.93 hours. Figure 4. Rotational lightcurve of 14P with the data from 2004. The lightcurve is folded with the LS best period of 8.93 hours. Figure 5. Results from the Monte Carlo simulations used to determine the rotation period of 14P from the datasets in 2004 (left) and 2007 (right). The resulting rotation periods for 2004 and 2007 are 8.93 ± 0.04 and 9.02 ± 0.04 respectively. thermore, the fact that the comet was inactive at both epochs suggests that 14P probably remained inactive around aphelion and a period change due to outgassing is unlikely to have occurred. Since we have no knowledge about the comet spin axis orientation, it is not possible to exclude the possi- bility that the viewing geometry changed between the two MNRAS 000, 1 -- 38 (2017) 0.112345ρ [arcsec]222426283032Surface brightness [mag arcsec-2]5101520Period [hours]0.00.10.20.30.40.50.60.70.80.9Lomb-Scargle Power0.00.20.40.60.81.0Rotational Phase22.222.422.622.8mr [mag]20/1/0421/1/048.908.928.948.96Period [hours]020Frequency20049.009.029.049.06Period [hours]02040602007 Rotation of Cometary Nuclei 17 Figure 6. Same as Fig. 2, for 14P on 18 May 2007. The co-added composite image of 14P was made up of 18 × 70 s exposures. The stellar appearance on the composite image and the surface bright- ness profile of the comet suggest that 14P was inactive during the observations in 2007. Figure 9. Monte Carlo simulation results for the phase function and the rotation period of 14P for the combined dataset from 2004 and 2007. The determined linear phase function slope is β = 0.060 ± 0.005 (left) and the rotation period is Prot = 9.02 ± 0.01 hours (right). Figure 7. Lomb-Scargle periodogram for 14P with the dataset from 2007. The highest peak corresponds a period Prot = 9.02 hours. Figure 8. Rotational lightcurve of 14P with the data from 2007. The lightcurve is folded with period 9.02 hours. Figure 10. Phase function of comet 14P. The absolutely cali- brated comet magnitudes corrected for heliocentric and geocen- tric distance are plotted versus phase angle α. The linear phase function with the best-fitting slope β = 0.060 ± 0.005 mag deg-1 is plotted as a solid line. two datasets in order to determine a phase function and a common rotation period. We ran the Monte Carlo simulation on the combined dataset and determined a phase function slope β = 0.060 ± 0.005 mag/deg and period Prot = 9.02 ± 0.01 hours (Fig. 9, 10). The Lomb-Scargle periodogram of the combined datasets (Fig. 11) has a pronounced peak at Pfit = 4.51 hours which corresponds to the best period from the Monte Carlo simulation. The lightcurve phased with the best period Prot = 9.02 (Fig. 12) has a range ∆Hr = 0.37 ± 0.05 mag corresponding to a/b ≥ 1.41 ± 0.06 and DN ≥ 0.19 ± 0.03 g cm−3. The mean absolute magnitude was Hr(1,1,0) = 14.87 ± 0.05 mag. Using eq. 4 and the radius from Fern´andez et al. (2013), we estimated the comet's albedo to be Ar = 4.3±0.6%. epochs. However, both individual lightcurves have the same peak-to-peak brightness variation (within the corresponding uncertainties), and therefore we can assume that the change in geometry did not influence the observed lightcurve. With these assumptions at hand, we proceeded to combine the 4.2 47P/Ashbrook-Jackson The first attempt to determine the rotation rate of 47P was made by Snodgrass et al. (2006) using data from two observ- ing nights in 2005. However, the resulting time series were not sufficient to choose between four possible periods: 11.2, MNRAS 000, 1 -- 38 (2017) 0.112345ρ [arcsec]222426283032Surface brightness [mag arcsec-2]5101520Period [hours]0.00.10.20.30.40.50.60.70.80.9Lomb-Scargle Power0.00.20.40.60.81.0Rotational Phase20.620.821.021.221.4mr [mag]15/05/200718/05/200719/05/20070.0550.0600.065β [mag deg-1]0200Frequency9.019.029.03Period [hours]01005.56.06.57.07.58.08.59.09.5α [deg]15.015.215.415.6Hr(1,1,α) [mag]20/1/0421/1/0415/5/0718/5/0719/5/07 18 R. Kokotanekova et al. Figure 11. Lomb-Scargle periodogram of 14P with the combined datasets from 2004 and 2007. The highest peak corresponds to the most likely period Prot = 9.02 hours. The periodogram is very densely populated with peaks from the aliases which are present due to the large time span between the two observing runs. Figure 13. Same as Fig. 2, for 47P on 6 March 2005. The co- added composite image of 47P is made up of 27 × 85 s expo- sures. The surface brightness profile of the comet slightly deviates from the stellar one beyond 2 arcseconds, which suggests that the comet was weakly active during the time of the observations. Figure 12. Rotational lightcurve of 14P/Wolf with the data from 2004 and 2007. The lightcurve is folded with period 9.02 hours. 15.5, 21.6 and 44 hours. Moreover, as discussed in Section 2.2.2, the attempts to determine the comet's phase function have also remained unconsolidated (Snodgrass et al. 2008b; Lamy et al. 2011). In order to address these inconsistencies, we obtained new time-series observations of the comet in April 2015. The new data were taken at a different apparition than those from 2005, and could not be used to look for a common pe- riod without introducing further uncertainties. Nevertheless, the two datasets could still be combined for an attempt to derive the phase function of the nucleus. Upon re-analysing the 2005 data set, we found that 47P was faintly active during the observing run. However, the inner surface brightness profile of the coma matched that of the comparison star well, suggesting that the activity was clearly weak (Fig. 13). We re-analysed the data from 2005 using our new absolute-photometry calibration method. The PS1 night-to- night calibration led to the identification of a smaller bright- ness variation and different possible periods than those in Snodgrass et al. (2006). The two strongest peaks of our LS periodogram were at Prot,1 = 10.8 and Prot,2 = 14.1 hours (Fig. 14), and it is impossible to choose between them unam- biguously (Fig. 15). The brightness variation of the resulting Figure 14. Lomb-Scargle periodogram of 47P with the data from 2005. The two highest peaks correspond to Prot,1 = 10.8 hours and Prot,2 = 14.1 hours. lightcurve was ∆mr = 0.33 ± 0.06 mag suggesting axis ratio of a/b ≥ 1.36 ± 0.07. When 47P was observed again in 2015, it appeared to be slightly active (Fig. 16). Nevertheless, the new time series showed sufficient brightness variation to enable a rotation period determination. The two highest peaks on the LS pe- riodogram of the 2015 dataset suggested Prot,1 = 15.6 hours or Prot,2 = 23.7 hours (Fig. 17). However, we consider that Prot,2 = 23.7 hours is an alias due to the nightly sampling of the observations. Phasing the lightcurve of the comet with 23.7 hours produced a non-realistic noisy lightcurve, and confirmed that this period does not correspond to the rota- tion rate of 47P. We ran the Monte Carlo simulation for periods between 3 and 23 hours (to avoid the 24-hour alias) and determined Prot = 15.6 ± 0.1 hours. The resulting plots of the MC sim- ulation here and for most objects below are not shown since they are similar to Fig. 9, and do not provide additional in- formation on the simulation outcomes. The brightness vari- ation of the lightcurve (Fig. 18) was ∆mr = 0.24 ± 0.06 mag, suggesting a/b ≥ 1.25 ± 0.07 and DN ≥ 0.06 ± 0.02 g cm−3. Besides deriving the lightcurve of the comet, one of the main aims of the new observations from 2015 was to con- MNRAS 000, 1 -- 38 (2017) 5101520Period [hours]0.00.10.20.30.40.50.60.70.80.9Lomb-Scargle Power0.00.20.40.60.81.0Rotational Phase14.414.614.815.015.2Hr(1,1,0) [mag]20/1/0421/1/0415/5/0718/5/0719/5/070.112345ρ [arcsec]222426283032Surface brightness [mag arcsec-2]5101520Period [hours]0.00.10.20.30.40.50.6Lomb-Scargle Power Rotation of Cometary Nuclei 19 Figure 18. Rotational lightcurve of 47P with the data from 2015. The lightcurve is folded with the period of 15.6 hours derived from the MC method. Figure 15. Rotational lightcurve of 47P with the data from 2005, folded with periods 10.8 hours (top) and 14.1 hours (bottom). It is impossible to select between these two periods. Figure 19. Same as Fig. 2, for 47P on 1 June 2006. The co- added composite image is made up of 4 × 300 s exposures. Due to the small number of frames, the composite image was made without subtraction of the average stellar background in order to avoid artefacts from the comet's slow position change. The comet appears active on the image, and its surface brightness profile deviates from the stellar PSF. strain the phase function of 47P. To address this, we first considered the previous brightness measurements from Li- candro et al. (2000), Lamy et al. (2011) and Snodgrass et al. (2008b). Their magnitude measurements were converted to PS1 magnitudes using the colour indices of 47P (B-V) = 0.78 ± 0.08 and (V-R) = 0.40 ± 0.08 (Lamy et al. 2011), and the conversions from Tonry et al. (2012). Additionally, we attempted to add an archival VLT data set from June 2006 when the comet was close to aphelion. However, these observations could not be used since the comet was clearly active on the frames (Fig. 19). Instead, these data complemented the data set from March 2006 (Snodgrass et al. 2008b), and confirmed that the comet had an outburst around aphelion. To derive the phase function coefficient β, we used the Monte Carlo approach considering only the long time-series from 2005 and 2015. We did not include the other observa- tions where the comet was active, or where the photomet- ric calibration had been done using different methods. The Monte Carlo method resulted in a coefficient β = 0.096 ± 0.004 mag deg-1. The derived phase function appears to be in good agreement with all previous observations (Fig.20), Figure 16. Same as Fig. 2, for 47P on 24 April 2015. The co- added composite image of 47P is made up of 26 × 80 s exposures. The comet appears to be slightly active with a tail detected to the north east. Figure 17. Lomb-Scargle periodogram of 47P with the data from 2015. The two highest peaks correspond to Prot = 23.7 hours and Prot = 15.6 hours, although the period of 23.7 is most likely a 24-hour alias. MNRAS 000, 1 -- 38 (2017) 21.621.822.0mr [mag]05/3/0506/3/050.00.20.40.60.81.0Rotational Phase21.621.822.0mr [mag]05/3/0506/3/050.112345ρ [arcsec]212223242526272829Surface brightness [mag arcsec-2]5101520Period [hours]0.00.10.20.30.40.50.6Lomb-Scargle Power0.00.20.40.60.81.0Rotational Phase20.821.021.221.4mr [mag]19/4/1521/4/1522/4/1523/4/1524/4/150.112345ρ [arcsec]2224262830Surface brightness [mag arcsec-2] 20 R. Kokotanekova et al. Figure 20. Phase function of comet 47P derived from the ob- serving runs in 2005 and 2015. The symbols from 2005 and 2015 correspond to these used on Figs. 15 and 18. The linear phase function slope β determined with the MC method is 0.096 ± 0.004 mag deg-1. Despite being unusually steep, the phase function is consistent with the previous observations of the comet from Li- candro et al. (2000); Snodgrass et al. (2008b); Lamy et al. (2011). However, since the comet was probably active in 2005 and 2015, the derived phase function slope is not conclusive. although it is unusually steep compared to the typical phase function for JFCs (see Table 5). Using that value for β to convert the observed magni- tude, we calculated Hr(1,1,0) = 14.59 ± 0.06 mag. Using the radius from SEPPCoN and Eq. 4, we derived an albedo Ar = 5.0±0.7 %. We interpret these results with caution because of the the slight activity detected on the stacked frames from 2005 and 2015, as well as the unusually steep phase function. If the coma contribution was large and/or the actual nucleus phase function slope was shallower, we would expect the absolute magnitude of 47P to be fainter. In that case, the comet must also have a smaller albedo (Ar ≤ 5.0 %). Similarly, the derived period Prot = 15.6 ± 0.1 hours must also be regarded as uncertain. The comet was found to be active at the time of the observations and therefore the nucleus signal was likely dampened by the present coma making the brightness variation more difficult to detect. Since the periods from both epochs were uncertain due to the limited sampling and the potential activity, we could not search for period changes occurring between 2005 and 2015. 4.3 93P/Lovas 93P/Lovas was observed with three different instruments during six nights in January 2009 as part of SEPPCoN. The observations were taken at heliocentric distance of 3.8 au when 93P was outbound. The composite images of the comet from each night contained traces of activity, and a tail to the west could clearly be resolved on the VLT frames (Fig. 21). Despite the weak activity, the brightness variation in the time series from each night suggested that the nucleus signal could still be detected. The LS periodogram of the combined dataset can be seen in Fig. 22. The strongest peak at ∼ 24 hours does not produce a typical lightcurve and corresponds to a 24-hour alias. From the remaining peaks, those at Prot = Figure 21. Same as Fig. 2, for the VLT observations of 93P on 24 January 2009. Due to the small number of frames, the com- posite image was made without subtraction of the average stellar background in order to avoid artefacts from the comet's slow po- sition change. The co-added composite image is made up of 8 × 250 s exposures. A tail to the west can be clearly distinguished. The comet profile appears stellar close to the centre but deviates from that of the comparison star at larger radii. 18.2 hours and Prot = 13.2 hours result in possible lightcurves (Fig. 23). We used the MC method to look for the best period between 3 and 23 hours (to avoid the aliasing at 24 hours). The simulation resulted in possible periods between 13.1 and 19.7 with the most frequently preferred period of 18.2 hours (29% of the iterations, Fig. 24). It is impossible to deduce the precise spin rate of 93P from these data, but the period can be constrained to the range Prot = 18.2+1.5−5 hours. The brightness variation of 93P is ∆mr = 0.21 ± 0.05 mag and suggests an axis ratio a/b ≥ 1.21 ± 0.06. The mean magnitude of the comet is mr = 21.09 ± 0.05 mag which cor- responds to Hr(1,1,0) = 15.17 ± 0.05 mag, for a typical phase function β = 0.04 mag/deg. Using Eq. 4 and the SEPPCoN radius from Fern´andez et al. (2013), we estimate that the albedo of 93P is Ar = 4.2±0.9 %. Since the comet showed signatures of activity during the time of the observations, the brightness and albedo values we have derived need to be treated as upper limits. If the coma contribution of the frames is significant, the absolute magnitude of the nucleus must be larger, and therefore the resulting albedo must be smaller. In order to derive more certain estimates of the nucleus parameters, the comet needs to be observed at higher heliocentric distances where it is more likely to be inactive. 4.4 94P/Russell In the analysis described here, we attempted to determine the rotation rate of 94P/Russell after combining three data sets from 2005, 2007 and 2009. The observations were taken before and after the same aphelion passage in 2007. The dataset from 2005 was previously used to deter- mine a period of ∼ 33 hours Snodgrass et al. (2008b). We re-processed the data and used our method for absolute cali- bration to combine the observations from the four observing nights in 2005. The surface brightness profile presented in Snodgrass et al. (2008b) suggested that the comet could have been MNRAS 000, 1 -- 38 (2017) 0246810α [deg]14.614.815.015.215.415.6mr(1,1,α) [mag]Li00La11S0801/06/2006β = 0.096 ± 0.0040.112345ρ [arcsec]20222426283032Surface brightness [mag arcsec-2] Rotation of Cometary Nuclei 21 Figure 22. Lomb-Scargle periodogram of 93P showing the LS power versus period. The highest peak corresponds to a 24-hour alias. The next three peaks correspond to Prot = 18.2, 13.2 and 15.8 hours. Figure 24. Monte Carlo simulation results for the rotation pe- riod of 93P. The most frequently preferred rotation period is 18.2 hours, but the large range of possible periods does not allow us to uniquely determine the rotation rate of the comet. Figure 23. Rotational lightcurve of 93P folded with the two most likely periods 18.2 hours (top) and 13.2 hours (bottom). The dashed line corresponds to second-order Fourier series which aim to reproduce an asymmetric double-peaked lightcurve. The lightcurve phased with 13.2 hours shows less scatter, but the data are not sufficient to discriminate between the two periods. weakly active at the time of the observations. We performed a careful background subtraction of the comet composite images for each night, and concluded that 94P appeared stellar on each night of the run (see Fig. 25). The Lomb-Scargle periodogram of the data taken in 2005 has two strong peaks corresponding to 20.43 and 14.31 hours (Fig. 26). The lightcurves phased with these periods are plotted in Fig. 27. It is not possible to reject the second- best period based on the appearance of the lightcurve. How- ever, in all iterations of the MC simulation the larger period was preferred and therefore the period was determined to be Prot = 20.43 ± 0.05 hours. The resulting lightcurve had a brightness variation ∆mr = 0.7 ± 0.1 mag. This corresponds to an axis ratio a/b ≥ 1.9 ± 0.2 and density DN ≥ 0.05 ± 0.01 g cm−3 . The data taken during the SEPPCoN runs in 2007 and MNRAS 000, 1 -- 38 (2017) Figure 25. Same as Fig. 2, for the observations of 94P in 2005. The co-added composite image is made up of 15 × 75 s exposures taken on 7 July 2005. 2009 were also checked for the presence of activity (Fig. 28 and 29). Due to the faintness of the comet, in both cases its surface brightness profiles levelled out within 5 arcseconds from the nucleus. However, we can conclude that 94P was inactive in both epochs considering the good matches with the stellar PSF close to the centre, as well as the appearance of the composite images. Neither of the two datasets from 2007 and 2009 were sufficient to derive the rotation rate of 94P independently. We therefore only used them to estimate the nucleus magni- tude and the minimum brightness variation at each epoch. We measured mr = 22.6 ± 0.2 and ∆mr = 1.0 ± 0.2 mag for 2007, and mr = 21.30 ± 0.05 and ∆mr = 0.80 ± 0.05 mag for 2009. We combined all three datasets to determine the pre- cise rotation rate of the comet. The analysis of the joined datasets was done under the following assumptions: 1) the comet was inactive during all observations and the measured magnitudes had no coma contributions; 2) the rotation pe- riod remained constant during the entire aphelion passage, and 3) the changing viewing geometry between the different observations did not affect the lightcurve shape significantly. With these assumptions in mind, we used the MC method to derive a phase function with a slope β = 0.039 5101520Period [hours]0.00.10.20.30.40.50.60.70.8Lomb-Scargle Power20.821.021.2Hr [mag]21/1/0922/1/0924/1/0927/1/0928/1/0929/1/090.00.20.40.60.81.0Rotational Phase20.821.021.2Hr [mag]1214161820Period [hours]0.00.10.20.30.40.50.60.70.80.91.0Frequency0.112345ρ [arcsec]222426283032Surface brightness [mag arcsec-2] 22 R. Kokotanekova et al. Figure 26. Lomb-Scargle periodogram of 94P from the dataset taken in 2005. The highest peaks correspond to the most likely periods Prot,1 = 20.43 hours and Prot,2 = 14.31 hours. Figure 28. Same as Fig. 2, for the observations of 94P in 2007. The co-added composite image is made up of 8 × 400 s exposures taken on 20 July 2007. Figure 27. Rotational lightcurve of 94P from the data obtained in 2005. The lightcurve is folded with Prot,1 = 20.43 hours (top) and Prot,2 = 14.31 hours (bottom). We cannot choose between the two periods based on the appearance of the two lightcurves. However, Prot,1 = 20.43 hours is preferred by the MC method, and is therefore considered as more likely. ± 0.002 mag deg-1 (Fig. 30). The LS periodogram of the combined dataset on Fig. 31 peaks at Prot = 20.70 hours. The period Prot = 20.70 hours was also suggested by PDM and SLM. The other two peaks of the LS periodogram close to 38 and 40 hours were also inspected but their lightcurves were significantly noisier. The period of 20.70 hours was preferred in 86% of the MC iterations, which allows us to set the rotation rate of 94P to Prot = 20.70 ± 0.07 hours. The corresponding lightcurve plotted in Fig. 32 shows a very good agreement between the separate datasets. The absolute magnitude of 94P from the combined dataset was Hr(1,1,0) = 15.50 ± 0.09 mag. The albedo of 94P was determined with Eq.4 to be Ar = 4.0±0.6 %. The only data sets which deviate from the first-order Fourier series in Fig. 32 are the ones from July 2007. These points are fainter than the comet magnitude from the rest of Figure 29. Same as Fig. 2, for the observations of 94P in 2009. The co-added composite image is made up of 8 × 100 s exposures taken on 28 January 2009. the nights. There were no indications of problems with the images or the photometric calibration during these nights. We can conclude that the lightcurve must be asymmetric, with one of the minima being sharper and deeper than the other one. Such a lightcurve would have ∆mr = 1.11 ± 0.09 mag which corresponds to a/b ≥ 2.8 ± 0.2 and density DN ≥ 0.07 ± 0.02 g cm−3. Another effect which could produce the observed lightcurve is the change of viewing geometry. Comet 94P moved approximately 120◦ along its orbit between 2005 and 2009, which could be sufficient to produce a noticeable varia- tion in the total surface area of the nucleus for an observer on Earth. Alternatively, the shift in brightness might be caused by weak activity in the 2005 and 2009 data when the comet was closer to the Sun. Such activity is not evident in the profiles on Figs. 25, 28 and 29 but it is possible for some weak activity to be hidden within the seeing disc of distant comets (e.g. Snodgrass et al. 2016). With the limited data here, we cannot determine whether the deep minimum in the lightcurve is a feature of the nucleus or if it is caused by other effects. MNRAS 000, 1 -- 38 (2017) 5101520Period [hours]0.00.10.20.30.40.50.60.70.8Lomb-Scargle Power21.021.522.0mr [mag]04/7/0505/7/0506/7/0507/7/050.00.20.40.60.81.0Rotational Phase21.021.522.0mr [mag]0.112345ρ [arcsec]23242526272829303132Surface brightness [mag arcsec-2]0.112345ρ [arcsec]222426283032Surface brightness [mag arcsec-2] Rotation of Cometary Nuclei 23 4.5 110P/Hartley 3 Comet 110P/Hartley 3 was observed with VLT-FORS2 and NTT-EFOSC2 during 8 nights between June and August 2011. The aim of the observations was to sample the comet's phase function in the phase angle range between 1◦ and 10◦. The method for precise absolute photometric calibra- tion with PS1 allowed us to combine these datasets and to derive the comet's phase function as well as to study its rotational lightcurve. We looked for signatures of activity on comet compos- ite images for each individual night, and on Fig. 33 we have presented an example for the middle of the observing pe- riod. The comet did not show any indication of coma pres- ence throughout the observing period, and we assume that the derived photometry from each night contains only signal from the nucleus. We used the MC method to derive a phase function for 110P. The determined phase function with linear slope β = 0.069 ± 0.002 mag deg-1 is in excellent agreement with all individual datasets (Fig. 34). All datasets were used to derive the comet's lightcurve under the same assumptions as those described earlier for 14P, 47P and 94P. The LS periodogram in Fig. 35 has three pronounced peaks at Prot,1 = 10.153 hours, Prot,2 = 8.375 hours and Prot,3 = 6.779 hours. The MC method out- lines Prot,1 = 10.153 ± 0.001 hours (75% of the iterations) and Prot,2 = 8.375 ± 0.001 hours (17% of the iterations) as most likely solutions (Fig. 36). Qualitatively, the lightcurve phased with Prot,1 = 10.153 ± 0.001 hours presents less scat- ter of the points and agrees with the trends in the individual observing blocks better. Since Prot,1 is also preferred by the MC method, we report 10.153 ± 0.001 hours as the most likely period of 110P. The brightness variation of the resulting lightcurve is ∆mr = 0.20 ± 0.03 which puts a lower limit on the comet axis ratio a/b ≥ 1.20 ± 0.03. Using Prot,1, we can estimate the nucleus density DN ≥ 0.13 ± 0.02 g cm−3. The mean absolute magnitude of the comet was Hr(1,1,0) = 15.47 ± 0.03 mag, which corresponds to a nucleus radius rN = 2.31 ± 0.03 km, assuming an albedo of 4%. Our results are in good agreement with those of Lamy et al. (2011) (see Section 2.2.5). This validates our results and confirms that it is possible to constrain both the phase function and the lightcurve of the comet from sparse obser- vations spread over months. Although the two observations were taken at different apparitions and a small period change could have occurred during the active phase of the comet, due to the large uncertainty in the period from Lamy et al. (2011), we cannot search for period changes between the two epochs. 4.6 123P/West-Hartley This SEPPCoN target was observed on three consecutive nights in July 2007 while it was at heliocentric distance of 5.6 au. A careful examination of the images indicated that despite the large heliocentric distance at the time of the observations, 123P was weakly active (Fig. 37). The observations from the individual nights clearly in- dicated a brightness variation of the nucleus. However, the LS periodogram of the data did not reveal any pronounced Figure 30. Phase function of comet 94P combining the datasets from 2005, 2007, and 2009. The linear phase function coefficient derived with the Monte Carlo method is β = 0.039 ± 0.002 mag deg-1. Figure 31. Lomb-Scargle periodogram of 94P with the datasets from 2005, 2007, and 2009 combined. The highest peak corre- sponds to the most likely period Prot = 20.70 hours. Figure 32. Rotational lightcurve of 94P with the combined datasets from 2005, 2007 and 2009. The symbols of each data set correspond to those used on Fig. 30. The lightcurve is folded with the best-fitting period Prot = 20.70 hours. The fitted first- order Fourier series (dashed line) agree with all points except for the ones from 18 July 2007. These fainter points could be inter- preted as signatures of an asymmetric lightcurve with one deep minimum, or alternatively as results from the changing viewing geometry between the three epochs. MNRAS 000, 1 -- 38 (2017) 4681012141618α [deg]14.515.015.516.016.517.0Hr(1,1,α) [mag]04/7/0505/7/0506/7/0507/7/0517/7/0718/7/0719/7/0720/7/0722/1/0927/1/0928/1/0929/1/09510152025Period [hours]0.00.10.20.30.40.50.60.70.8Lomb-Scargle Power0.00.20.40.60.81.0Rotational Phase15.015.215.415.615.816.016.216.4Hr [mag] 24 R. Kokotanekova et al. Figure 33. Same as Fig. 2, for 110P on 15 July 2012. The co- added composite image is made up of 18 × 70 s exposures. The comet appears inactive and its surface brightness profile follows that of the comparison star. Figure 34. Phase function of comet 110P. The linear slope β de- rived with the Monte Carlo method is 0.069 ± 0.002 mag deg-1. The NTT-EFOSC2 points from 17 and 18 June 2012 were binned since the S/N of the individual points was low due to bad observ- ing conditions. Figure 35. Lomb-Scargle periodogram of 110P for the combined dataset with all observations from 2012. The three highest peaks correspond to Prot,1 = 10.153 hours, Prot,2 = 8.375 hours and Prot,3 = 6.779 hours. Figure 36. Rotational lightcurve of 110P with all of the data from 2012. The lightcurve is folded with the two most-likely peri- ods 10.153 h (top) and 8.375 hours (bottom) derived from the MC method. The lightcurve with Prot,1 = 10.153 hours is preferred by the MC method (in 75% of the iterations) and it is in better agree- ment with the brightness variation within the individual nights. The symbols are the same as in Fig. 34. The NTT-EFOSC2 points from 17 and 18 June 2012 were binned since the S/N of the indi- vidual points was low due to bad observing conditions. peaks with significant power (Fig. 38). The two highest peaks correspond to 3.7 and 10.3 hours. Those two peri- ods were also preferred by the MC simulation, which picked Prot = 3.70 ± 0.02 hours in 66% of the iterations and Prot = 10.27 ± 0.05 hours (34%). The lightcurves resulting from these two periods are plotted in Fig. 39. Both periods appear to be in agreement with the data, and it is not possible to choose between them. Moreover, the data phased with other periods selected by the periodogram produce lightcurves with similar quality. Therefore, we conclude that the collected data are not suf- ficient to determine the spin rate of 123P. We estimated a brightness variation ∆mr = 0.5 ± 0.1 mag which corresponds to an axis ratio a/b ≥ 1.6 ± 0.1. The mean measured magnitude of 123P was mr = 23.3 ± 0.1 mag which converts to Hr(1,1,0) = 15.7 ± 0.1 mag if a phase function with β = 0.04 mag deg-1 is used. Our absolute mag- nitude and the radius measured by Fern´andez et al. (2013) convert to an albedo Ar = 3.6 ± 0.8% (Eq. 4). It is how- ever important to note that the surface brightness profile of 123P indicated a weak activity, which implies that the abso- lute magnitude Hr(1,1,0) of the nucleus could be fainter and the determined albedo must be treated as an upper limit. 4.7 137P/Shoemaker-Levy 2 Comet 137P was observed during one night in 2005 and two nights in 2007 as part of SEPPCoN. It appeared inactive during both observing epochs (Figs. 40 and 41). We applied the MC method on the combined dataset from all three nights to determine the comet's phase function MNRAS 000, 1 -- 38 (2017) 0.112345ρ [arcsec]222426283032Surface brightness [mag arcsec-2]0246810α [deg]15.415.615.816.016.216.4Hr(1,1,α) [mag]17/6/1218/6/1222/6/1224/6/1212/7/1215/7/1226/7/1219/8/12510152025Period [hours]0.00.10.20.30.40.50.60.7Lomb-Scargle Power15.415.6Hr [mag]0.00.20.40.60.81.0Rotational Phase15.415.6Hr [mag] Rotation of Cometary Nuclei 25 Figure 37. Same as Fig. 2, for 123P on 18 July 2007. The co- added composite image is made up of 23 × 110 s exposures. The comet appears stellar on the composite image, however its surface brightness profile deviates from that of the comparison star, which indicates that the comet was weakly active during the time of the observations. Figure 40. Same as Fig. 2 for 137P on 6 March 2005. The co- added composite image is made up of 23 × 110 s exposures. The comet appears inactive and its surface brightness profile follows that of the comparison star close to the centre before it levels out at the background noise level. Figure 38. Lomb-Scargle periodogram of 123P. The two highest peaks correspond to Prot,1 = 3.7 hours and Prot,2 = 10.7. Figure 41. Same as Fig. 2 for 137P on 13 July 2007. The co- added composite image is made up of 20 × 75 s exposures. The comet appears inactive and its surface brightness profile matches that of the comparison star. (Fig. 42). The derived phase function slope was β = 0.035 ± 0.004 mag deg-1. Next, we attempted to determine the lightcurve period from the data taken in 2005. The highest peak of the peri- odogram in Fig. 43 corresponds to a rotation period of 7.7 hours. However, all peaks on the periodogram have low pow- ers which are not sufficient to determine the rotation rate of 137P. The lightcurve phased with a period of 7.7 hours is plot- ted in Fig. 44. Its brightness variation is ∆mr = 0.18 ± 0.05 mag, which converts to a/b ≥ 1.18 ± 0.05. The uncertainties of the individual points are large in comparison with the de- tected brightness variation. Therefore, it is not possible to derive a precise rotation rate for the comet from this data set. We attempted to improve the period determination by combining all data from 2005 and 2007. However, the pho- tometry from 2007 has even larger photometric uncertainties and does not lead to improvement of the period estimation. The absolute magnitude of 137P is Hr(1,1,0) = 14.63 ± 0.05 mag. Using Eq. 4 and the SEPPCoN radius from Fern´andez et al. (2013), we estimated and albedo Ar = 2.8±0.5%. Figure 39. Rotational lightcurve of 123P with all of the data from 2007. The lightcurve is folded with the most-likely periods 3.7 h (top) and 10.7 hours (bottom). MNRAS 000, 1 -- 38 (2017) 0.112345ρ [arcsec]2426283032Surface brightness [mag arcsec-2]5101520Period [hours]0.000.050.100.150.200.250.300.35Lomb-Scargle Power22.823.023.223.423.6mr [mag]17/7/0718/7/0720/7/070.00.20.40.60.81.0Rotational Phase22.823.023.223.423.6mr [mag]0.112345ρ [arcsec]2426283032Surface brightness [mag arcsec-2]0.112345ρ [arcsec]222426283032Surface brightness [mag arcsec-2] 26 R. Kokotanekova et al. Figure 42. Phase function of comet 137P. The linear phase func- tion coefficient derived from the Monte Carlo simulations is β = 0.035 ± 0.004 mag deg-1. Figure 45. Same as Fig. 2 for 149P on 23 January 2009. The co- added composite image is made up of 15 × 80 s exposures. The comet appears inactive and its surface brightness profile matches that of the comparison star. Figure 43. Lomb-Scargle periodogram of 137P from the 2007 dataset. The highest peak corresponds to a period of Prot = 7.7 hours. Figure 46. Phase function of comet 149P. The linear phase func- tion coefficient derived from the Monte Carlo simulations is β = 0.03 ± 0.02 mag deg-1. degrees between the first and the last observing night. We used the MC method to constrain the phase function slope of the comet as β = 0.03 ± 0.02 mag deg-1. The periodogram of the time series corrected for geo- metric effects peaks at Prot = 11.9 hours. The period of 11.9 ± 0.1 is preferred by the MC simulation in 84% of the itera- tions. However, the power of the peaks on the periodogram is too small and we cannot select the best period unam- biguously. A rotation period near 12 hours would make this measurement for 149P difficult, and a clear determination of such a period using an Earth-based facility would require a longer photometric time sequence. Figure 48 shows the lightcurve of 149P with the best fit from the MC method. The photometric uncertainty of the individual points is large with respect to the total brightness variation of the lightcurve, which confirms that the derived lightcurve is uncertain. The brightness variation of the comet is ∆mr = 0.11 ± 0.04 mag which converts to a/b ≥ 1.11 ± 0.04. The observed mean magnitude of 149P was mr = 22.14 ± 0.04 mag which corresponds to Hr(1,1,0) = 16.93 ± 0.04 if the derived phase function with β = 0.03 ± 0.02 mag deg-1 is used. Using Eq. 4, we can calculate that the albedo of 149P is Ar = 2.8 ± 0.4%. MNRAS 000, 1 -- 38 (2017) Figure 44. Rotational lightcurve of 137P with all of the data from 2007 folded with one of the possible periods, 7.7 hours. The uncertainty of the points is large in comparison to the brightness variation of the comet, which obstructs the period determination. 4.8 149P/Mueller 4 Comet 149P was observed using NTT, WHT and VLT dur- ing 7 nights at the end of January 2009. The surface bright- ness profiles of the comet for each night indicated that it was not active at the time of the observations (see Fig. 45). The phase angle of 149P changed between 8.5 and 10 0123456α [deg]14.414.614.815.015.2Hr(1,1,α) [mag]06/3/0513/5/0714/5/075101520Period [hours]0.000.050.100.150.200.250.300.350.400.45Lomb-Scargle Power0.00.20.40.60.81.0Rotational Phase21.221.421.6mr [mag]13/5/0714/5/070.112345ρ [arcsec]222426283032Surface brightness [mag arcsec-2]789101112α [deg]17.017.117.217.317.4Hr(1,1,α) [mag]21/1/0922/1/0923/1/0924/1/0927/1/0928/1/0929/1/09 Rotation of Cometary Nuclei 27 Figure 47. Lomb-Scargle periodogram of the combined datasets for 149P showing the LS power versus period. The highest peak corresponds to the most likely period Prot = 11.88 hours. Since all peaks have low power, the spin period of the comet cannot be determined unambiguously. Figure 49. Same as Fig. 2, for 162P on 18 May 2007. The co- added composite image is made up of 10 × 110 s exposures. The comet appears inactive and its surface brightness profile agrees with that of the comparison star. Figure 48. Rotational lightcurve of 149P with all of the data from 2009. The points from WHT and NTT were binned The lightcurve is folded with the most-likely period of 11.88 hours. Figure 50. Lomb-Scargle periodogram of the 2007 dataset for 162P showing the LS power versus period. The highest peak cor- responds to the most likely period Prot = 32.6 hours. 4.9 162P/Siding Spring Comet 162P was observed in 2007 around its aphelion, and again in 2012 close to its next aphelion passage. The first set of observations aimed to determine the comet's lightcurve, while the second data set focused on its phase function. The comet had a stellar profile and appeared to be in- active in 2007 (Fig. 49). The LS periodogram of the data from the three observing nights in 2007 is shown in Fig. 50. The most pronounced peak in the periodogram corresponds to Prot = 32.6 hours, and the lightcurve phased with that pe- riod can be seen in Fig. 51. Using the MC method without phase function correction, we determined the rotation pe- riod of the comet to be Prot = 32.6 ± 1 hours. This period is in good agreement with the value of ∼ 33 hours determined by the team of La Canada observatory (see Section 2.2.9). From the observations in 2007, we measured the mean magnitude of 162P to be mr = 20.63 ± 0.05 mag. The bright- ness variation of the comet was ∆mr = 0.45 ± 0.05 mag, which corresponds to a/b ≥ 1.51 ± 0.07. Comet 162P was also inactive during all observations in 2012, which is demonstrated by the surface brightness plot in Fig. 52. Since the observations were taken at a large phase angle range (4-12◦), we could only combine the data MNRAS 000, 1 -- 38 (2017) Figure 51. Rotational lightcurve of 162P with the data from 2007. The lightcurve is folded with period 32.6 hours. after deriving the comet's phase function. The MC method determined a phase function coefficient β = 0.039 ± 0.002 mag deg-1. The LS periodogram of the combined data set from 2012 suggested multiple possible rotation periods for 162P (Fig. 53). The MC method preferred Prot,1 = 33.237 ± 0.008 hours in 62% of the iterations and Prot,2 = 32.852 ± 0.003 hours 510152025Period [hours]0.000.050.100.150.200.250.300.350.40Lomb-Scargle Power0.00.20.40.60.81.0Rotational Phase16.817.0Hr [mag]21/1/0922/1/0923/1/0924/1/0927/1/0928/1/0929/1/090.112345ρ [arcsec]222426283032Surface brightness [mag arcsec-2]5101520Period [hours]0.00.20.40.60.81.0Lomb-Scargle Power0.00.20.40.60.81.0Rotational Phase20.220.420.620.821.021.2mr [mag]17/5/0718/5/0719/5/07 28 R. Kokotanekova et al. Figure 52. Same as Fig. 2, for 162P on 23 April 2012. The co- added composite image is made up of 5 × 60 s exposures. The comet appears inactive and its surface brightness profile generally agrees with that of the comparison star. The narrower profile of the comet is most likely an artefact of the position uncertainty of the comet on the frames. Figure 54. Rotational lightcurve of 162P with the data from 2012. The lightcurve is folded with Prot,1 = 33.237 hours (top) and Prot,2 = 32.852 hours (bottom). It is not possible to choose between the two periods from the data set collected in 2012. Figure 53. Lomb-Scargle periodogram of the 2012 dataset for 162P showing the LS power versus period. There are a number of possible periods as well as secondary peaks caused by aliasing. The highest peaks correspond to rotation periods of 32.852 hours and 33.237 hours. in 35% of the iterations. The lightcurves in Fig. 54 confirm that due to the limited sampling of the lightcurve, it is im- possible to choose between these two possibilities, although it is worth noting that the points from 24 May 2012 agree better with Prot,2 = 32.852. The brightness variation in the 2012 observations was ∆mr = 0.59 ± 0.04 mag, which corresponds to a/b ≥ 1.72 ± 0.06. The absolute magnitude of 162P from the 2012 dataset was Hr(1, 1, 0) = 13.91 ± 0.04 mag. If we use Eq. 4, we can estimate the albedo of 162P to be Ar = 1.8 ± 0.3%. This result makes comet 162P the JFC with the lowest known albedo (see Section 5.4). As a final step in the analysis of the data for 162P, we combined the two datasets from 2007 and 2012 in order to attempt constraining the comet's lightcurve and phase func- tion better. It is possible that the period of 162P slightly changed between 2007 and 2012 while the comet was ac- tive close to perihelion. Besides, it is not excluded that since the two observations were done at different geometries, the resulting lightcurves can appear different. Nevertheless, it is worth attempting to combine the two data sets as the Figure 55. Phase function of comet 162P. The linear phase function slope derived from the Monte Carlo simulations is β = 0.038 ± 0.002 mag deg-1. increased number of observations can provide a better un- derstanding of the nucleus' properties. With these caveats in mind, we proceeded to analyse the combined data from 2007 and 2012. The MC method suggested a phase function with a slope β = 0.038 ± 0.002 mag deg-1 and a lightcurve with period Prot = 32.853± 0.002 hours. This period corresponds to the highest peak of the LS periodogram in Fig. 56. The derived parameters from the combined data set are very close to those of the 2012 data set alone (See. Table 4). However since they were derived using data from two different apparitions, we consider the values from just the 2012 data set to be less uncertain. 5 DISCUSSION In Table 1, we summarised the physical characteristics of all JFCs with known rotation rates. With the newly anal- MNRAS 000, 1 -- 38 (2017) 0.112345ρ [arcsec]20222426283032Surface brightness [mag arcsec-2]5101520Period [hours]0.00.20.40.60.81.0Lomb-Scargle Power13.613.814.014.214.4Hr [mag]Prot = 33.23723/4/1224/5/1214/6/1217/6/1223/6/120.00.20.40.60.81.0Rotational Phase13.613.814.014.214.4Hr [mag]Prot = 32.852468101214α [deg]13.613.814.014.214.414.614.8Hr(1,1,α) [mag]17/5/0718/5/0719/5/0723/4/1224/5/1214/6/1217/6/1223/6/12 Table 4. Derived physical parameters for all observed comets. Rotation of Cometary Nuclei 29 Comet Epoch β [mag/deg]3 rN [km]4 14P 47P 93P 94P 110P 123P 137P 149P 162P 2004 2007 Combined 2005* 2006* 2015* 2005 + 2015** - 2009* 2005 2007 2009 Combined 2012 2007* 2007 2005 + 2007 2009 2007 2012 Combined** Hr(1,1,0)1 14.59±0.06 14.87±0.05 mr1 Prot [h]2 22.58±0.05 - 8.93±0.04 21.06±0.05 - 9.02±0.04 9.02±0.01 - 21.83±0.06 - 10.8/14.1 21.55±0.04 - - 21.11±0.06 14.58±0.06a 15.6±0.1 21.09±0.05 15.17±0.05b 18.2+1.5−15 21.3±0.1 - 22.6±0.2 - 21.30±0.05 - 15.50±0.09 - 15.47±0.03 - 15.7±0.1b 23.3±0.1 21.39±0.05 - 14.63±0.05 - 22.14±0.04 16.93±0.04 20.63±0.05 - 13.91±0.04 - 13.90±0.05 - - - 0.060±0.005 - - - 0.096±0.004 - - 20.43±0.05 - - - - - 20.70±0.07 0.039±0.002 10.153±0.001 0.069±0.002 - - - - 0.035±0.004 - 0.03±0.02 - 32.6±1 - 33.237/32.852 0.039±0.002 0.038±0.002 32.853±0.002 a/b DN [g cm−3] Ar [%]5 - ∆mr 0.36±0.05 1.39±0.06 0.19±0.04 - - 0.39±0.05 1.43±0.07 0.19±0.04 - - 0.37±0.05 1.41±0.06 0.19±0.03 4.3±0.6 - 0.33±0.06 1.36±0.07 - - - - - - - 0.24±0.06 1.25±0.07 0.06±0.02 - - 5.0±0.7c - - - - 4.2±0.9c 0.21±0.05 1.21±0.06 - - 1.9±0.2 0.7±0.1 0.05±0.01 - - 1±0.2 2.5±0.5 - - - 0.80±0.05 2.09±0.10 - - - 1.11±0.09 2.8±0.2 4.0±0.6 0.07±0.02 - 0.20±0.03 1.20±0.03 0.13±0.02 2.31±0.03 - 3.6±0.8c 0.5±0.1 1.6±0.1 - - 0.18±0.05 1.18±0.05 - - - 2.8±0.5 - - - 0.11±0.04 1.11±0.04 - 2.8±0.4 - 0.45±0.05 1.51±0.07 - - - 1.8±0.3 0.59±0.04 1.72±0.06 0.017±0.003 - 0.62±0.05 1.77±0.08 0.018±0.003 1.8±0.3 - - 1 Magnitudes in PS1 system. 2 The synodic rotation periods and their uncertainties were derived from the mean and standard deviation from the MC method (see Section 3.4.8). 3 The linear phase function coefficients and their uncertainties were derived from the mean and standard deviation from the MC method (see Section 3.4.8). 4 Calculated from Hr(1,1,0) assuming an albedo A=4%. 5 Calculated using Eq. 4 from Hr(1,1,0) and the effective radius Reff from Fern´andez et al. (2013) (see Tab. 1). * The comet was weakly active. The results do not include corrections for the presence of a near-nucleus coma. ** The data are from different apparitions. a The β value for the Hr(1,1,0) was taken from the phase function fit of the combined 2005 and 2015 datasets. b Calculated for β = 0.04 mag deg-1. c The comet was weakly active at the time of the observation. The albedo estimates are therefore upper limits. Figure 56. Lomb-Scargle periodogram of the combined datasets of 162P from 2007 and 2012 showing the LS power versus period. The highest peak corresponds to Prot = 32.853 hours. Figure 57. Rotational lightcurve of 162P with the data from 2007 and 2012. The lightcurve is folded with the most likely period of 32.853 hours. ysed lightcurves in Section 4, we have added six additional lightcurves, seven phase functions and eight albedo esti- mates. Here, we compare our newly obtained results with the overall JFC characteristics and use the expanded sample to draw conclusions about the collective population properties. MNRAS 000, 1 -- 38 (2017) 5.1 Spin rate distribution The distribution of the rotation rates of comets can be used to study their collisional history. Fig. 58 shows a histogram of all known spin rates of JFCs. We have plotted the rota- tion frequency f = 1/Prot which was normalised using the geometric mean (cid:104) f (cid:105) of the whole sample. Similar plots for asteroids have shown that the distribution of asteroid spin 5101520Period [hours]0.00.20.40.60.81.0Lomb-Scargle Power0.00.20.40.60.81.0Rotational Phase13.413.613.814.014.214.4Hr [mag]17/5/0718/5/0719/5/0723/4/1224/5/1214/6/1217/6/1223/6/12 30 R. Kokotanekova et al. rates is Maxwellian which has suggested that asteroids are a collisionally evolved population (Harris 1996; Pravec et al. 2002). The best-fitting Maxwell distribution in Fig. 58 does not show good agreement with the measured spin rates. We performed Kolmogorov-Smirnov tests comparing the nor- malised frequency distribution in Fig. 58 to Maxwell distri- bution and flat distribution with the same mean and stan- dard deviation. The resulting D statistics were 0.20 (p = 0.09) and 0.13 (p = 0.44) for the uniform and Maxwell dis- tributions respectively. We cannot reject the null hypothesis in either of the cases, and therefore both distributions can possibly describe the data. The cumulative size distribution (CSD) of JFCs was found to be very close to the one expected for a collisionally relaxed population of strengthless bodies (Lamy et al. 2004; Snodgrass et al. 2011; Fern´andez et al. 2013, and references therein). However, this result has a large uncertainty and cannot be used as a proof that JFCs originate from disrupted larger bodies (e.g KBOs). In turn, it suggests that due to the continuous mass loss of JFCs their size distribution can be shaped by a complex combination of collisional processes in the past and activity in the present epoch (Snodgrass et al. 2011). Similarly, our results for the spin distribution of comets suggest that their rotation can be determined by the ongoing activity. The mass lost through activity jets is able to exert a torque on the nucleus, which in turn changes the spin rate of the comet on orbital timescales (e.g. Samarasinha et al. 2004). This mechanism can be responsible for reshaping the original distribution of the spin rates, and could explain the current spin rate distribution of JFC. However, it is impor- tant to know that Fig. 58 includes data from just 37 comets, many of which have lightcurve periods with large uncertain- ties. This highlights the need to increase the sample of JFCs with known rotational properties in order to enable the un- derstanding of the population history. It is worth noting that evidence from Rosetta, such as the low density/high porosity, and presence of hypervolatiles like O2 and N2, suggests that 67P is not a collisional frag- ment (see Davidsson et al. 2016, and references therein). The apparent coincidence of sizes and spin rates of JFC nu- clei being consistent with collisional evolution, while in situ measurements of their bulk properties suggest otherwise, is surprising. This may instead support the hypothesis by Jutzi et al. (2017) that JFCs have undergone significant collisional evolution, but the distributions presented here do not yet al- low a definitive conclusion. 5.2 Shapes Fig. 59 shows the distribution of the axis ratios of all comets. Most a/b values are smaller than a/b = 2 and the median of the distribution is at a/b = 1.5. However, all comets with shape models obtained from in situ observations (9P, 19P, 67P, 81P, 103P) have significantly higher axis ratios (see Ta- ble 1). For all other objects the axis ratio is a lower limit since it was calculated from the lightcurve brightness vari- ation. It is therefore possible that the typical elongation of JFCs is higher than the one we estimated from the current distribution, suggesting that bilobate shapes (like those seen Figure 58. Histogram of the normalised rotation rates of 37 JFCs. The normalised spin rate is calculated as f /(cid:104) f (cid:105) where f = 1 / Prot and (cid:104) f (cid:105) is the geometric mean of f . The dashed line corresponds to the best-fitting Maxwellian distribution. Figure 59. Distribution of the axis ratios a/b of JFCs. The ver- tical line corresponds to the median value of a/b = 1.5. For all comets (except 9P, 19P, 67P, 81P, 103P), the given axis ratio is obtained from ground- and space-based telescope and is therefore just a lower limit of the elongation. by spacecraft at 67P and 103P) may be common, in agree- ment with recent formation models (Davidsson et al. 2016). 5.3 Bulk densities and stability against rotational splitting We attempted to use the expanded sample of JFCs with estimated rotation rates and elongations to constrain the comet density and tensile strength. As we discussed in Sec- tion 3.4.7, it is commonly assumed that comets have negli- gible tensile strengths. Under this assumption, it is possible to set a lower limit on the density necessary to keep JFCs stable against rotational instabilities (Eq. 6; Pravec et al. 2002). In Fig. 60 we plot the rotation versus projected axis ratio for all comets in the expanded sample. Using a similar plot, Lowry & Weissman (2003) discovered that comets do not require densities higher than approximately 0.6 g cm-3 in order to be stable against rotational instabilities. Here we confirm this result for all objects except for 322P, 73P-C and 147P. As we discussed in section 2.1.17, according to Knight MNRAS 000, 1 -- 38 (2017) 0.00.51.01.52.02.53.0Normalised spin rate0123456Number1.01.52.02.53.03.5a/b 02468Number et al. (2016) it is not clear whether 322P has asteroidal or cometary origin. Therefore, the fact that it requires higher density can be interpreted as evidence in favour of the hy- pothesis that it is an asteroid. Comet 147P lies very close to the limit of 0.6 g cm-3 and has a large period uncertainty. Therefore, we do not consider it as an outlier. Additionally, 147P belongs to the class of quasi-Hilda comets and might have asteroidal origin (Ohtsuka et al. 2008). Comet 73P-C on the other hand clearly has a JFC origin and therefore should be similar to the other objects in our sample. How- ever, since it seems to be continuously disintegrating (see section 2.1.7), it cannot be used to study the stability crite- rion. It is also possible that the breakup of the comet exposed the innermost part of the pre-breakup nucleus which could have a larger tensile strength (see Gundlach et al. 2016, and references therein) If we exclude these three comets, our expanded sam- ple confirms the density limit of 0.6 g cm-3 discovered by Lowry & Weissman (2003). By analogy with the clear cut off in rotation rates of asteroids at 2.2 g cm-3 (Pravec et al. 2002), we interpret the cut-off for comets as an indication that 0.6 g cm-3 is a typical density for JFCs. This agrees with the density estimates from recent spacecraft measure- ments (Richardson et al. 2007; Jorda et al. 2016). Further insights into the material properties of JFCs can be determined from comparing their rotation rates and sizes. In previous studies, Davidsson (1999, 2001) and Toth & Lisse (2006) already explored the location of comets and other primitive minor bodies in the radius-rotation period plane. In Fig. 61 we plot the distribution of rotation rates with radius for all comets. A key feature of the distribution of comets in the plot is that the domain in the lower right corner is not populated. In order to interpret this observation, we employ the recent results from the Rosetta mission. The in situ mea- surements of comet 67P provide precise estimates of the nu- cleus bulk parameters. It has density of 0.532 ± 0.007 g cm-3 (Jorda et al. 2016), axis ratio a/b = 2.05 ± 0.06 (calculated from the axis estimates in Jorda et al. 2016), and tensile strength of 3-15 Pa with an upper limit of 150 Pa (Groussin et al. 2015). If we assume that 67P is a representative exam- ple for JFCs, we can use these values to study the properties of the whole population. In Fig. 61, we have plotted the asteroid spin barrier (Pravec et al. 2002) which corresponds to the minimum ro- tation period of a strengthless body with density ∼ 3 g cm-3. For a comparison, we have also plotted the rotation limit for a spherical object with density of 0.6 g cm-3. The position of the limit for comets will change for different elongations and densities since less dense and more elongated objects are easier to disrupt. So far in the analysis, we have treated comets as strengthless, however the measurements of the tensile strength of 67P allow us to explore more complicated mod- els which take the material strength of comets into account. We have used the analytical models developed by Davids- son (1999, 2001) to determine the maximum rotation rate of prolate ellipsoids which are stable against rotational in- stabilities using the density, axis ratio and tensile strength of 67P (Fig. 61, solid green curve). This curve agrees very well with the observed data and puts 73P-C right at the limit of stability, which agrees with its frequent fragmenta- MNRAS 000, 1 -- 38 (2017) Rotation of Cometary Nuclei 31 tion events. Although comet 31P lies below the stability line, its projected axis ratio is lower than that of 67P. We have therefore investigated the stability limit for objects with density of 0.5 g cm-3 and a typical axis ratio of a/b = 1.6 (equal to the elongation of 31P). We deter- mine that under these assumptions none of the comets re- quires tensile strength higher than ∼ 10 Pa to remain stable against rotational instabilities (Fig. 61, dashed blue curve). We varied the axis-ratio parameter of the model for ratios a/b ≤ 2.0 and concluded that none of the observed comets requires a tensile strength larger than 25 Pa to remain sta- ble against rotational splitting. This confirms the low-tensile strength estimates of 67P by Groussin et al. (2015) and of Shoemaker-Levy 9 (Asphaug & Benz 1996). An interesting test of this model would come from fu- ture observations of the rotation rate of 31P. The comet's period was previously very well determined by Luu & Je- witt (1992). If new observations of its lightcurve show that the nucleus is spinning up, this comet would be a strong candidate for future rotational splitting. Despite the small number of nuclei with radii larger than 3 km in the sample, it is noticeable that all of them lie far above the stability limit. The simplest explanation for this effect could be deduced from the understanding of activity-induced rotational changes. According to the rela- tions derived in Samarasinha & Mueller (2013), the rota- tion changes induced by outgassing are proportional to the square of the rotation period and inversely proportional to the square of the radius. In this scenario, if a large nucleus is spinning up due to reaction torques, the faster it gets, the less it can spin up with every orbit. Therefore, it is very hard to spin up the large nuclei which already rotate with relatively short periods. At this stage, we cannot evaluate this hypothesis further since spin changes are poorly investigated and to this date only 5 comets have confirmed period changes (see Samaras- inha & Mueller 2013, and references therein). To improve the understanding of the rotation of large comets, we need to measure the rotation rates of more large nuclei and to increase the number of comets with period determinations at multiple apparitions. Finally, in Fig. 61, we have also plotted all active aster- oids with known periods and radii (Jewitt et al. 2015). Most of them lie in the lower right domain of the plot where no JFCs can be found. However, it is particularly interesting to note that 107P fulfils the stability criteria for comets too. This object has sparked a long-standing debate on whether it is a comet or an active asteroid (see Jewitt et al. 2015, and references therein) Since 107P is above the stability limit for typical JFCs, we cannot reject the possibility that it has a cometary origin. 5.4 Surface Properties Prior to this work, there were only nine comets for which both the albedo and the phase function were known (Snod- grass et al. 2011). We have significantly increased this num- ber by updating the values for one comet and adding the measurements for five additional comets from this work. We have summarized the albedos and phase function coefficients for 24 comets in Table 5. The median of all known linear phase function slopes is 0.046 mag/deg and the standard 32 R. Kokotanekova et al. Figure 60. Rotation period against projected axis ratio for JFC nuclei. The grey triangles denote comets with parameters determined from lightcurve or radar measurements. The orange circles are the comets from this work. For these points, the axis ratio is a lower limit and the uncertainties are plotted when they were stated by the authors. The blue diamonds correspond to comets visited by spacecraft with precise shape models. The diagonal lines indicate the minimum density (denoted in g cm-3 to the right), which a strengthless body of the given axis ratio and spin period requires to remain intact. Apart from the unusual cases of 323P and 73P, which are discussed in the text, no comet requires a density greater than ∼ 0.6 g cm-3 to remain stable against rotational splitting. deviation is 0.017 mag/deg. The median of all albedos is 4.2% and the standard deviation is 1.3%. We have looked for possible correlations between the surface properties of the comets and their sizes. In Fig. 62 it can be seen that large JFCs tend to have low albedos and small phase function coefficients. The albedo distribu- tion with size agrees with the one presented by Fernandez et al. (2016), which consisted of a larger sample of approxi- mately 50 comets with albedos derived within the SEPPCoN program. We note a possible correlation between the phase func- tion coefficient and the albedo in Fig. 62 (top panel). It is well established that similar correlations exist between albedo or spectral type and phase functions for asteroids (e.g. Oszkiewicz et al. 2012). Comet 47P was determined to be active at the time of the observations which were used to determine its albedo and phase function. Under these conditions, it is possible that we have overestimated the nucleus brightness and therefore un- derestimated its albedo. Additionally, the activity possibly led us to determine an inaccurate phase function. Due to these concerns, we prefer to exclude it from the analysis. We performed a Spearman rank correlation test be- tween the phase function coefficient and the albedo of all comets (excluding 47P). The test produced rank ρ of 0.82 and p-value of 0.0005 which suggests a possible correlation between the phase function coefficients and albedos. In or- der to confirm this possible correlation and to be able to interpret it, we need to increase the number of JFCs with well-determined surface properties. 6 SUMMARY We have developed a method for precise absolute calibration of photometric time series using Pan-STARRS DR1 stars. With this technique we achieved photometric calibration with uncertainty as low as 0.02 mag. Thus we were able to MNRAS 000, 1 -- 38 (2017) 1.01.52.02.53.03.5Projected axial ratio510203040Period [hours]2.00.60.20.060.022P6P7P9P10P17P19P21P22P28P31P36P46P48P49P61P67P73P-C76P81P82P87P92P103P121P143P147P169P209P260P322P14P47P93P94P110P162P Rotation of Cometary Nuclei 33 Figure 61. Rotation period against effective radius of the JFC nuclei. The blue diamonds are comets visited by spacecraft; the grey squares are comets observed from ground and the orange circles are the comets added in this work. For comparison we plotted active asteroids with known rotation rates (pink pentagons). The lower horizontal dotted line corresponds to the asteroid spin barrier (Harris 1996; Pravec et al. 2002). The upper dashed pink line shows the minimum possible rotation rate for strengthless spherical bodies with density ρ = 0.6 g cm-3. The curves are derived from the model for prolate ellipsoids stable against rotational instability by Davidsson (2001). The solid green line is the model for density ρ = 532 kg m-3, axis ratio a/b = 2 and tensile strength T = 15 Pa, which corresponds to the parameters measured for 67P from Rosetta (Jorda et al. 2016; Groussin et al. 2015). The dashed blue curve is for the same density but a/b = 1.6 (the value for 31P) and T = 10 Pa. By varying the model parameters, we can conclude that for typical densities and axis ratios (a/b ≤ 2.0), none of the observed comets requires a tensile strength larger than 25 Pa to remain stable against rotational splitting. study the rotation, shapes, and surfaces of nine Jupiter fam- ily comets, most of which were observed at multiple epochs using different instruments. We have collected an up-to-date sample of JFCs with published rotational properties and ex- panded it with the measurements from this work. We used the extended sample to characterise the bulk properties of JFCs. The results are as follows: (i) We have used time-series photometry of nine JFCs taken in the period 2004-2015 to study their lightcurves. We have derived the rotation rates of six comets (14P, 47P, 93P, 94P, 110P, and 162P). For comets 123P, 137P and 149P the collected data were insufficient to derive unambiguous rotation periods. To our knowledge, for comets 93P, 94P and 162P these are the first published rotation rates. Comets 14P, 47P and 110P had previous lightcurves but our results significantly improved the period estimates. (ii) Lower limits on the axis ratios of all observed comets have been derived from the brightness variation of the time series. Three of the comets, 47P, 93P and 123P, were most likely active at the times of the observations and therefore their brightness variation was most likely underestimated. (iii) We have determined the linear phase function coeffi- cients of seven of the observed comets - 14P, 47P, 94P, 110P, 137P, 149P, and 162P. To our knowledge, for all comets except 47P, this is the first phase function determination. Our results have increased the number of comets with well- constrained phase function coefficients from 13 to 19. (iv) The derived phase function coefficients have been used in the calculation of the absolute magnitudes Hr(1,1,0). For comets 93P and 123P, we used a phase function coeffi- cient β=0.04 mag/deg. All comets except for 110P were part of SEPPCoN and had radius measurements derived from MNRAS 000, 1 -- 38 (2017) 110Effective Radius [km]2510203040Period [hours]Asteroid Spin Barrier2P6P7P9P10P17P19P21P22P28P31P36P46P48P49P61P67P73P-C76P81P82P87P92P103P121P143P147P169P209P260P322P14P47P93P94P110P162P107P133P(3200) PhaethonP/2012 F52000 SY178176P 34 R. Kokotanekova et al. Table 5. Albedo and phase function measurements for JFCs. Comet A [%] Reference 2P " " 9P " " " 10P 14P 19P " " 22P 28P 36P 45P 47P " 48P 49P 67P " 81P 93P 94P 103P 110P 123P 137P 143P 149P 162P " 169P 5.0 ± 2.0 - - 6.1 ± 0.8 6.4 ± 1.3 4.6 ± 1.5 7.2 ± 1.6 3.0 ± 1.2 4.3 ± 0.6 3.3 ± 0.6 2.9 ± 0.6 7.2 ± 2.0 4.8 ± 1.0 3.0 ± 1.0 - - ≤ 5.0 ± 0.7 This work - - 4.5 ± 1.9 6.5 ± 0.2 5.4 ± 0.6 6.4 ± 1.0 4.2 ± 0.9 4.0 ± 0.6 4.5 ± 0.9 - 3.6 ± 0.8 2.8 ± 0.5 - 2.8 ± 0.4 1.8 ± 0.3 3.7 ± 1.4 3.0 ± 1.0 Fern´andez (2000) - - Weighted mean Li et al. (2007a) Lisse et al. (2005) Fern´andez et al. (2003) A'Hearn et al. (1989) This work Weighted mean Buratti et al. (2004) Li et al. (2007b) Lamy et al. (2002) Jewitt & Meech (1988) - - - - Campins et al. (1995) Fornasier et al. (2015) Kelley et al. (2009) Li et al. (2009) This work This work Li et al. (2013) - This work This work - This work This work Fernandez et al. (2006) DeMeo & Binzel (2008) β [mag/deg] 0.053 ± 0.003 0.060 ± 0.005 0.060 ± 0.005 0.046 ± 0.007 - - - 0.037 ± 0.004 0.060 ± 0.005 0.043 ± 0.009 - - - 0.025 ± 0.006 0.060 ± 0.019 ∼0.06 0.096 ± 0.004 0.083 ± 0.006 0.059 ± 0.002 - 0.074 ± 0.006 0.076 ± 0.003 0.0513 ± 0.0002 - 0.039 ± 0.002 0.046 ± 0.002 0.069 ± 0.002 - 0.035 ± 0.004 0.043 ± 0.001 0.03 ± 0.02 0.039 ± 0.002 - - α Range [deg] Reference. - 0-110 4-28 4-117 - - - 9-28 5-9 13-80 - - - 0-15 1-11 88-93 3-9 2-9 5-16 - 1-10 0-11 0-100 - 5-17 79-95 1-9 - 0.5-6 5-13 8-10 4-12 - - Weighted mean Fern´andez (2000) Boehnhardt et al. (2008) Li et al. (2007a) - - - Sekanina & Zdenek (1991) This work Li et al. (2007b) - - - Delahodde et al. (2001) Snodgrass et al. (2008b) Lamy et al. (2004) This work Snodgrass et al. (2008b) Jewitt & Sheppard (2004) - Fornasier et al. (2015) Tubiana et al. (2008) Li et al. (2009) - This work Li et al. (2013) This work - This work Jewitt et al. (2003) This work This work - - Spitzer infrared observations (Fern´andez et al. 2013). Using these radii and the derived absolute magnitudes, we have estimated the albedos of all eight comets. (v) We derived a geometric albedo of 1.8 ± 0.3 % for comet 162P. This makes 162P the JFC with lowest mea- sured geometric albedo to date. (vi) Prior to this work, there were nine comets for which both the albedo and the phase function coefficient were known (see Snodgrass et al. 2011). We have updated the values for 47P and added five more comets (14P, 94P, 137P, 149P, 162P) to this sample. The increased number of comets has allowed us to look for correlations between the surface properties of JFCs. Large nuclei (Reff ≥ 5 km) appear to have low albedos (≤ 3 %) and low phase function coefficients (≤ 0.04 mag/deg). How- ever, since only three comets in that size range have been observed, this needs to be confirmed with future observa- tions. We have discovered a possible correlation between the phase function coefficient and the albedo, where comets with larger albedos have steeper phase functions. In order to con- firm as well as to interpret this result, we would require further phase function observations. (vii) In Table 5 we have collected the known albedos and phase functions of 24 JFCs. The distribution of the linear phase function slopes has a median of 0.046 mag/deg and standard deviation of 0.017 mag/deg. The known albedos have a median of 4.2% and standard deviation of 1.3% (viii) We have reviewed the properties of all JFCs which (to the extent of our knowledge) had published rotation rates. After adding the six comets from this work, the total size of the sample has reached 37 comets. (ix) We have attempted to use the distribution of spin rates to improve the understanding of JFC evolution. The employed Kolmogorov-Smirnov tests determined that the normalised spin rates of comets is consistent both with a Maxwell distribution and a flat distribution. Therefore, we cannot distinguish between the possibilities that JFCs are a collisionally-dominated population like asteroids or their spin rate distribution is dominated by other processes, such as activity-driven spin changes. (x) The distribution of the axis ratios shows that the ma- jority of comets have projected axis ratios smaller than 2. The median of the whole JFC sample is 1.5. However, ground observations only give a lower limit to the axis ratio. All five comets with shape models determined from in situ space craft observations have axis ratios larger than 1.6. (xi) Under the assumption that JFCs have negligible ten- sile strengths, we have used their axis ratios and periods to constrain their bulk densities. We have confirmed the result from Lowry & Weissman (2003) that a density of 0.6 g cm−3 MNRAS 000, 1 -- 38 (2017) Rotation of Cometary Nuclei 35 instabilities. An interesting outcome of this analysis is that comet 31P also lies very close to the stability limit which makes it a candidate for potential future splitting. ACKNOWLEDGEMENTS 082.C-0517(B), We would like to thank the referee, Imre Toth, for the helpful suggestions that improved the manuscript. This work is based on observations collected at the European Organisation for Astronomical Research in the Southern Hemisphere under ESO programmes 072.C-0233(A), 074.C- 0125(A), 077.C-0609(B), 079.C-0297(A), 079.C-0297(B), 082.C-0517(A), 089.C- 0372(B), and 194.C-0207(C). We thank the observatory staff who helped us obtain the various data sets in this paper, especially those collected in service mode at the VLT. Also based on observations made with the WHT and INT operated on the island of La Palma by the Isaac Newton Group of Telescopes in the Spanish Observatorio del Roque de los Muchachos of the Instituto de Astrofisica de Canarias, under UK PATT programmes I/2005A/11, W/2007A/20 and W/2008B/23. 089.C-0372(A), CS is funded by a STFC Ernest Rutherford fellowship. AF acknowledges support from STFC grant ST/L000709/1. The Pan-STARRS1 Surveys (PS1) and the PS1 pub- lic science archive have been made possible through contri- butions by the Institute for Astronomy, the University of Hawaii, the Pan-STARRS Project Office, the Max-Planck Society and its participating institutes, the Max Planck In- stitute for Astronomy, Heidelberg and the Max Planck In- stitute for Extraterrestrial Physics, Garching, The Johns Hopkins University, Durham University, the University of Edinburgh, the Queen's University Belfast, the Harvard- Smithsonian Center for Astrophysics, the Las Cumbres Ob- servatory Global Telescope Network Incorporated, the Na- tional Central University of Taiwan, the Space Telescope Science Institute, the National Aeronautics and Space Ad- ministration under Grant No. NNX08AR22G issued through the Planetary Science Division of the NASA Science Mis- sion Directorate, the National Science Foundation Grant No. AST-1238877, the University of Maryland, Eotvos Lorand University (ELTE), the Los Alamos National Laboratory, and the Gordon and Betty Moore Foundation. REFERENCES A'Hearn M. F., Campins H., Schleicher D. G., Millis R. L., 1989, The Astrophysical Journal, 347, 1155 A'Hearn M. F., et al., 2005, Science (New York, N.Y.), 310, 258 A'Hearn M. F., et al., 2011, Science, 332, 1396 Appenzeller I., et al., 1998, The Messenger, vol. 94, p. 1-6, 94, 1 Asphaug E., Benz W., 1996, Icarus, 121, 225 Belskaya I. N., Levasseur-Regourd A.-C., Shkuratov Y. G., Muinonen K., 2008, The Solar System Beyond Neptune, M. A. Barucci, H. Boehnhardt, D. P. Cruikshank, and A. Mor- bidelli (eds.), University of Arizona Press, Tucson, 592 pp., p.115-127, pp 115 -- 127 Belton M., Samarasinha N., Fernandez Y., Meech K., 2005, Icarus, 175, 181 Belton M. J., et al., 2011, Icarus, 213, 345 Belton M. J., et al., 2013, Icarus, 222, 595 Figure 62. Surface properties of all JFCs with known radius, albedo and phase function. The orange circles correspond to comets with properties derived in this work. Top: Phase function slope versus albedo. There is a trend of increasing phase function slope with increasing albedo. Comet 47P was active at the time of the observations, so in reality its phase function coefficient might be smaller and its albedo might be higher. Middle: albedo versus radius. Bottom: phase function coefficient versus radius. is sufficient to keep all of the studied nuclei stable against rotational instabilities. (xii) If we instead model JFCs as prolate ellipsoids with non-negligible tensile strengths using the model from Davidsson (2001), we conclude that none of the observed comets requires tensile strength higher than 10-25 Pa in or- der to be stable against rotational splitting. Comet 73P-C lies very close to the stability limit we derived, which sug- gests that its ongoing splitting might be due to rotational MNRAS 000, 1 -- 38 (2017) 234567Albedo [%]0.020.040.060.080.10β [mag/deg]2P9P10P19P28P67P81P103P14P47P94P137P149P162P246810Radius [km]234567Albedo [%]2P9P10P19P28P67P81P103P14P47P94P137P149P162P246810Radius [km]0.020.040.060.080.10β [mag/deg]2P9P10P19P28P67P81P103P14P47P94P137P149P162P 36 R. Kokotanekova et al. Boehnhardt H., et al., 2002, Astronomy and Astrophysics, 387, Groussin O., et al., 2015, Astronomy & Astrophysics, Volume 583, 1107 Boehnhardt H., Tozzi G. P., Bagnulo S., Muinonen K., Nathues A., Kolokolova L., 2008, Astronomy and Astrophysics, Volume 489, Issue 3, 2008, pp.1337-1343, 489, 1337 Bohnhardt H., Kaufl H. U., Keen R., Camilleri P., Carvajal J., Hale A., 1995, IAU Circ., No. 6274, #1 (1995). Edited by Green, D. W. E., 6274 id.A32, 12 pp., 583 Gundlach B., Blum J., Blum J., 2016, Astronomy & Astrophysics, 589, A111 Gutierrez P. J., de Leon J., Jorda L., Licandro J., Lara L. M., Lamy P., 2003, Astronomy and Astrophysics, 407, L37 Harmon J., Nolan M., 2005, Icarus, 176, 175 Harmon J. K., Nolan M. C., Howell E. S., Giorgini J. D., Taylor Buratti B., Hicks M., Soderblom L., Britt D., Oberst J., Hillier P. A., 2011, The Astrophysical Journal, 734, L2 J., 2004, Icarus, 167, 16 Harris A. W., 1996, Lunar and Planetary Science, volume 27, Buzzoni B., et al., 1984, ESO Messenger (ISSN 0722-6691), Dec. page 493, 27 1984, p. 9-13., 38, 9 Campins H., 1988, Icarus, 73, 508 Campins H., Osip D. J., Rieke G., Rieke M., 1995, Planetary and Space Science, 43, 733 Campins H., Ziffer J., Licandro J., Pinilla-Alonso N., Fern´andez Y., de Le´on J., Moth´e-Diniz T., Binzel R. P., 2006, The As- tronomical Journal, 132, 1346 Chambers K. C., et al., 2016, http://arxiv.org/abs/1612.05560 Chesley S., et al., 2013, Icarus, 222, 516 Ciarniello M., et al., 2015, Astronomy & Astrophysics, 583, A31 Crovisier J., Biver N., Bockelee-Morvan D., Colom P., Gerard E., Jorda L., Rauer H., 1995, IAU Circ., No. 6227, #1 (1995). Edited by Green, D. W. E., 6227 Davidsson B. J. R., 1999, Icarus, 142, 525 Davidsson B. J. R., 2001, Icarus, 149, 375 Davidsson B. J. R., Sierks H., Guttler C., Marzari F., Pajola M., Rickman H., 2016, Astronomy & Astrophysics Davis L. E., 1999, in Craine E. R., Tucker R. A., Barnes J. V., eds, Vol. 189, Precision CCD Photometry. ASP Conference Series, p. 35 DeMeo F., Binzel R. P., 2008, Icarus, 194, 436 Delahodde C. E., Meech K. J., Hainaut O. R., Dotto E., 2001, Astronomy and Astrophysics, 376, 672 Drahus M., Kuppers M., Jarchow C., Paganini L., Hartogh P., Villanueva G. L., 2010, Astronomy and Astrophysics, 510, A55 Drahus M., et al., 2011, The Astrophysical Journal Letters, Vol- ume 734, Issue 1, article id. L4, 6 pp. (2011)., 734 Duncan M., Levison H., Dones L., 2004, Comets II, M. C. Festou, H. U. Keller, and H. A. Weaver (eds.), University of Arizona Press, Tucson, 745 pp., p.193-204, pp 193 -- 204 Duxbury T. C., Newburn R. L., Brownlee D. E., 2004, Journal of Geophysical Research, 109, E12S02 Dworetsky M. M., 1983, Monthly Notices of the Royal Astronom- ical Society, 203, 917 Farnham T. L., 2001, American Astronomical Society, DPS Meet- ing #33, id.12.10; Bulletin of the American Astronomical So- ciety, Vol. 33, p.1047, 33, 1047 Fern´andez Y., 2000, Icarus, 147, 145 Fern´andez Y., Meech K., Lisse C., A'Hearn M., Pittichov´a J., Belton M., 2003, Icarus, 164, 481 Fern´andez Y., Lowry S., Weissman P., Mueller B., Samarasinha N., Belton M., Meech K., 2005, Icarus, 175, 194 Fernandez Y. R., Campins H., Kassis M., Hergenrother C. W., Binzel R. P., Licandro J., Hora J. L., Adams J. D., 2006, The Astronomical Journal, Volume 132, Issue 3, pp. 1354-1360., 132, 1354 Fern´andez Y., et al., 2013, Icarus, 226, 1138 Fernandez Y. R., Weaver H. A., Lisse C. M., Meech K. J., Lowry S. C., Bauer J. M., Fitzsimmons A., Snodgrass C., 2016, American Astronomical Society, AAS Meeting #227, id.141.22, 227 Fornasier S., et al., 2015, Astronomy & Astrophysics, 583, A30 Glassmeier K.-H., Boehnhardt H., Koschny D., Kuhrt E., Richter Hergenrother C., 2014, IAU CBET, 3870 Hoenig S. F., 2005, Astronomy and Astrophysics, Volume 445, Issue 2, January II 2006, pp.759-763, 445, 759 Howell S. B., 1989, Publications of the Astronomical Society of the Pacific, 101, 616 Howell E. S., et al., 2014, American Astronomical Society, DPS meeting #46, id.209.24, 46 Jehin E., Manfroid J., Hutsemekers D., Gillon M., Magain P., 2010, Central Bureau Electronic Telegrams, No. 2589, #1 (2010). Edited by Green, D. W. E., 2589 Jester S., et al., 2005, The Astronomical Journal, Volume 130, Issue 3, pp. 873-895., 130, 873 Jewitt D., 2009, The Astronomical Journal, Volume 137, Issue 5, pp. 4296-4312 (2009)., 137, 4296 Jewitt D., Luu J., 1989, The Astronomical Journal, 97, 1766 Jewitt D., Meech K., 1987, The Astronomical Journal, 93, 1542 Jewitt D. C., Meech K. J., 1988, The Astrophysical Journal, 328, 974 Jewitt D., Sheppard S., 2004, The Astronomical Journal, 127, 1784 Jewitt D., Sheppard S., Fernndez Y., 2003, The Astronomical Journal, 125, 3366 Jewitt D., Hsieh H., Agarwal J., 2015, Asteroids IV, Patrick Michel, Francesca E. DeMeo, and William F. Bottke (eds.), University of Arizona Press, Tucson, 895 pp. ISBN: 978-0- 816-53213-1, 2015., p.221-241, pp 221 -- 241 Jorda L., Lamy P., Groussin O., Toth I., A'Hearn M. F., Peschke S., 2000, ISO Beyond Point Sources: Studies of Extended In- frared Emission, September 14-17, 1999, ISO Data Centre, Villafranca del Castillo, Madrid, Spain. Edited by R. J. Lau- reijs, K. Leech and M. F. Kessler, ESA-SP 455, 2000. p. 61., 455, 61 Kaiser N., et al., 2002, Jorda L., et al., 2016, Icarus, 277, 257 Jutzi M., Benz W., Toliou A., Morbidelli A., Brasser R., 2017, Astronomy & Astrophysics, Volume 597, id.A61, 13 pp., 597 in Tyson J. A., Wolff S., eds, and Other Telescope Technolo- Vol. gies and Discoveries. Edited by Tyson, J. Anthony; Wolff, Sidney. Proceedings of the SPIE, Volume 4836, pp. doi:10.1117/12.457365, http://proceedings.spiedigitallibrary.org/proceeding. aspx?doi=10.1117/12.457365 154-164 4836, Survey (2002).. p. 154, Kaiser N., et al., 2010, in Stepp L. M., Gilmozzi R., Hall H. J., eds, Vol. 7733, Ground-based and Airborne Telescopes III. Edited by Stepp, Larry M.; Gilmozzi, Roberto; Hall, Helen J. Proceedings of the SPIE, Volume 7733, article id. 77330E, 14 pp. (2010).. p. 77330E, doi:10.1117/12.859188, http://proceedings.spiedigitallibrary.org/proceeding. aspx?doi=10.1117/12.859188 Kamoun P. G., Campbell D. B., Ostro S. J., Pettengill G. H., Shapiro I. I., 1982, Science, 216, 293 Kasuga T., Balam D. D., Wiegert P. A., 2010, The Astronomical Journal, 140, 1806 Keller H. U., Mottola S., Skorov Y., Jorda L., 2015, Astronomy I., 2007, Space Science Reviews, 128, 1 & Astrophysics, 579, L5 Groussin O., Lamy P., Jorda L., Toth I., 2004, Astronomy and Astrophysics, 419, 375 Kelley M. S., Wooden D. H., Tubiana C., Boehnhardt H., Wood- ward C. E., Harker D. E., 2009, The Astronomical Journal, MNRAS 000, 1 -- 38 (2017) Rotation of Cometary Nuclei 37 Volume 137, Issue 6, pp. 4633-4642 (2009)., 137, 4633 Knight M. M., Farnham T. L., Schleicher D. G., Schwieterman E. W., 2011, The Astronomical Journal, 141, 2 Knight M. M., Schleicher D. G., Farnham T. L., Schwieterman E. W., Christensen S. R., 2012, The Astronomical Journal, 144, 153 Knight M. M., Mueller B. E. A., Samarasinha N. H., Schleicher D. G., 2015, The Astronomical Journal, Volume 150, Issue 1, article id. 22, 14 pp. (2015)., 150 Knight M. M., Fitzsimmons A., Kelley M. S. P., Snodgrass C., 2016, The Astrophysical Journal Letters, Volume 823, Issue 1, article id. L6, 6 pp. (2016)., 823 Lamy P. L., Toth I., Jorda L., Weaver H. A., A'Hearn M., 1998a, Astronomy and Astrophysics, v.335, p.L25-L29 (1998), 335, L25 Lamy P. L., Toth I., Weaver H. A., 1998b, Astronomy and Astro- physics, v.337, p.945-954 (1998), 337, 945 Snodgrass C., Hsieh H. H., Hainaut O., 2012, Astronomy & Astrophysics, 548, A12 Luu J., Jewitt D., 1990, Icarus, 86, 69 Luu J. X., Jewitt D. C., 1992, The Astronomical Journal, 104, 2243 Manzini F., Oldani V., Crippa R., Borrero J., Bryssink E., Mob- berley M., Nicolas J., 2014, Astrophysics and Space Science, 351, 435 Masoumzadeh N., et al., 2017, Astronomy & Astrophysics, 599, A11 Meech K., Hainaut O., Marsden B., 2004, Icarus, 170, 463 Meech et al., 2005, Science, 310, 265 Meech K. J., et al., 2009, American Astronomical Society, DPS meeting #41, id.20.07, 41 Meech K., et al., 2011a, Icarus, 213, 323 Meech K. J., et al., 2011b, The Astrophysical Journal, 734, L1 Millis R. L., A'Hearn M. F., Campins H., 1988, The Astrophysical Lamy P., Toth I., A'Hearn M. F., Weaver H. A., Weissman P. R., Journal, 324, 1194 2001, Icarus, 154, 337 Lamy P., Toth I., Jorda L., Groussin O., A'Hearn M. F., Weaver H. A., 2002, Icarus, 156, 442 Lamy P. L., Toth I., Fernandez Y. R., Weaver H. A., 2004, Comets II Lamy P. L., Toth I., Weaver H. A., Jorda L., Kaasalainen M., Guti´errez P. J., 2006, Astronomy and Astrophysics, 458, 669 Lamy P. L., Toth I., Weaver H. A., A'Hearn M. F., Jorda L., Mottola S., et al., 2014, Astronomy & Astrophysics, 569, L2 Mueller B. E. A., 1992, In Lunar and Planetary Inst., Asteroids, Comets, Meteors 1991 p 425-428 (SEE N93-19113 06-90) Mueller B. E., Ferrin I., 1996, Icarus, 123, 463 Mueller B. E. A., Samarasinha N. H., 2002, Earth, Moon, and Planets, v. 90, Issue 1, p. 463-471 (2002)., 90, 463 Mueller B. E. A., Samarasinha N. H., 2015, American Astronom- ical Society 2009, Astronomy and Astrophysics, 508, 1045 Mueller B. E. A., Samarasinha N. H., Fernandez Y. R., 2008, Lamy P. L., Toth I., Weaver H. A., A'Hearn M. F., Jorda L., 2011, Monthly Notices of the Royal Astronomical Society, 412, 1573 Lamy P., Faury G., Llebaria A., Knight M., A'Hearn M., Battams K., 2013, Icarus, 226, 1350 Leibowitz E. M., Brosch N., 1986, Icarus, 68, 430 Levison H. F., 1996, in Completing the Inventory of the Solar System, Astronomical Society of the Pacific Conference Pro- ceedings, volume 107, T.W. Rettig and J.M. Hahn, Eds., pp. 173-191.. Astronomical Society of the Pacific (ASP), pp 173 -- 191, http://adsabs.harvard.edu/abs/1996ASPC..107..173L Levison H. F., Terrell D., Wiegert P. A., Dones L., Duncan M. J., 2006, Icarus, 182, 161 Li J.-Y., et al., 2007a, Icarus, 187, 41 Li J., A'Hearn M., McFadden L., Belton M., 2007b, Icarus, 188, 195 Li J.-Y., A'Hearn M. F., Farnham T. L., McFadden L. A., 2009, Icarus, 204, 209 Li J.-Y., et al., 2013, Icarus, 222, 559 Licandro J., Tancredi G., Lindgren M., Rickman H., Hutton R. G., 2000, Icarus, 147, 161 Lisse C. M., et al., 2005, The Astrophysical Journal, 625, L139 Lisse C. M., et al., 2009, Publications of the Astronomical Society of Pacific, Volume 121, Issue 883, pp. 968-975 (2009)., 121, 968 Lomb N. R., 1976, Astrophysics and Space Science, 39, 447 Lowry S. C., Fitzsimmons A., 2001, Astronomy and Astrophysics, 365, 204 Lowry S. C., Weissman P. R., 2003, Icarus, 164, 492 Lowry S. C., Weissman P. R., 2007, Icarus, 188, 212 Lowry S. C., Fitzsimmons A., Cartwright I. M., Williams I. P., 1999, Astronomy and Astrophysics, v.349, p.649-659 (1999), 349, 649 American Astronomical Society, 40 Mueller B. E. A., Farnham T. L., Samarasinha N. H., A'Hearn M. F., 2010a, American Astronomical Society, DPS meeting #42, id.28.31; Bulletin of the American Astronomical Society, Vol. 42, p.966, 42, 966 Mueller B. E. A., Samarasinha N. H., Rauer H., Helbert J., 2010b, Icarus, 209, 745 Nolan M. C., Harmon J. K., Howell E. S., Benner L. A., Giorgini J. D., Ostro S. J., Campbell D. B., Margot J. L., 2006, Amer- ican Astronomical Society, DPS meeting #38, id.12.06; Bul- letin of the American Astronomical Society, Vol. 38, p.504, 38, 504 Ohtsuka K., Ito T., Yoshikawa M., Asher D. J., Arakida H., 2008, Astronomy and Astrophysics, Volume 489, Issue 3, 2008, pp.1355-1362, 489, 1355 Oszkiewicz D. A., Bowell E., Wasserman L. H., Muinonen K., Penttila A., Pieniluoma T., Trilling D. E., Thomas C. A., 2012, Icarus, Volume 219, Issue 1, p. 283-296., 219, 283 Pravec P., Harris A. W., 2000, Icarus, 148, 12 Pravec P., Sarounova L., Wolf M., 1996, Icarus, 124, 471 Pravec P., Harris A. W., Michalowski T., 2002, Asteroids III, W. F. Bottke Jr., A. Cellino, P. Paolicchi, and R. P. Binzel (eds), University of Arizona Press, Tucson, p.113-122, pp 113 -- 122 Reyniers M., Degroote P., Bodewits D., Cuypers J., Waelkens C., 2009, Astronomy and Astrophysics, Volume 494, Issue 1, 2009, pp.379-389, 494, 379 Richardson J. E., Melosh H. J., Lisse C. M., Carcich B., 2007, Icarus, 190, 357 Samarasinha N. H., Mueller B. E. A., 2013, The Astrophysical Journal, 775, L10 Samarasinha N. H., Mueller B. E. A., Belton M. J. S., Jorda L., Lowry S. C., Fitzsimmons A., Collander-Brown S., 2003, Astron- 2004, Comets II omy and Astrophysics, 397, 329 Lowry S. C., Fitzsimmons A., Jorda L., Kaasalainen M., Lamy P., Toth I., 2006, American Astronomical Society, 38 Lowry S., Fitzsimmons A., Lamy P., Weissman P., 2008, in , The Solar System Beyond Neptune, M. A. Barucci, H. Boehn- hardt, D. P. Cruikshank, and A. Morbidelli (eds.), University of Arizona Press, Tucson, 592 pp., p.397-410. pp 397 -- 410, http://adsabs.harvard.edu/abs/2008ssbn.book..397L Lowry S., Duddy S. R., Rozitis B., Green S. F., Fitzsimmons A., Samarasinha N. H., Mueller B. E. A., A'Hearn M. F., Farnham T. L., 2010, IAU Circ., No. 9178, #1 (2010). Edited by Green, D. W. E., 9178 Samarasinha N. H., Mueller B. E. A., A'Hearn M. F., Farnham T. L., Gersch A., 2011, The Astrophysical Journal, 734, L3 Samarasinha N. H., et al., 2012, American Astronomical Society, DPS meeting #44, id.506.03, 44 Scargle J. D., 1982, The Astrophysical Journal, 263, 835 Schleicher D. G., Knight M. M., 2016, eprint arXiv:1605.01705 MNRAS 000, 1 -- 38 (2017) 38 R. Kokotanekova et al. Schleicher D. G., Knight M. M., Levine S. E., 2013, The Astro- nomical Journal, Volume 146, Issue 5, article id. 137, 8 pp. (2013)., 146 Scotti J. V., et al., 1996, IAU Circ., No. 6301, #1 (1996). Edited by Marsden, B. G., 6301 Sekanina Z., 1987, In ESA, 278 Sekanina Z., Zdenek 1991, The Astronomical Journal, 102, 350 Sekanina Z., Brownlee D. E., Economou T. E., Tuzzolino A. J., Green S. F., 2004, Science, 304, 1769 Sierks H., et al., 2015, Science (New York, N.Y.), 347, aaa1044 Snodgrass C., Carry B., 2013, Automatic Removal of Fringes from EFOSC Images. Vol. 152, European Southern Observatory, http://adsabs.harvard.edu/abs/2013Msngr.152...14S Snodgrass C., Fitzsimmons A., Lowry S. C., 2005, Astronomy and Astrophysics, 444, 287 Snodgrass C., Lowry S. C., Fitzsimmons A., 2006, Monthly No- tices of the Royal Astronomical Society, 373, 1590 Snodgrass C., Saviane I., Monaco L., Sinclaire P., 2008a, The Messenger, vol. 132, p. 18-19, 132, 18 Snodgrass C., Lowry S. C., Fitzsimmons A., 2008b, Monthly No- tices of the Royal Astronomical Society, 385, 737 Snodgrass C., Fitzsimmons A., Lowry S. C., Weissman P., 2011, Monthly Notices of the Royal Astronomical Society, 414, 458 Snodgrass C., et al., 2016, Astronomy & Astrophysics, Volume 588, id.A80, 12 pp., 588 Soderblom L. A., et al., 2002, Science, 296, 1087 Stellingwerf R. F., 1978, The Astrophysical Journal, 224, 953 Storm S. P., et al., 2006, American Astronomical Society, DPS meeting #38, id.12.08; Bulletin of the American Astronomical Society, Vol. 38, p.504, 38, 504 Tancredi G., Fern´andez J. A., Rickman H., Licandro J., 2000, Astronomy and Astrophysics Supplement Series, 146, 73 Thomas P., et al., 2013a, Icarus, 222, 453 Thomas P. C., et al., 2013b, Icarus, 222, 550 Tody D., 1986, in Crawford D. L., ed., Vol. 627, IN: Instru- mentation in astronomy VI; Proceedings of the Meeting, Tucson, AZ, Mar. 4-8, 1986. Part 2 (A87-36376 15-35). Bellingham, WA, Society of Photo-Optical Instrumentation Engineers, 1986, p. 733.. pp 733 -- 748, doi:10.1117/12.968154, http://proceedings.spiedigitallibrary.org/proceeding. aspx?articleid=1242189 Tody D., 1993, Astronomical Data Analysis Software and Systems II, A.S.P. Conference Series, Vol. 52, 1993, R. J. Hanisch, R. J. V. Brissenden, and Jeannette Barnes, eds., p. 173., 52, 173 Tonry J. L., et al., 2012, The Astrophysical Journal, 750, 99 Toth I., Lisse C., 2006, Icarus, 181, 162 Toth I., Lamy P., Weaver H., 2005, Icarus, 178, 235 Tubiana C., Barrera L., Drahus M., Boehnhardt H., 2008, As- tronomy and Astrophysics, 490, 377 Tubiana C., Bohnhardt H., Agarwal J., Drahus M., Barrera L., Ortiz J. L., 2011, Astronomy & Astrophysics, 527, A113 Tubiana C., Snodgrass C., Michelsen R., Haack H., Boehnhardt H., Fitzsimmons A., Williams I. P., 2015, Astronomy & As- trophysics, Volume 584, id.A97, 10 pp., 584 VanderPlas J. T., Ivezic Z., 2015, The Astrophysical Journal, Vol- ume 812, Issue 1, article id. 18, 15 pp. (2015)., 812 Volk K., Malhotra R., 2008, The Astrophysical Journal, 687, 714 Warner B. D., 2006, The Minor planet bulletin bulletin of the Mi- nor Planets Section of the Association of Lunar and Planetary Observers : MPB.. Vol. 33, Univ, http://adsabs.harvard. edu/abs/2006MPBu...33...35W Warner B. D., Fitzsimmons A., 2005, IAU Circ., No. 8578, #1 (2005). Edited by Green, D. W. E., 8578 Weaver H. A., Stern S. A., Parker J. W., 2003, The Astronomical Journal, 126, 444 Weissman P. R., Doressoundiram A., Hicks M. D., Chamberlin A., Sykes M. V., Larson S., Hergenrother C., 1999, Bulletin of the Astronomical Society, Vol. 31, No. 4, p. 1121, id.30.03, 31, 1121 Williams G. V., 2017, Central Bureau Electronic Telegrams, 4359, 1 (2017). Edited by Green, D. W. E., 4359 This paper has been typeset from a TEX/LATEX file prepared by the author. MNRAS 000, 1 -- 38 (2017)
1107.5309
1
1107
2011-07-26T20:00:03
A Southern Sky and Galactic Plane Survey for Bright Kuiper Belt Objects
[ "astro-ph.EP", "astro-ph.IM", "astro-ph.SR" ]
About 2500 square degrees of sky south of declination -25 degrees and/or near the galactic plane were surveyed for bright outer solar system objects. This survey is one of the first large scale southern sky and galactic plane surveys to detect dwarf planets and other bright Kuiper Belt objects in the trans-Neptunian region. The survey was able to obtain a limiting R-band magnitude of 21.6. In all, 18 outer solar system objects were detected, including Pluto which was detected near the galactic center using optimal image subtraction techniques to remove the high stellar density background. Fourteen of the detections were previously unknown trans-Neptunian objects, demonstrating that the southern sky had not been well-searched to date for bright outer solar system objects. Assuming moderate albedos, several of the new discoveries from this survey could be in hydrostatic equilibrium and thus be considered dwarf planets. Combining this survey with previous surveys from the northern hemisphere suggests that the Kuiper Belt is nearly complete to around 21st magnitude in the R-band. All the main dynamical classes in the Kuiper Belt are occupied by at least one dwarf planet sized object. The 3:2 Neptune resonance, which is the innermost well-populated Neptune resonance, has several large objects while the main outer Neptune resonances such as the 5:3, 7:4, 2:1, and 5:2 do not appear have any large objects. This indicates that the outer resonances are either significantly depleted in objects relative to the 3:2 resonance or have a significantly different assortment of objects than the 3:2 resonance. For the largest objects (H<4.5 mag), the scattered disk population appears to have a few times more objects than the main Kuiper Belt population, while the Sedna population could be several times more than that of the main Kuiper Belt.
astro-ph.EP
astro-ph
A Southern Sky and Galactic Plane Survey for Bright Kuiper Belt Objects Scott S. Sheppard1, Andrzej Udalski2, Chadwick Trujillo3, Marcin Kubiak2, Grzegorz Pietrzynski2, Radoslaw Poleski2, Igor Soszynski2, Michal K. Szyma´nski2, and Krzysztof Ulaczyk2 ABSTRACT About 2500 square degrees of sky south of declination -25 degrees and/or near the galactic plane were surveyed for bright outer solar system objects. This survey is one of the first large scale southern sky and galactic plane surveys to detect dwarf planets and other bright Kuiper Belt objects in the trans-Neptunian region. The survey was able to obtain a limiting R-band magnitude of 21.6. In all, 18 outer solar system objects were detected, including Pluto which was detected near the galactic center using optimal image subtraction techniques to remove the high stellar density background. Fourteen of the detections were previously unknown trans-Neptunian objects, demonstrating that the southern sky had not been well-searched to date for bright outer solar system objects. Assuming moderate albedos, several of the new discoveries from this survey could be in hydrostatic equilibrium and thus be considered dwarf planets. Combining this survey with previous surveys from the northern hemisphere suggests that the Kuiper Belt is nearly complete to around 21st magnitude in the R-band. All the main dynamical classes in the Kuiper Belt are occupied by at least one dwarf planet sized object. The 3:2 Neptune resonance, which is the innermost well-populated Neptune resonance, has several large objects while the main outer Neptune resonances such as the 5:3, 7:4, 2:1, and 5:2 do not appear have any large objects. This indicates that the outer resonances are either significantly depleted in objects relative to the 3:2 resonance or have a significantly different assortment of objects than the 3:2 resonance. For the largest objects (H < 4.5 mag), the scattered disk population appears to have a few times more objects than the main Kuiper Belt population, while the Sedna population could be several times more than that of the main Kuiper Belt. 1Department of Terrestrial Magnetism, Carnegie Institution of Washington, 5241 Broad Branch Rd. NW, Washington, DC 20015, USA, [email protected] 2Warsaw University Observatory, Al. Ujazdowskie 4, 00-478 Warszawa, Poland 3Gemini Observatory, 670 North A'ohoku Place, Hilo, HI 96720, USA – 2 – Subject headings: Kuiper Belt – Oort Cloud – comets: general – minor planets, asteroids – solar system: general – planetary formation 1. Introduction The strong dynamical connection that the trans-Neptunian objects (TNOs) have to the planets makes determining their population and orbital structures valuable for gaining insight into solar system formation and planet evolution. The Kuiper Belt, a remnant of the original protoplanetary disk, has a "fossilized" record of the original solar nebula and subsequent evolution of the solar system. TNOs are likely primitive with significant amounts of volatiles. The largest TNOs or dwarf planet sized objects are rare but extremely important for several reasons: 1) The brightest few objects are the only ones accessible to high signal to noise spectroscopy techniques that are required to determine surface compositions, such as methane and water ice (Barucci et al. 2008; Trujillo et al. 2011). These physical characteristics are important in order to understand the formation, origin and composition of the objects and gain insight into planet formation and chemistry in the original solar nebula. 2) The size distribution of the biggest objects in the Kuiper Belt determines if the mass in the Kuiper Belt is dominated by the largest or smallest objects, which is a key metric of planetismal growth scenarios (Kenyon et al. 2008, 2010; Cuzzi et al. 2010). The size and number of the biggest objects constrain the density and thus planet formation ability of the original solar nebula in the outer solar system. 3) Occultations of stars by the biggest TNOs are possible to predict and observe in order to probe the TNOs sizes, shapes, albedos, and atmospheres (Elliot and Kern 2003; Elliot et al. 2010). The Palomar 48 inch Schmidt telescope in the northern hemisphere, with one of the largest CCD cameras in the world, was used to survey most of the sky north of -25 degrees declination for the brightest (mR . 21 mags) TNOs (Trujillo and Brown 2003; Brown et al. 2004, 2005; Brown 2008; Schwamb et al. 2009, 2010). In these surveys tens of bright TNOs including likely dwarf planets Eris, Makemake, Haumea, Orcus, Quaoar, Sedna and 2007 OR10 were discovered. These surveys showed that many of the largest Kuiper Belt Objects (KBOs) have relatively large inclinations with the vast majority of KBOs expected to be found within about 20 degrees of the ecliptic (Brown 2008). Extrapolating the Cumulative Luminosity Function (CLF) to the bright end of the KBOs indicates several large KBOs should be discovered in the southernmost parts of the sky that the surveys from the northern telescopes did not image. The southern hemisphere has not been well-surveyed for distant solar system objects until now because in the past there were no sensitive, wide-field digital imagers on suitable – 3 – telescopes in the south. This changed in 2009 when a large wide-field imager was put onto the 1.3 meter Warsaw telescope at Las Campanas in Chile. The OGLE-Carnegie Kuiper Belt Survey (OCKS) was implemented to search the Kuiper Belt for dwarf planets and bright TNOs through a shallow survey to fainter than 21st magnitude in the R-band from the southern hemisphere. OCKS covered the area within a few tens of degrees of the ecliptic for declinations less than -25 degrees and the crowded galactic plane fields in the north and south. Another independent southern sky survey for KBOs was started in late 2009 with the Schmidt telescope at La Silla (Rabinowitz 2010). This is the first time most of this sky area was searched for outer solar system objects with modern digital CCD detectors. 2. Observations The vast majority of the survey fields were obtained with the Warsaw 1.3 meter telescope at Las Campanas observatory in Chile. The telescope is also known as the OGLE telescope (Optical Gravitational Lensing Experiment; Udalski et al. 1994) and OCKS is considered part of the OGLE-IV project. OGLE-IV commenced with the successful commissioning of the new wide-field 1.4 square degree imager at the beginning of 2010. The southern sky Kuiper Belt survey observations at the Warsaw telescope occurred between March and September 2010 while the northern galactic plane fields near the ecliptic were imaged in December 2010 and January 2011. The 1.4 square degree imager has 32 E2V44-82 2048×4102 CCD chips with 0.′′26/pixel. There are four rows and 9 columns of chips. Gaps are generally only a few arcseconds between chips except between the first and second rows and third and fourth rows the gaps are a bit wider at several tens of arcseconds. Readout time for the detector is about 20 seconds. All fields were within about 2.5 hours of opposition with most being within 1.5 hours. At these opposition distances, the apparent motion of an outer solar system object is dominated by the parallax from the Earth's movement, making confusion of outer solar system objects with foreground main belt asteroids minimal (Luu and Jewitt 1988). Las Campanas is a very dark site with excellent seeing conditions (Thomas-Osip et al. 2011). Most images were obtained with the seeing around 1 arcsecond or less. If the seeing was much worse than 1 arcsecond or if the conditions were not photometric on a given night, observations were not taken. Integrations were 180 seconds with the telescope tracking at sidereal rates. Since there was no preferred VR or R-band filter for the 1.4 square degree imager, a V-band filter was used at the start of the survey for fields West of the Galactic plane. Because of the better seeing conditions in the I-band, the I-band filter was used for fields in the Galactic plane as well as fields East of the Galactic plane. It was found that the V-band and I-band – 4 – images obtained similar depths but the I-band was preferred since it was less sensitive to moderate moon brightness. Image reduction was performed by first bias subtracting and then flat-fielding the images. In addition to the Warsaw data, about 100 square degrees were surveyed using the CTIO 4 meter Blanco telescope with its MOSAIC II camera that covers about a third of a square degree. These data were obtained in June 2009 and 2010 in order to see how well such a program would work on the 4 meter telescope. Images were only 20 seconds in length and reached magnitudes of about 22 in the R-band. Recovery was mostly done at the Warsaw 1.3 meter telescope but some recovery took place at the CTIO 4 meter and Magellan 6.5 meter. 3. Analysis In total about 2500 square degrees of sky were surveyed in the southern hemisphere or near the galactic plane (Figure 1). Each survey field had at least two hours between the first and last image of a three image sequence. Outer solar system objects were searched for in the survey fields in two complementary ways. One technique used a computer algorithm specifically designed to detect the apparent motion of trans-Neptunian objects (Trujillo and Jewitt 1998; Sheppard and Trujillo 2010) while a second technique used a differencing algo- rithm (Udalski et al. 1997,2003; Wozniak 2000) on the three images in order to remove the steady state of background stars to look for moving or transient objects. The differencing algorithm was used on all fields and was the only technique used on fields within 15 degrees of the galactic plane. Both computer algorithms were calibrated to detect moving objects that appeared in all three images from one night and had a motion consistent with being beyond 10 AU (motion slower than about 10 arcseconds per hour). Because of the fine pixel scale and relatively good seeing, the survey was sensitive to objects moving as slow as 0.5 arcseconds per hour. This apparent motion corresponds to objects out to about 300 AU. Since the survey covers many nights, the data from night to night is not all of the same quality. In order to make the data as consistent as possible over the nights, survey fields were only taken in moderate seeing (∼ 1 arcsecond) or better conditions and only when conditions were photometric. If the seeing was significantly worse than about one arcsecond, the survey was not continued for that night. The limiting magnitude of the survey was determined by placing artificial objects in the fields matched to the point spread function of the images with motions mimicking that – 5 – of a TNO (4 to 0.5 arcseconds per hour). A 50% detection efficiency at an R-band limiting magnitude of about 21.6 magnitudes was found for fields with good seeing conditions about 15 degrees or more from the galactic plane (Figure 2). For fields with moderate seeing conditions the R-band limiting magnitude was found to be about 21.2 magnitudes, where the typical color of a moderately red KBO was used to convert the I-band survey fields to the R-band (R-I=0.5 mags) in order to better compare the survey with previous survey depths. For images near the galactic plane the stellar confusion would limit the detection of moving solar system objects in previous surveys. In this survey the optimal PSF matching image subtraction techniques developed by Alard and Lupton (1998) and Alard (2000) and implemented through the previous OGLE phases were used (Wozniak 2000; Wozniak et al. 2001; Udalski 2003). PSF matching and image subtraction removed the stellar confusion from the galactic plane. Thus, this is the first survey to be sensitive to TNOs near the galactic center where the ecliptic plane crosses the galaxy. To test the moving object algoritm with differenced images, Pluto was observed early in the survey and easily found in the dense galactic plane (Figures 3 and 4). The survey depth near the galactic center was similar to the depth of the fields off the galactic plane, but the survey efficiency of detection was decreased by about 15 percent. 4. Results and Discussion 4.1. Completion Limits of the Kuiper Belt Eighteen outer solar system objects were detected in this survey. Fourteen of these objects were new discoveries showing that this region of sky had not been well-searched for bright, distant objects in the past (Tables 1 and 2). Combining this southern sky and galactic plane survey with the previous large area northern sky surveys (Trujillo and Brown 2003; Brown 2008; Schwamb et al. 2009, 2010) and a recent large Kuiper Belt survey in the south started by Rabinowitz (2010) makes it likely that the Kuiper Belt is now nearly complete to about 21st magnitude in the R-band. To date, only three areas have not been well searched for bright outer solar system objects: 1) southern fields very distant from the ecliptic (> 20 degrees ecliptic latitude) and thus unlikely to harbor many bright KBOs, 2) a few hundred square degrees in the northern section of the north galactic plane near the ecliptic and 3) a few hundred square degrees in the northern section of the south galactic plane near the ecliptic (Figure 1). There are around 70 known TNOs with apparent magnitudes brighter than 21 in the R-band (see section 4.2), almost all within 20 degrees of the ecliptic. Thus, there is on average one KBO brighter than 21st mag every few hundred square degrees of sky near the ecliptic. This means there is likely to be only one or two KBOs brighter than – 6 – 21st magnitude in the few remaining areas yet to be searched. Though nearly complete to 21st magnitude now, some objects, especially Centaurs and scattered disk objects, have large eccentricities and thus could become brighter than 21st magnitude in the future as they approach perihelion. The size of an object at the completeness limit depends on the distance and albedo of the object (Figure 5). The largest few objects, with radii greater than about 500 km, have been found to have very high albedos (ρR ∼ 0.6 − 0.8), while smaller objects appear to have moderate albedos (ρR ∼ 0.1 − 0.2) (Stansberry et al. 2008). The high albedos of the largest objects is likely due to atmospheres and/or surface processes such as cryovolcanism (Licandro et al. 2006; Dumas et al. 2007; Rabinowitz et al. 2007; Sheppard 2007). Assuming a typical albedo of ρR = 0.15 for the moderate sized KBOs of 21st magnitude, the completeness limit at 30 AU is about 80 km in radius, while at 50 AU it is about 225 km in radius (Figure 6). In absolute magnitude, H, this would be 6.6 and 4.4 magnitudes respectively (Figure 7). It is clear that further Pluto or larger sized objects could remain undetected if beyond a few hundred AU. 4.2. Size Distribution Figure 8 shows the cumulative number of all known TNOs versus their absolute magni- tude, H. Objects with absolute magnitudes H > 7 mags appear to have a roll-over in their size distribution because of detection biases. The largest KBOs with H < 3 mags do not follow the simple power-law found for the objects with 3 < H < 7 mags (Brown 2008). The largest KBOs have been found to have preferentially higher albedos, likely because of atmo- sphere effects and surface activity that keep the surfaces young and bright (Jewitt and Luu 2004; Lykawka and Mukai 2005; Schaller and Brown 2007; Stansberry et al. 2008; Desch et al. 2009). The absolute magnitudes the largest KBOs would have if they had a more typical albedo of 0.15 (Brown and Trujillo 2004; Brown et al. 2006; Stansberry et al. 2008) are shown by squares in Figure 8. The squares in Figure 8 fit a simple power-law for all objects with H < 7 magnitudes (r < 60 km assuming 0.15 albedo). The points in a cumulative distribution are heavily correlated with one another, tending to give excess weight to the faint end of the distribution. A differential distribution does not suffer from this problem. Figure 9 shows the differential number of all known TNOs versus their absolute magnitude, where, like in Figure 8, the largest few objects have had their absolute magnitudes adjusted for their abnormally high albedos compared to smaller objects. It is clear there is a turnover around an absolute magnitude of 7 mags (r ∼ 60 km) showing observational bias beyond this magnitude. The best fit power-law for the differential – 7 – points finds q = 3.0 ± 0.5 for H < 7 magnitudes, where n(r)dr ∝ r−qdr is the differential power-law radius distribution with n(r)dr describing the number of TNOs with radii in the range r to r+dr. This is slightly lower than most previous fits (q ∼ 4) that were more heavily dependent on fainter (smaller) objects (Jewitt et al. 1998; Trujillo et al. 2001; Petit et al. 2008; Fraser et al. 2008; Fuentes and Holman 2008; Fraser and Kavelaars 2009; Fuentes et al. 2009). As the scattered disk and Sedna populations are not close to completion on the large end (H < 4.5 mags), including such objects (Eris, 2007 OR10 and Sedna), as done here, likely results in a shallower measured slope. The size distributions of individual dynamical classes are likely more informative (see section 4.2.2). There are no obvious discontinuities at the large end of the KBO size distribution when including all dynamical classes of TNOs (Figures 8 and 9). 4.2.1. Dwarf Planets A dwarf planet is defined by the International Astronomical Union (IAU) as an object that is in hydrostatic equilibrium and has not cleared the neighborhood around its heliocen- tric orbit of other similarly sized objects. Though the dwarf planet definition is imprecise, it is clear that Ceres in the main asteroid belt as well as Pluto and Eris in the outer solar system are bonafide dwarf planets. Makemake and Haumea are also likely dwarf planets as are the next largest bodies in the outer solar system such as Sedna, 2007 OR10, Orcus and Quaoar. Though the lower size limit of an object in hydrostatic equilibrium is not well defined, Lineweaver and Norman (2010) suggest it could be as small as 200 km in radius for an icy body in the outer solar system. This would put tens more objects in the outer solar system into the dwarf planet category, including three objects discovered in this survey (Table 1: 2010 EK139, 2010 KZ39 and 2010 FX86). The actual sizes and shapes of these bodies are not well known to date and will depend heavily on their albedos and compositions. Further detailed observations are required to determine the true sizes and shapes of the new discoveries. With most of the biggest Kuiper Belt objects likely known, it is interesting to compare where the largest (H ≤ 4.5 mags) objects reside dynamically in the Kuiper Belt (Figure 10). At least one of the largest objects can be found in most of the TNO dynamical populations (Tables 3 and 4). The scattered disk population (Gomes et al. 2008) has Eris and 2007 OR10, while Sedna is in its own dynamical class (Morbidelli and Levison 2004; Gladman and Chan 2006) that resides significantly beyond the Kuiper Belt edge (Trujillo and Brown 2001; Allen et al. 2001). The high inclination classical Kuiper belt (Gomes 2003) has several large objects including Makemake, Haumea, Varuna and (278361) 2007 JJ43. Even the low – 8 – inclination classical Kuiper belt population, generally known for its smaller sized objects (Levison and Stern 2001), appears to have Quaoar. Further confirming Quaoar's status as a low inclination Kuiper belt object is Quaoar's ultra-red surface (Jewitt and Luu 2004), which is a characteristic generally associated with the low inclination classical Kuiper Belt (Tegler and Romanishin 2000; Trujillo and Brown 2002; Stern 2002; Doressoundiram et al. 2008; Peixinho et al. 2008). The actual number of Pluto sized bodies is now known (Table 3). Previous authors have argued that the Kuiper Belt likely lost a substantial amount of its mass through collisional grinding and dynamical interactions with the planets (Kenyon and Luu 1999; Levison et al. 2008; Morbidelli et al. 2008; Stewart and Leinhardt 2009). Observationally, many more objects appear to be required in order to produce the observed angular momentum of the largest KBOs (Jewitt and Sheppard 2002; Rabinowitz et al. 2006) and binaries (Noll et al. 2008). Detailed simulations show that Kuiper Belt formation is possible with only the small number of Pluto sized objects observed (Kenyon and Bromley 2008; Schlichting and Sari 2011). A significant number of Pluto sized objects likely exist in the populations beyond 100 AU such as the Sedna types and Oort cloud objects, which are currently too faint to be efficiently detected to date. It is important to determine if the Pluto sized objects formed in the Kuiper Belt as we see it today or if they originated much closer to the Sun and were later transported to their current orbits. 4.2.2. TNO Population Ratios On the large size end (H . 4.5 mags), the ratio of the (Plutinos):(Main Kuiper Belt):(Scattered Disk):(Sedna Types) was found to be (1) : (2.6) : (7 ± 3) : (75±+115 −55 ), respectively (Table 4). Thus the Sedna population could be the dominant observed small body population for dwarf sized planets (Figure 11). The scattered disk population is likely bigger than the main Kuiper Belt (MKB) population by a factor of a few. The Plutino pop- ulation is smaller by a factor of a few compared to the main Kuiper Belt. Both the scattered disk and main Kuiper Belt populations on the large end of the size distribution (H . 4.5 mags) are consistent with q = 3.3 ± 0.7 while the Plutino population appears significantly shallower than this with q = 2.2 ± 0.5. The scattered disk population size determined from the largest objects (H < 4.5 mags) is consistent with Trujillo et al. (2001) estimated from smaller objects in the scattered disk when using a q ∼ 3.3 size distribution. – 9 – 4.2.3. The Main Kuiper Belt The main Kuiper Belt (39 < a < 48 AU) appears to be divided into three distinct dynamical classes (Figure 12). The high inclination and low inclination ("cold") classical classes have been suggested for a decade, with the largest objects preferentially in high inclination orbits (Levison and Stern 2001; Brown 2001). When plotting only the largest few objects (H < 4.5 mags), there appears to also be both low eccentricity and higher eccentricity classes (Figure 10). All three of the low inclination objects (i < 10 degs) with H < 4.5 mags have low eccentricities (e < 0.05). The high inclination objects with H < 4.5 mags in the main Kuiper Belt appear to have either low eccentricities (0.03 < e < 0.07; 6 observed) or significantly higher eccentricities (0.13 < e < 0.16; 8 observed). Only one of the twenty main Kuiper Belt objects with H < 4.5 mags has an eccentricity between these two ranges (Salacia (120347) 2004 SB60 which has e = 0.10) while two others have slightly higher eccentricities (Haumea with e = 0.20 and (230965) 2004 XA192 with e = 0.25). The Hartigan and Hartigan (1985) dip test for bimodality shows a strong bimodality in eccentricity when including all main Kuiper Belt objects with H < 4.5 mags except for the interesting binary object Salacia (these nineteen objects give a dip statistic of 0.145 which corresponds to a confidence of 0.997 for bimodality, a 3 sigma result). Including Salacia in the dip test gives a less significant result of only 0.990 confidence in bimodality, or slightly less than 3 sigma. Including smaller main Kuiper Belt objects decreases the bimodality significance even further. If real, the low versus higher eccentricity populations of highly inclined large objects could have different origins, such as forming in different regions of the solar system or originally from different scattering events during the migration of the planets. The largest bodies (H < 4.5 mags) are too few for meaningful statistics, but it appears that the high inclination main Kuiper Belt does not have a significantly shallower power-law distribution than the low inclination population, as was found for the smaller objects of these two populations by Fraser et al. (2010) (Figure 12). There is a possible deficiency of objects in the main Kuiper Belt between 2.5 < H < 3.5 magnitudes, but this is likely not statistically significant as it is just small number statistics. 4.2.4. Resonance Populations A surprising result is the absence of large objects in all the main Neptune resonance populations except the 3:2 resonance (see Gladman et al. 2008 and Elliot et al. 2005 for resonance calculations as well as the updated version of Elliot et al. 2005 kept by Marc Buie – 10 – at www.boulder.swri.edu/buie/kbo/astrom). The Plutinos or 3:2 resonance objects include some of the largest known KBOs such as Pluto, Orcus and Ixion (Table 3) while the other observed heavily populated resonances such as the 5:3, 7:4, 2:1, and 5:2 have no known large KBOs (Table 4). Any object in the Neptune resonances brighter than 21st magnitude (r & 200 km), would likely have been detected by now (Figure 7 and Table 4). The 5:2 has a few sizable objects with absolute magnitudes of around 3.8 and 5.1 mags, (84522) 2002 TC302 and (26375) 1999 DE9, respectfully. The largest 2:1 object appears to be (119979) 2002 WC19 with an absolute magnitude of 5.1 mags. None of the other resonances have any objects with absolute magnitudes brighter than 5 mags. 2010 EK139, discovered in this survey, appears to be one of the only known objects in the very distant 7:2 resonance (based on orbit calculations from Marc Buie's website at www.boulder.swri.edu/buie/kbo/astrom that has up to date information first published in Elliot et al. 2005). The relative populations of the various Neptune resonances are currently not well con- strained since observational biases make discoveries easier in the closer 3:2 resonance (Jewitt et al. 1998; Trujillo et al. 2001). There is also a strong longitude and latitude dependence on discovery of resonance populations (Chiang and Jordan 2002; Chiang et al. 2003). Previous observational works have suggested that the 2:1 resonance appears to have less objects than the 3:2 resonance (Jewitt et al. 1998; Chiang and Jordan 2002). Numerical simulations of resonance sweeping (Hahn and Malhotra 2005) have shown that the main Neptune resonance populations relative to the 3:2 resonance population may be 2:1 (x2), 7:4 (x0.8), 5:3 (x0.6), 5:2 (x0.5). These simulations suggest that the outer resonances should have a factor of four more objects than the 3:2 resonance. Thus, if 3 very large objects such as Pluto, Orcus and Ixion were found in the 3:2 resonance population, based on Poisson statistics, one would expect 12 ± 3.5 objects of similar size in the other resonances. This is not the case, and such a scenario can be rejected with 3.5 sigma confidence, so either the outer resonances are significantly less populated than the 3:2 resonance or the outer resonance bodies have a different size distribution than the 3:2 resonance. It is likely that the 3:2 resonance is populated by objects that formed significantly closer to the Sun than the outer resonances. Objects forming closer to the Sun would likely accrete more material in a shorter amount of time allowing them to become larger before they were captured in the Neptune resonances. With the Kuiper Belt nearly complete to 21st magnitude, it is unlikely that a planet larger than Mercury within a few 100 AU currently exists. Lykawka and Mukai (2008) suggested such a planet could have disrupted the outer resonance populations. It is still possible that a close stellar encounter or now defunct outer planet could have disrupted or depleted the outer resonances early in the solar system's history. – 11 – 5. Summary The OGLE Carnegie Kuiper belt Survey (OCKS) is one of the first southern sky and galactic plane surveys for bright outer solar system objects. Eighteen bright Trans-Neptunian objects were discovered, including some of the most southern outer solar system objects ever detected as well as the intrinsically brightest solar system objects discovered in several years (2010 EK139 with H = 3.8 and 2010 KZ39 with H = 3.9 mags). 1) A total of 2500 square degrees was searched in the survey. About 2200 square degrees of the survey was south of declination -25 degrees, where northern KBO surveys cannot efficiently observe. The surveyed area includes almost all of the southern sky within about 20 degrees of the ecliptic. Another 300 square degrees of sky was surveyed in the northern galactic plane near the ecliptic using optimal image subtraction techniques to remove the stellar background. 2) The survey obtained a limiting R-band magnitude of 21.6 during optimal observing conditions using the 1.3 meter Warsaw telescope at Las Campanas observatory in Chile. In moderate seeing the survey limit was 21.2 magnitudes in the R-band. During bad seeing conditions the survey was not performed. 3) Kuiper Belt surveys are now nearly complete to about 21st magnitude in the R-band. The corresponding size of an object at 21st magnitude depends on the distance and albedo of the object. At 30 AU 21st magnitude corresponds to about H = 6.6 mags while at 50 AU H = 4.4 mags, which when assuming a moderate albedo of ρR = 0.15 correspond to radii of 80 km and 225 km respectively. Through looking at the cumulative luminosity function of the Kuiper Belt objects, significant incompleteness in the main Kuiper Belt probably starts around a radius of 100 km (H ∼ 6 mags) and becomes drastic around a radius of 60 km (H ∼ 7 mags.). 4) For the largest objects (H . 4.5 mag), the ratio of the population sizes for the various dynamical reservoirs in the outer solar system were found to be (1) : (2.6) : (7±3) : (75±+115 −55 ), for the (Plutinos):(Main Kuiper Belt):(Scattered Disk):(Sedna Types), respectively. Thus the scattered disk population is likely a few times larger than the main Kuiper Belt population and several times larger than the Plutino population. The Sedna type population likely is the biggest of all the observed outer solar system reservoirs but remains largely unknown because of the strong observational bias against finding very distant objects. 5) Beyond the Kuiper Belt edge, at a few hundred AU or so, there could easily be more Pluto, Mercury or even larger sized objects in Sedna-like orbits. No new Sedna-like objects were detected even though the survey was sensitive to objects up to about 300 AU. Sedna is likely one of the larger and thus one of the brighter members of its population. Any – 12 – further Sedna-like object detections will likely require significantly fainter magnitudes while still covering large areas of sky. Pan-Starrs has a chance to detect some Sedna like objects since it will survey large areas of sky to around a magnitude fainter than this survey, but LSST will be needed to find significant numbers of Sedna like objects since sensitivity and large areas of sky are needed to probe this distant, faint population. 6) All the major populated dynamical reservoirs in the Kuiper Belt, including the scat- tered disk, high inclination classical belt, low inclination classical belt (Quaoar), Sedna and the Plutinos are occupied by dwarf planet sized objects. Only the well-populated outer Neptune mean motion resonances such as the 2:1, 7:4, 5:2, and 5:3 are not occupied by a dwarf planet sized object. Any dwarf planet in these outer resonances would likely have been found to date, suggesting the outer resonances are occupied by a different mix of objects than the 3:2 resonance population or are significantly depleted in objects relative to the 3:2 resonance. 7) The scattered disk and main Kuiper Belt were found to have a power-law size dis- tribution of q = 3.3 ± 0.7 for the largest few objects (H < 4.5 mags), while the Plutino population has a shallower slope of q = 2.2 ± 0.5. The high and low inclination main Kuiper Belt populations appear to have similar slopes in their size distributions. 8) The main Kuiper Belt could have three distinct dynamical classes: (1) low inclination with low eccentricity (e < 0.05), (2) high inclination with low eccentricity (e < 0.07), and (3) high inclination with higher eccentricities (e > 0.13). Acknowledgments The OGLE project has received funding from the European Research Council under the European Community's Seventh Framework Programme (FP7/2007-2013) / ERC grant agreement no. 246678 to AU. C.T. was supported by the Gemini Observatory, which is operated by the Association of Universities for Research in Astronomy, Inc., on behalf of the international Gemini partnership of Argentina, Australia, Brazil, Canada, Chile, the United Kingdom, and the United States of America. REFERENCES Alard, C. 2000, A&AS, 144, 363. Alard, C. and Lupton, R. 1998, ApJ, 503, 325. – 13 – Allen, L., Bernstein, G. and Malhotra, R. 2001, ApJ, 549, L241. Barucci, M., Brown, M., Emery, J. and Merlin, F. 2008, in The Solar System Beyond Nep- tune, ed. M. Barucci, H. Boehnhardt, D. Cruikshank and A. Morbidelli (Tucson: Univ of Arizona Press), 143-160. Brown, M. 2001, AJ, 121, 2804. Brown, M. and Trujillo, C. 2004, AJ, 127, 2413. Brown, M., Trujillo, C. and Rabinowitz, D. 2004, ApJ, 617, 645. Brown, M., Trujillo, C. and Rabinowitz, D. 2005, ApJ, 635, L97. Brown, M. Schaller, E., Roe, H., Rabinowitz, D. and Trujillo, C. 2006, ApJ, 643, L61. Brown, M. 2008, in The Solar System Beyond Neptune, ed. M. Barucci, H. Boehnhardt, D. Cruikshank and A. Morbidelli (Tucson: Univ of Arizona Press), 335-344. Chiang, E. and Jordan, A. 2002, AJ, 124, 3430. Chiang, E., Jordan, A., Millis, R. et al. 2003, AJ, 126, 430. Cuzzi, J., Hogan, R., and Bottke, W. 2010, Icarus, 208, 518. Desch, S., Cook, J., Doggett, T. and Porter, S. 2009, Icarus, 202, 694. Doressoundiram, A., Boehnhardt, H., Tegler, S. and Trujillo, C. 2008, in The Solar System Beyond Neptune, ed. M. Barucci, H. Boehnhardt, D. Cruikshank and A. Morbidelli (Tucson: Univ of Arizona Press), 91-104. Dumas et al. 2007, AA, 471, 331. Elliot, J. and Kern, S. 2003, EM&P, 92, 375. Elliot, J., Kern, S., Clancy, K. et al. 2005, AJ, 129, 1117 Elliot et al. 2010, Nature, 465, 897. Fraser, W. et al. 2008, Icarus, 195, 827. Fraser, W. & Kavelaars, J. 2009, AJ, 137, 72. Fraser, W., Brown, M. & Schwamb, M. 2010, 210, 944. Fuentes, C. & Holman, M. 2008, AJ, 136, 83 Fuentes, C., George, M., & Holman, M. 2009, ApJ, 696, 91 Gladman, B. and Chan, C. 2006, ApJ Lett., 643, L135. Gladman, B. Marsden, B. and VanLaerhoven, C. 2008, in The Solar System Beyond Neptune, ed. M. Barucci, H. Boehnhardt, D. Cruikshank and A. Morbidelli (Tucson: Univ of Arizona Press), 43-57. – 14 – Gomes, R. 2003, Earth Moon Planets, 92, 29. Gomes, R., Levison, H., Tsiganis, K. and Morbidelli, A. 2005, Nature, 435, 466. Gomes, R., Fernandez, J., Gallardo, T. and Brunini, A. 2008, in The Solar System Beyond Neptune, ed. M. Barucci, H. Boehnhardt, D. Cruikshank and A. Morbidelli (Tucson: Univ of Arizona Press), 259-273. Hahn, J. and Malhotra, R. 2005, AJ, 130, 2392. Hartigan, J. and Hartigan, P. 1985, The Annals of Statstics, 13, 70. Iorio, L. 2009, MNRAS, 400, 346. Jewitt, D., Luu, J., Trujillo, C. 1998, AJ, 115, 2125. Jewitt, D. and Sheppard, S. 2002, AJ, 123, 2110. Jewitt, D. and Luu, J. 2004, Nature, 432, 731. Kenyon, S. and Luu, J. 1999, AJ, 118, 1101. Kenyon, S. and Bromley, B. 2008, ApJ Supplement, 179, 451. Kenyon, S., Bromely, B., O'Brien, D. and Davis, D. 2008, in The Solar System Beyond Neptune, ed. M. Barucci, H. Boehnhardt, D. Cruikshank and A. Morbidelli (Tucson: Univ of Arizona Press), 293-313. Kenyon, S and Bromley, B. 2010, ApJ Supplement, 188, 242. Levison, H. and Stern, S. A. 2001, AJ, 121, 1730. Levison, H., Morbidelli, A., Vanlaerhoven, C., Gomes, R., and Tsiganis, K. 2008, Icarus, 196, 258. Licandro, J. et al. 2006, AA, 445, 35. Lineweaver, C. and Norman, M. 2010, arXiv:1004.1091 Luu, J. and Jewitt, D. 1988, AJ, 95, 1256. Lykawka, P. S., and Mukai, T. Planetary and Space Science, 53, 1319. Malhotra, R. 1995, AJ, 110, 420. Melott, A. and Bambach, R. 2010, MNRAS, 407, L99. Morbidelli, A.and Levison, H. 2004, AJ, 128, 2564. Morbidelli, A., Levison, H. and Gomes, R. 2008, in The Solar System Beyond Neptune, ed. M. Barucci, H. Boehnhardt, D. Cruikshank and A. Morbidelli (Tucson: Univ of Arizona Press), 275-292. – 15 – Noll, K., Grundy, W., Chiang, E., Margot, J. and Kern, S. 2008, in The Solar System Beyond Neptune, ed. M. Barucci, H. Boehnhardt, D. Cruikshank and A. Morbidelli (Tucson: Univ of Arizona Press), 345-363. Peixinho, N., Lacerda, P., and Jewitt, D. 2008, AJ, 136, 1837-1845. Petit, J., Kavelaars, J., Gladman, B. and Loredo, T. 2008, in The Solar System Beyond Neptune, ed. M. Barucci, H. Boehnhardt, D. Cruikshank and A. Morbidelli (Tucson: Univ of Arizona Press), 71-87. Rabinowitz, D., Barkume, K., Brown, M., Roe, H., Schwartz, M., Tourtellotte, S. and Trujillo, C. 2006, ApJ, 639, 1238. Rabinowitz, D., Schaefer, B. and Tourtellotte, S. 2007, AJ, 133, 26. Rabinowitz, D. 2010, Poster presentation at the TNO 2010: Dynamical and Physical Prop- erties of Trans-Neptunian Objects workshop in Philadelphi, PA, USA Ragozzine, D. and Brown, M. 2007, AJ, 134, 2160. Schaller, E. and Brown, M. 2007, ApJ, 659, L61. Schlichting, H. and Sari, R. 2011, ApJ, 728, 68. Schwamb, M., Brown, M., and Rabinowitz, D. 2009, ApJ, 694, L45. Schwamb, M., Brown, M., Rabinowitz, D. and Ragozzine, D. 2010, ApJ, 720, 1691. Sheppard, S. 2007, AJ, 134, 787. Sheppard, S. and Trujillo, C. 2010, ApJ, 723, 233. Stansberry, J., Grundy, W., Brown, M., Cruikshank, D., Spencer, J., Trilling, D. and Margot, J. 2008, in The Solar System Beyond Neptune, ed. M. Barucci, H. Boehnhardt, D. Cruikshank and A. Morbidelli (Tucson: Univ of Arizona Press), 161-179. Stern, S. A. 2002, AJ, 124, 2297. Stewart, S. and Leinhardt, Z. 2009, ApJ Letters, 691, L133. Tegler, S. and Romanishin, W. 2000, Nature, 407, 979-981. Thomas-Osip, J., McCarthy, P., Prieto, G., Phillips, M., and Johns, M. 2011, arxiv:1101.2340 Trujillo, C. and Jewitt, D. 1998, AJ, 115, 1680. Trujillo, C., Jewitt, D. and Luu, J. 2000, ApJ, 529, L103. Trujillo, C., Jewitt, D. and Luu, J. 2001, AJ, 122, 457. Trujillo, C. and Brown, M. 2001, ApJ, 554, L95. Trujillo, C. and Brown, M. 2002, ApJ, 566, L125. – 16 – Trujillo, C. and Brown, M. 2003, EM&P, 92, 99. Trujillo, C., Sheppard, S. and Schaller, E. 2011, ApJ, 730, 105 Udalski, A., Szymanski, M., Stanek, K. et al. 1994, Acta Astronomica, 44, 165. Udalski, A., Kubiak, M., and Szymanski, M. 1997, Acta Astronomica, 47, 319. Udalski, A. 2003, Acta Astronomica, 53, 291. Wozniak, P. 2000, Acta Astronomica, 50, 421. Wozniak, P., Udalski, A., Szymanski, M., Kubiak, M., Pietrzynski, G., Soszynski, I. and Zebrun, K. 2001, Acta Astronomica, 51, 175. This preprint was prepared with the AAS LATEX macros v5.2. – 17 – Table 1. New outer Solar System objects discovered in this survey Name H mR a e i R r (mag) (mag) (AU) (deg) (AU) (km) 2010 EK139 2010 KZ39 2010 FX86 2010 EL139 2010 HE79 2010 PU75 2010 JK124 2009 MF10 2010 HD112 2010 JJ124 2009 MG10 2010 HG109 2010 HU113 2009 ME10 3.8 3.9 4.3 5.0 5.1 5.3 5.4 6.0 6.5 6.6 7.0 7.3 7.4 7.5 19.5 20.1 20.7 20.1 19.8 20.9 21.2 21.1 22.2 20.1 21.7 21.7 22.1 21.0 69.1 45.8 47.0 39.2 39.3 43.4 39.7 57.5 44.5 83.0 47.5 39.8 36.2 27.8 0.53 0.15 0.08 0.07 0.20 0.08 0.09 0.52 0.03 0.72 0.34 0.23 0.03 0.18 29.5 26.1 25.2 23.0 15.7 10.2 15.6 26.1 3.9 37.8 19.9 29.2 11.3 14.7 40.5 46.3 46.8 36.6 34.9 40.0 40.3 36.1 43.1 24.1 32.8 30.5 35.3 23.1 310a 300a 230a 190 180 150 140 120 100 80 70 60 60 50 Note. - Orbital elements are from the Minor Planet Center and are the semimajor axis (a), inclination (i), and eccentricity (e). The radii (r) of the new objects were determined assuming an albedo of 0.15 and using the equation, r = (2.25 × 1016R2∆2/pRφ(0))1/2100.2(m⊙−mR) where R is the heliocentric distance in AU, ∆ is the geocentric distance in AU, m⊙ is the apparent red magnitude of the sun (−27.1), pR is the red geometric albedo, mR is the apparent red magnitude of the object and φ(0) = 1 is the phase function at opposition. H is the absolute magnitude of the object. aThese objects could be labeled as dwarf planets since their radii are larger than 200 km assuming a moderate or lower albedo. – 18 – Table 2. Known KBOs and Centaurs detected in this survey Name H mR a e i R r (mag) (mag) (AU) (deg) (AU) (km) (134340) Pluto 2007 JJ43 (10199) Chariklo (55576) Amycus -0.7 3.2 6.4 7.8 13.6 19.4 17.5 19.6 39.6 48.0 15.8 25.0 0.25 0.16 0.17 0.39 17.1 12.1 23.4 13.3 31.8 41.7 13.8 16.8 1150 350 100 50 Note. - See Table 1 for comments and definitions. Table 3. Ten Intrinsically Brightest TNOs Name H a e i Class (mag) (AU) (deg) (136199) Eris (134340) Pluto (136472) Makemake (136108) Haumea (90377) Sedna (225088) 2007 OR10 (90482) Orcus (50000) Quaoar (28978) Ixion -1.2 -0.7 -0.3 0.2 1.6 1.9 2.3 2.5 3.2 68.0 39.7 45.4 43.0 510 67.3 39.2 43.5 39.6 0.43 0.25 0.16 0.20 0.85 0.50 0.23 0.04 0.25 Scattered 43.9 3:2 Resonance 17.1 29.0 High i Classical 28.2 High i Classical 11.9 30.7 20.6 8.0 19.6 3:2 Resonance Low i Classical 3:2 Resonance Sedna Scattered Note. - The orbital elements are from the Minor Planet Center and are the semimajor axis (a), inclination (i), and eccentricity (e). H is the absolute magnitude and Class is the dynamical classification of the object. – 19 – Table 4. Bright Kuiper Belt Population Statistics Class Hcomp (mag) rcomp (km) N Pop Ratio (N/N3:2) 3:2 5:3 7:4 2:1 5:2 Main Kuiper Belt (MKB) Scattered Disk Sedna Type MKB Low i & e MKB High i, All e MKB High i & e MKB High i & Low e 6 0 0 0 0 13a 1 0/5 0/5 0/3 0/2 13/5a 4.5 4.2 4.1 3.3 2.5 4.1 4.1b N/A 35 ± 15b 35 ± 15/5 1.6c N/A 75+115 75+115 −55 /1 −55 4.6 3/6 3 11/5 11 4.1 8/5 8 4.1 4.6 6 6/6 210 250 260 380 550 260 200 260 260 200 c Note. - Hcomp is the absolute magnitude completion limit for the particular dynamical class while rcomp is the radius completion limit assuming an albedo of 0.15. N is the number of objects known within each class with an aboslute magnitude equal to or brighter than Hcomp. The Pop Ratio is the population number ratio of each dynamical class relative to the 3:2 resonance number population at the Hcomp of that particular dynamical class (i.e. N/N3:2). aNone of the Haumea family members, except for Haumea itself, are in- cluded. The Haumea family members are likely pieces of Haumea and have very high albedos unlike most of the other moderately sized objects with ab- solute magnitudes around 3 or 4 (see Raggozine et al. 2007; Trujillo et al. 2011). bSince the scattered disk objects spend most of their time near aphelion, which can be up to a few hundred AU, the absolute magnitude completion number here is for objects currently within about 50 AU of the Sun. The to- tal number of possible scattered disk objects with absolute magnitdue brighter than this was determined by taking the number of known objects of this bright- – 20 – ness or brighter and a Poisson probability statistic of how many more are cur- rently unobservable in the distant solar system based on the percent of time the known objects would be brighter than 21st magnitude in their orbit. cike the scattered disk objects, Sedna is only brighter than 21st magnitude near perihelion. Thus for most of Sedna's orbit it would not be detected by the current large area surveys. To account for this, a Poisson probability statistic of how many more Sedna type objects of similar size are unobservable in the distant solar system was determined based on Sedna's orbit. – 21 – Fig. 1.- The black shaded regions represent the sky area surveyed in this work. The horizontal axis is the right ascension in hours and the vertical axis is the declination in degrees. The blue solid line shows the ecliptic, the purple dashed line shows −20 degrees from the ecliptic and the cyan dotted dashed line shows −30 degrees from the ecliptic. Areas more than −40 degrees south of the ecliptic are shown with brown horizontal stripes. The area within 15 degrees of the galactic plane is shown with red crossed stripes while the area covered by wide-field KBO surveys from the north (Trujillo and Brown 2003; Brown 2008; Schwamb et al. 2009, 2010) are shown with green angled stripes. Almost all KBOs are expected to be within 20 degrees of the ecliptic with it highly unlikely any KBO is beyond 40 degrees from the ecliptic (Brown 2008). – 22 – Fig. 2.- Detection efficiency of the KBO survey versus the apparent red magnitude using the Warsaw 1.3 meter telescope. In good seeing (0.8 arcseconds FWHM) the 50% detection efficiency is at about 21.6 mags while in moderate seeing (∼ 1 arcsecond) it is about 21.2 mags in the R-band. Effective radii of the apparent magnitudes were calculated assuming the object has an albedo of 0.15 and is at 40 AU. – 23 – Fig. 3.- A small portion of an image showing Pluto in the galactic plane from the 1.3 meter Warsaw telescope. Pluto is in the center of this image as revealed in Figure 4. – 24 – Fig. 4.- A difference image of the Pluto fields (Figure 3) showing the removal of the steady state background of stars. The motion of Pluto is clearly revealed in the difference image as a positive (bright) and negative (dark) point from the subtraction process of the two individual images. – 25 – Fig. 5.- The outer solar system has now been surveyed to a completeness limit of about 21 magnitudes in the R-band. This figure shows what size and distance an object would be for several different albedos for a 21st magnitude object. Shaded areas correspond to completeness limits for albedo < 0.1 (light), albedo < 0.25 (medium) and albedo < 0.625 (dark). It is clear that a Pluto (1161 km) or even larger sized object could easily have gone undetected to date if beyond a few hundred. – 26 – Fig. 6.- The radius of an object is shown assuming a moderate albedo of 0.15 for various heliocentric distances and apparent red magnitudes. The known objects in the Kuiper Belt region are complete to about 21st magnitude, shown by the shaded region below the dashed line. It is likely that everything under the dashed line at 21 magnitudes is known. The radii used for the named objects in the figure are Pluto (1161 km), Mercury (2440 km), Mars (3396 km), Earth (6371 km) and an arbitrary lower limit on a hypothetical eccentric giant planet or companion to our Sun, sometimes called Nemesis or Tyche, with 10,000 km radius (Iorio 2009; Melott and Bambach 2010). In the distant solar system very large objects would easily be undetected to date. – 27 – Fig. 7.- The heliocentric distance versus the completeness of absolute magnitude, H. The shaded region shows where the outer solar system should be complete in discoveries. The effective radius on the right side assumes an albedo of 0.15. The average semi-major axis for the various major Neptune resonance populations are shown as vertical dashed lines for reference. – 28 – Fig. 8.- The absolute magnitude versus the cumulative number of all known trans- Neptunian objects (solid line). The absolute magnitudes for the largest objects appear overly bright since these objects have much higher albedos than most smaller KBOs. Squares show the absolute magnitudes that the largest KBOs would have if their albedos were 0.15 and not around 0.7 as has been found for Eris, Pluto, Makemake and Haumea. Squares also show the absolute magnitudes the moderately sized KBOs would have if their albedos were not around 0.25 but 0.15 for Sedna, 2007 OR10, Orcus, and Quaoar (Stansberry et al. 2008). – 29 – Fig. 9.- The absolute magnitude versus the differential number of known trans-Neptunian objects. Objects are binned in 1 magnitude bins. The largest few objects have had their absolute magnitudes adjusted fainter as in Figure 8 to account for their higher albedos compared to the smaller objects. The dashed line shows the best fit to the largest objects. It is apparent that the Kuiper Belt is nearly complete to about an absolute magnitude of around 5-6 mags after which a turnover shows significant incompleteness. – 30 – Fig. 10.- The semi-major axis versus eccentricity of multi-opposition trans-Neptunian objects. Large circles represent TNOs with H ≤ 4.5 mags. This figure shows several distinct dynamical KBO populations. Vertical dashed lines show the main resonances with Neptune as well as the Neptune Trojans in the 1 : 1 resonance. Scattered disk objects have perihelia 30 . peri . 45 AU as shown between the dashed lines. Classical objects are in the lower center portion of the figure and include the the Main Kuiper Belt (MKB) with its high and low inclination populations. There also appears to be a high and low eccentricity population of large objects. An edge near 50 AU can clearly be seen for low eccentricity objects. Centaurs are on unstable orbits between the giant planets. Sedna stands out as being significantly below the perihelion line shown at 40 AU demonstrating its decoupled influence from Neptune unlike the scattered disk objects. – 31 – Fig. 11.- The absolute magnitude versus the cumulative number of objects for the various completeness limits (see Table 4) of the dynamical classes in the Kuiper Belt. The largest few objects have had their absolute magnitudes adjusted as in Figure 8 to account for their higher albedos compared to the smaller objects. The Sedna type (square) and scattered disk objects (dotted line) have significant error bars as these populations are not complete for even the largest objects. Poisson statistics were used to extrapolate the total scattered disk and Sedna populations from the known objects brighter than 21st magnitude using the amount of time the objects would be detectable in their eccentric orbits (Table 4). The scattered population is likely larger than either the main Kuiper Belt (solid line) or the 3:2 resonance population (dashed line). The Sedna type population appears to be the largest of all the populations by a factor of ten or more. The 3:2 resonance population has a shallower size distribution slope (q = 2.2 ± 0.5) than the other populations (q = 3.3 ± 0.7). – 32 – Fig. 12.- The absolute magnitude versus the cumulative number of objects in the main Kuiper Belt (solid line) and its subcategories. The largest few objects have had their absolute magnitudes adjusted as in Figure 8 to account for their higher albedos compared to the smaller objects. The high inclination (i > 10 degrees) and eccentricity (e > 0.13) objects (dotted dashed line) dominate the main Kuiper Belt on the large end. There appears to be a sizable group of main Kuiper Belt objects that have high inclinations but low eccentricities (e < 0.07) (triple dotted dashed line). These high i and low e objects could be related to either of the other subcategories in this figure. Large low inclination and low eccentricity objects (long dashed line) are very rare with only Quaoar being brighter than an absolute magnitude of 4.2.
1510.07689
1
1510
2015-10-26T21:19:46
Hydrogen-Water Mixtures in Giant Planet Interiors Studied with Ab Initio Simulations
[ "astro-ph.EP" ]
We study water-hydrogen mixtures under planetary interior conditions using ab initio molecular dynamics simulations. We determine the thermodynamic properties of various water-hydrogen mixing ratios at temperatures of 2000 and 6000 K for pressures of a few tens of GPa. These conditions are relevant for ice giant planets and for the outer envelope of the gas giants. We find that at 2000 K the mixture is in a molecular regime, while at 6000 K the dissociation of hydrogen and water is important and affects the thermodynamic properties. We study the structure of the liquid and analyze the radial distribution function. We provide estimates for the transport properties, diffusion and viscosity, based on autocorrelation functions. We obtained viscosity estimates of the order of a few tenths of mPa.s for the conditions under consideration. These results are relevant for dynamo simulations of ice giant planets.
astro-ph.EP
astro-ph
Hydrogen-Water Mixtures in Giant Planet Interiors Studied with Ab Initio Simulations F. Soubiran Department of Earth and Planetary Science, University of California, Berkeley, CA 94720, U.S.A. B. Militzer Department of Earth and Planetary Science, Department of Astronomy, University of California, Berkeley, CA 94720, U.S.A. 5 1 0 2 t c O 6 2 . ] P E h p - o r t s a [ 1 v 9 8 6 7 0 . 0 1 5 1 : v i X r a Abstract We study water-hydrogen mixtures under planetary interior conditions using ab initio molecular dynamics simula- tions. We determine the thermodynamic properties of various water-hydrogen mixing ratios at temperatures of 2000 and 6000 K for pressures of a few tens of GPa. These conditions are relevant for ice giant planets and for the outer envelope of the gas giants. We find that at 2000 K the mixture is in a molecular regime, while at 6000 K the dissociation of hydrogen and water is important and affects the thermodynamic properties. We study the structure of the liquid and analyze the radial distribution function. We provide estimates for the transport properties, diffusion and viscosity, based on autocorrelation functions. We obtained viscosity estimates of the order of a few tenths of mPa s for the conditions under consideration. These results are relevant for dynamo simulations of ice giant planets. Keywords: hydrogen, water, planetary interior, trans- port properties. 1. Introduction The extraordinary discovery of more than one thou- sand confirmed exoplanets in the past decade [1] under- lines the necessity for a better description of their struc- ture, formation, and evolution. The discovery of an un- expectedly rich population of Neptune and Sub-Neptune exoplanets [2] stressed the need for a better understand- ing of these kinds of planets. Based on the gravitational moments measured by Voyager II and orbital observations of their satellites [3, 4], different models have been pro- posed for the two ice giants of our solar system, Uranus and Neptune [5, 6]. Still, considerable uncertainties about their internal composition remain, and a more accurate equations of state (EOS) is needed to improve the models. Furthermore, magnetic field observations provide ad- ditional information and add some constraints on interior models. Dynamo simulations [7, 8, 9] predicted that wa- ter in different phases is important for the magnetic field generation. So far, such dynamo simulations rely on ice giant interior models that assume water to be fully phase separated from other compounds including hydrogen. Re- cent ab initio computer simulations predicted that water and metallic hydrogen readily mix at pressure-temperature conditions in the cores of giant planets. Experimental data Email address: [email protected] (F. Soubiran) obtained at lower pressures [10] predicted the true picture to be more complex and suggested partial mixing of wa- ter and hydrogen. These results may also become relevant for the interiors of gaseous giant planets like Saturn and Jupiter because the presence of a small amount of water would affect the physical properties of their envelopes. The properties of hydrogen-water mixtures in different concen- trations are thus of significant interest in planetary science. Here we performed computer simulations of fluid water- hydrogen mixtures at different concentrations under the pressure-temperature conditions found in Neptune's and Uranus' upper mantle and the outer envelope of Jupiter and Saturn. We explored the effects of the concentration, temperature, and pressure on both the thermodynamic properties and the structure of the fluid. The most sig- nificant changes are introduced when the hydrogen and water dissociate at high temperature and pressure. We also computed the transport properties, such as particle diffusion and viscosity in order to better constrain the dy- namo simulations. 2. Simulation method We performed molecular dynamics simulations based on finite temperature density functional theory (DFT) us- ing the Vienna Ab initio Simulation Package (VASP) [11]. We studied four different mixtures from H2O:H2=16:64, Preprint submitted to Elsevier July 22, 2021 24:48, 32:32, and 40:16 molecules in simulation cell. We define the water concentration as x = NH2O/(NH2O + NH2 ). Simulation results for pure water and hydrogen can be found in Refs. [12, 13, 14] and [15, 16, 17, 18, 19]. Our simulations were at least 3 ps long for a time-step of 0.2 fs, short enough to accurately describe the H-H and O-H bonds vibrations. The temperature was kept constant using a Nos´e thermostat [20, 21]. The electronic structure was computed at each time step using Mermin's finite temperature approach [22] of the Kohn-Sham scheme [23], with a Fermi-Dirac occu- pation distribution. We used projector augmented wave (PAW) pseudopotentials [24] with a cut-off radius of rcut = 0.8 a0 for hydrogen and rcut = 1.1 a0 for oxygen. For the plane-wave basis, an energy cut-off of 1100 eV was needed to reach a convergence with less than 1% inaccuracy on both the energy and the pressure. We used the generalized gradient approximation (GGA) with the Perdew, Burke, and Ernzerhof (PBE) [25] exchange-correlation functional that gave reasonable results in similar systems [26, 12]. All simulations were performed using periodic boundary con- ditions. We chose to use the Baldereschi k-point [27] for sampling the Brillouin zone because it gave similar ther- modynamic functions (within 1%) to a 2×2×2 Monkhorst- Pack grid [28]. In this article we will present results at two temper- atures: 2000 and 6000 K with pressures ranging from 10−35 GPa and 10−70 GPa for the lower and higher tem- peratures, respectively. These conditions are close to what we find in Neptune's or Uranus' upper mantle [29, 30, 6] or in the outer envelop of Jupiter and Saturn [31, 32]. 3. Results and discussion 3.1. Thermodynamic properties We extracted the pressure and the internal energy time- averages from the different molecular dynamics simula- tions. The error bars presented in the figures were com- puted using the block averaging method [33]. In Fig. 1, we report the evolution of the molecular density as a function of the pressure at 2000 K. The trend is quite similar for each mixing ratio. Molecular densities are much higher for hydrogen-rich mixtures, which simply reflects the fact the hydrogen molecules are smaller than the water molecules. In Fig. 2, we plotted the internal energy per molecule as function of pressure for two temperatures. The curves for each concentration have been shifted in order to superpose them for each isotherm and compare their evolution as a function of the pressure. At 2000 K, the linear behavior of the energy is very similar for all concentrations, indicating that the thermal excitations of H2 and H2O molecules make similar con- tributions to both the internal energy and the pressure. In contrast, we find significant deviations from the lin- ear relationship at 6000 K. More importantly, there is a 3 l ] Å / s e u c e o m l [ y t i s n e d r a u c e o M l l 0.11 0.10 0.09 0.08 0.07 0.06 0.05 0.04 5 0.20 0.33 0.50 0.71 10 15 20 25 30 35 40 Pressure [GPa] Figure 1: (color online) Molecular density as a function of the pres- sure at 2000 K. The four colors are for the four different H2O con- centrations, x, given in the caption. dependence on the concentration, indicating a change of structure of the liquid, which we discuss in section 3.2. 0.20 0.33 0.50 0.71 1.2 1.0 0.8 0.6 0.4 0.2 0.0 l l / ] e u c e o m V e [ y g r e n E −0.2 0 10 20 30 40 50 60 70 80 Pressure [GPa] Figure 2: (color online) Internal energy per molecule as a function of the pressure at 2000 K (•) and 6000 K ((cid:4)). The energies have been manually shifted for each concentration and temperature to better compare them as function of pressure. At 2000 K, the energy origins for each concentration have been chosen so that at 8.6 GPa, the energy is 0.2 eV/mol. For the 6000 K isotherm, the energy at 13.5 GPa is 0.0 eV/mol. The four colors are for the four different H2O concentrations, x, given in the caption. 3.2. Structure and composition The molecular dynamics simulations provide informa- tion about the microscopic structure of the fluid. Ra- dial distribution functions (RDF) are an established tool [34, 35] to analyze the structure. A first example is shown in Fig. 3, where we plotted the RDF of an equimolar mix- ture, x = 0.5, at three different pressures and a 2000 K 2 ) r ( g 12 10 8 6 4 2 0 0 7 GPa 14 GPa 34 GPa 2 4 6 8 10 r [bohr] ) r ( g 5 4 3 2 1 0 0 12 GPa 21 GPa 72 GPa 2 4 6 8 10 r [bohr] Figure 3: (color online) Radial distribution function of a x = 0.5 mixture at 2000 K and different pressures. The O-O structure is plotted in red, O-H in blue and H-H in cyan. Figure 4: (color online) Radial distribution function of a x = 0.5 mixture at 6000 K and different pressures. The O-O structure is plotted in red, O-H in blue and H-H in cyan. temperature. The large peaks in the O-H and H-H cor- relations underline the molecular character of the fluid, which is fully molecular up to at least 35 GPa. There ap- pear to be no other strong correlations in the fluid besides the molecular structure except at highest pressure where the correlation among the oxygen nuclei increases. Under these conditions, the fluid is close to phase transition to a superionic regime at 40-50 GPa at 2000 K in pure water [12]. In Fig. 4, we plotted the RDF for a x = 0.5 mixture at 6000 K. The size of the H-H and H-O peaks is significantly reduced compared to 2000 K indicating a partial dissocia- tion of the molecules. This dissociation is consistent with the behavior observed in pure hydrogen [26] and water [12]. We infer that the dissociation is the cause of the concentration dependence of the internal energy-pressure relationship in Fig. 2. We did not identify any other corre- lations besides the molecular peaks but it should be noted that the O-O correlation increases at 70 GPa. 3.3. Transport properties 3.3.1. Autocorrelation functions We derived the viscosity and diffusion coefficients of the hydrogen and oxygen species in each mixture in order to provide some dynamic information about the mixture and ultimately to improve models for planetary interiors. We used the autocorrelation functions for the diffusion and viscosity calculations. The diffusion coefficient of species is given by the Green-Kubo formula [36, 34]: Dα = 1 3Nα Xα Z +∞ 0 hvα(τ ) · vα(0)i dτ, (1) where we sum over all Nα nuclei of type α. vα(τ ) is the particle's velocity vector at time τ . The brackets mean that we average over many possible time origins assuming the fluid has reached an equilibrium in our simulations[37]. To reduce the fluctuations further, we averaged the result from different maximum integration times ranging from 75 to 100% of our time-window. The error bars in Figures 5-6 and Tables 1-2 represent one standard deviation from the mean. We computed the viscosity, η, from the autocorrelation function of the stress-tensor σαβ , η = V 3kBT X{αβ} Z +∞ 0 hσαβ (τ )σαβ (0)i dτ, (2) where the sum {αβ} includes all the deviatoric stress ten- sor components, {xy, xz, yz}. V is the volume of the sim- ulation cell and T the temperature of the system. There are still strong fluctuations in the stress tensor autocorre- lation function even when we use multiple origins because the stress-tensor is a bulk property and it is not possible to average over the multiple particles. In order to reduce the spreads in the results we fitted the autocorrelation function by a simple exponential function. This method provides more stable estimates of the viscosity. Nevertheless, we do not provide error bars in the figures because our viscosity values have to be considered to be order-of-magnitude esti- mates only because it is too difficult to obtain an accurate determination of the viscosity and its error bar. 3.3.2. Diffusion We analyze the diffusion of individual nuclei rather than tracking the motion of molecules because the lat- ter method becomes difficult as soon as dissociation and recombination effects occur at elevated temperature. In Fig. 5, we plotted the diffusion coefficient of oxy- gen nuclei as function of pressure for 2000 and 6000 K. For both temperatures, we find that diffusion slows down with increasing pressure because the particle interact more 3 0.0030 ] s / 2 m c [ O f o t n e i c i f f e o c n o i s u f f i D 0.0025 0.0020 0.0015 0.0010 0.0005 0.20 0.33 0.50 0.71 0.006 0.005 0.004 0.003 0.002 0.001 ] s / 2 m c [ H f o t n e i c i f f e o c n o i s u f f i D 0.20 0.33 0.50 0.71 0.0000 0 10 20 30 40 50 60 70 80 0.000 0 10 20 30 40 50 60 70 80 Pressure [GPa] Pressure [GPa] Figure 5: (color online) Oxygen diffusion coefficient as a function of the pressure at 2000 K (•) and 6000 K ((cid:4)).The four colors are for the four different H2O concentrations, x, given in the caption. Figure 6: (color online) Hydrogen diffusion coefficient as a function of the pressure at 2000 K (•) and 6000 K ((cid:4)). The four colors are for the four different H2O concentrations, x, given in the caption. strongly. At 6000 K, the diffusion is faster because the additional kinetic energy allows particles to hop over po- tential barriers more rapidly. The oxygen diffusion rate decreases with increasing water concentration presumably because water molecules interact more strongly with each other than they do with hydrogen molecules at the same pressure. The diffusion coefficient of the hydrogen nuclei are plot- ted in Fig. 6. For the dissociated fluids at 6000 K, the hydrogen atoms diffuse approximately twice as fast as the oxygen atoms that are heavier and larger. In the molecular regime, the hydrogen and oxygen diffusion rates are more similar because some hydrogen nuclei are bound to oxygen. Again we find that when the water concentration rises, the diffusion slows down. There is at least two reasons for that. First, an increasing fraction of hydrogen nuclei is bound in water molecules, which greatly restricts their mobility. Second the mobility of the hydrogen molecules is reduced by the interaction with water molecules. It is interesting to note that the diffusion rates at 6000 K are almost independent of pressure. This is most likely due to two counter-acting effects. With increase pressure, particles interact more strongly, which reduces diffusion rates. A pressure increase also introduces a rise in the dis- sociation fraction, which increases the mobility of different nuclei. 3.3.3. Viscosity In Fig. 7, we plotted our viscosity estimates for the H2- H2O mixtures. Values vary from 0.1 to 1 mPa s over the range of parameters that we explored here. The viscosity rises with increasing water concentration because water molecules interact more strongly than hydrogen molecules at the same pressure. The viscosity rises with increasing pressure and decreasing temperature, as expected. Vis- cosity estimates are of importance for the numerical sim- ulations of planetary interiors, especially for Neptune-like planets where the water content is presumably high. More- over, the viscosity plays a prominent role in the dynamics of the dynamo [7, 8, 9] and reliable estimates are therefore needed. ] s . a P m [ y t i s o c s i V 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0.0 0 0.20 0.33 0.50 0.71 10 20 30 40 50 60 70 80 Pressure [GPa] Figure 7: (color online) Viscosity as a function of the pressure at 2000 K (•) and 6000 K ((cid:4)). The four colors are for the four different H2O concentrations, x, given in the caption. 4. Conclusion We performed ab initio computer simulations of H2- H2O mixtures at 2000 K and 6000 K in order to provide information about the physics of such systems under con- ditions in the interior of icy or gaseous giant planets. We showed for instance that at 2000 K and a few tens of GPa, the mixture is purely molecular and its thermodynamic behavior depends only weakly on the water concentration. 4 Table 1: Hydrogen and oxygen diffusion coefficients and pressure ta- ble for different molecular densities and water concentrations xH2O at 2000 K. Mol. dens. (×10−2A−3) H diffusion O diffusion (×10−3 cm2/s) (×10−3 cm2/s) Pressure (GPa) 8.534(36) 16.73(11) 24.776(45) 36.694(75) 7.731(46) 15.936(56) 23.339(90) 6.763(30) 14.139(64) 21.66(13) 34.04(14) 5.667(51) 12.45(11) 19.40(14) 31.64(10) 1.197(84) 0.739(45) 0.573(10) 0.371(23) 1.099(80) 0.686(14) 0.459(16) 0.91(11) 0.461(19) 0.371(15) 0.281(12) 0.654(39) 0.357(38) 0.287(32) 0.205(23) 0.876(30) 0.441(80) 0.377(22) 0.219(59) 0.683(90) 0.427(52) 0.266(49) 0.622(85) 0.346(43) 0.252(7) 0.157(17) 0.560(14) 0.296(61) 0.060(18) 0.090(13) O diffusion xH2O 0.20 0.20 0.20 0.20 0.33 0.33 0.33 0.50 0.50 0.50 0.50 0.71 0.71 0.71 0.71 xH2O 0.20 0.20 0.20 0.20 0.20 0.33 0.33 0.33 0.33 0.33 0.50 0.50 0.50 0.50 0.50 0.71 0.71 0.71 0.71 0.71 6.0105 8.0000 9.3308 10.974 5.4095 7.2000 8.3977 4.8084 6.4000 7.4646 8.7792 4.2074 5.6000 6.5316 7.6818 6.0105 8.0000 9.3308 10.974 13.027 5.4095 7.2000 8.3977 9.8765 11.724 4.8084 6.4000 7.4646 8.7792 10.421 4.2074 5.6000 6.5316 7.6818 9.1187 Table 2: Same as Table 1 but at 6000 K. Pressure H diffusion Mol. dens. (×10−2A−3) (GPa) 13.47(11) 23.09(11) 31.80(12) 45.87(11) 68.514(96) 12.57(12) 22.26(11) 31.581(93) 45.95(16) 70.38(16) 11.570(81) 21.40(18) 30.83(15) 46.45(16) 72.25(16) 10.32(11) 20.24(17) 29.57(16) 45.78(19) 73.14(23) (×10−3 cm2/s) (×10−3 cm2/s) 5.47(23) 5.03(16) 4.729(83) 5.30(19) 4.86(17) 4.60(11) 4.236(63) 4.080(34) 4.011(75) 4.20(13) 4.02(15) 3.70(21) 3.54(40) 3.46(29) 3.35(28) 3.186(32) 2.92(25) 2.869(50) 2.888(67) 2.775(54) 2.54(12) 1.67(46) 1.834(72) 1.245(75) 1.190(64) 2.11(18) 1.39(13) 1.12(24) 1.01(13) 0.878(64) 1.95(26) 1.388(56) 1.05(11) 0.931(85) 0.616(51) 1.76(10) 1.132(48) 0.87(17) 0.568(72) 0.491(38) In contrast, at 6000 K we observe partial dissociation of both H2 and H2O molecules, which changes in the thermo- dynamic behavior. Because dissociation is often associated with a reduction in the electronic band gap, we can infer that the dissociated fluids [38] have a significantly higher electronic conductivity. This has potential implications for the magnetic field generation in water-rich gas giants planets. The autocorrelation function analysis provided robust results for the diffusion of oxygen and hydrogen nuclei in the mixtures. The diffusion of hydrogen is almost constant at 6000 K from 10 to 70 GPa range due to the dissociation process that add to the mobility of the protons. The dif- fusion coefficients are important for the planetary models because the diffusion influences the convective behavior in the semi-convective regime as shown by Leconte and Chabrier [31]. We also computed the viscosity which is of the order of a few tenths of mPa s at the conditions under consideration. This is not a very viscous fluid indicating a turbulent behavior in planetary interiors. 5 5. Acknoledgements This work has been supported by NSF and NASA. The simulations were performed on the NASA supercomputing facilities. References References [1] Exoplanet Archive, URL http://exoplanetarchive.ipac.caltech.edu , ???? [2] E. A. Petigura, G. W. Marcy, A. W. Howard, A Plateau in the Planet Population below Twice the Size of Earth, Astrophys. J. 770 69. [3] R. A. Jacobson, The gravity field of the Uranian system and the orbits of the Uranian satellites and rings, Bulletin of the American Astronomical Society 39 (2007) 453. [4] R. A. Jacobson, The orbits of the Neptunian satellites and the orientation of the pole of Neptune, Astrophys. J. 137 (2009) 4322. [5] R. Helled, J. D. Anderson, M. Podolak, G. Schubert, Interiors Models of Uranus and Neptune, Astrophys. J. 726 (2011) 15. [6] N. Nettelmann, R. Helled, J. Fortney, R. Redmer, New indi- cation for a dichotomy in the interior structure of Uranus and Neptune from the application of modified shape and rotation data, Planetary and Space Science 77 (0) (2013) 143 -- 151. [7] S. Stanley, J. Bloxham, Convective-region geometry as teh cause of Uranus' and Neptune's magnetic fields, Nature 428 (2004) 151. [8] S. Stanley, J. Bloxham, Numerical dynamo models of Uranus' and Neptune's unusual magnetic fields, Icarus 184 (2006) 556. [9] J. M. Bloxham, M. H. Heimpel, E. M. King, J. M. Aumou, Turfields models of ice giant internal dynamics: Dynamos, heat transfer, and zonal flows, Icarus 224 (2013) 97. [10] E. Bali, A. Aud´etat, H. Keppler, Waplanet hydrogen are im- miscible in Earth's mantle, Nature 495 (2013) 220. [11] Vienna Ab initio Simulation Package, URL http://www.vasp.at, ???? [12] M. French, T. R. Mattsson, N. Nettelmann, R. Redmer, Equa- tion of state and phase diagram of water at ultrahigh pressures as in planetary interiors, Phys. Rev. B 79 (2009) 054107, URL http://link.aps.org/doi/10.1103/PhysRevB.79.054107 . [13] H. F. Wilson, M. L. Wong, B. Militzer, Phys. Rev. Lett. 110 (2013) 151102. [14] B. Militzer, S. Zhang, "Ab initio Simulations of Superionic H2O, H2O2, and H9O4 Compounds", submitted to Phys. Rev. B (2014), ???? [15] B. Militzer, D. M. Ceperley, Phys. Rev. Lett. 85 (2000) 1890. [16] B. Militzer, D. M. Ceperley, Phys. Rev. E 63 (2001) 066404. [17] J. Vorberger, I. Tamblyn, B. Militzer, S. Bonev, Phys. Rev. B 75 (2007) 024206. [18] B. Militzer, Phys. Rev. B 87 (2013) 014202. [19] B. Militzer, W. B. Hubbard, Astrophys. J. 774 (2013) 148. [20] S. Nos´e, J. Chem. Phys 81 (1984) 511. [21] Nos´e, Prog. Theor. Phys. Suppl. 103 (1991) 1. [22] N. D. Mermin, Thermal Properties of the Inhomogeneous Elec- tron Gas, Phys. Rev. 137 (5A) (1965) A1441 -- A1443. [23] W. Kohn, L. J. Sham, Self-Consistent Equations Including Ex- change and Correlation Effects, Physical Review 140 (1965) 1133. [24] P. E. Blochl, Projector augmented-wave method, Phys. Rev. B 50 (24) (1994) 17953 -- 17979. [25] J. P. Perdew, K. Burke, M. Ernzerhof, Generalized Gradient Approximation Made Simple, Phys. Rev. Lett. 77 (18) (1996) 3865 -- 3868. [26] L. Caillabet, S. Mazevet, P. Loubeyre, Multiphase equation of state of hydrogen from ab initio calculations in the range 0.2 to 5 g/cc up to 10 eV, Phys. Rev. B 83 (2011) 094101, URL http://link.aps.org/doi/10.1103/PhysRevB.83.094101 . [27] A. Baldereschi, Mean-Value Point (1973) in Zone, 5212 -- 5215, http://link.aps.org/doi/10.1103/PhysRevB.7.5212. Phys. Rev. B 7 the Brillouin URL [28] H. J. Monkhorst, J. D. Pack, Special points for Brillouin-zone integrations, Phys. Rev. B 13 (12) (1976) 5188. [29] W. B. Hubbard, W. J. Nellis, A. C. Mitchell, N. C. Holmes, P. C. McCandless, S. S. Limaye, Interior structure of Neptune - Comparison with Uranus, Science 253 (1991) 648 -- 651. [30] M. Podolak, A. Weizman, M. Marley, Comparative models of Uranus and Neptune, Planet. Space Sci. 43 (1995) 1517 -- 1522. [31] J. Leconte, G. Chabrier, A new vision of giant planet interiors: Impact of double diffusive convection, A&A 540 A20. [32] B. Militzer, W. H. Hubbard, J. Vorberger, I. Tamblyn, S. A. Bonev, Astrophys. J. Lett. 688 (2008) L45. [33] D. C. Rapaport, The Art of Molecular Dynamics Simulation, Cambridge University Press, 2004. [34] D. Frenkel, B. Smit, Understanding molecular simulation: from algorithms to applications, Academic Press, 2002. [35] J.-P. Hansen, I. McDonald, Theory of simple liquids, Academic Press, second edn., 1986. [36] J.-F. Danel, L. Kazandjian, G. Z´erah, Numerical convergence of the self-diffusion coefficient and viscosity obtained with Thomas-Fermi-Dirac molecular dynamics, Phys. Rev. E 85 (2012) 066701. [37] J. M. Haile, Molecular Dynamics Simulation, Wiley- Interscience, 1997. [38] L. A. Collins, S. R. Bickham, J. D. Kress, S. Mazevet, T. J. Lenosky, N. J. Troullier, W. Windl, Dynamical and optical properties of warm dense hydrogen, Phys. Rev. B 63 (2001) 184110. 6
1703.10997
1
1703
2017-03-31T17:43:35
Evolution of major sedimentary mounds on Mars
[ "astro-ph.EP" ]
We present a new database of $>$300 layer-orientations from sedimentary mounds on Mars. These layer orientations, together with draped landslides, and draping of rocks over differentially-eroded paleo-domes, indicate that for the stratigraphically-uppermost $\sim$1 km, the mounds formed by the accretion of draping strata in a mound-shape. The layer-orientation data further suggest that layers lower down in the stratigraphy also formed by the accretion of draping strata in a mound-shape. The data are consistent with terrain-influenced wind erosion, but inconsistent with tilting by flexure, differential compaction over basement, or viscoelastic rebound. We use a simple landscape evolution model to show how the erosion and deposition of mound strata can be modulated by shifts in obliquity. The model is driven by multi-Gyr calculations of Mars' chaotic obliquity and a parameterization of terrain-influenced wind erosion that is derived from mesoscale modeling. Our results suggest that mound-spanning unconformities with kilometers of relief emerge as the result of chaotic obliquity shifts. Our results support the interpretation that Mars' rocks record intermittent liquid-water runoff during a $>$10$^8$-yr interval of sedimentary rock emplacement.
astro-ph.EP
astro-ph
Evolution of major sedimentary mounds on Mars: build-up via anticompensational stacking modulated by climate change Edwin S. Kite1,*, Jonathan Sneed1, David P. Mayer1, Kevin W. Lewis2, Timothy I. Michaels3, Alicia Hore4, Scot C.R. Rafkin5. 1. University of Chicago. 2. Johns Hopkins University. 3. SETI Institute. 4. Brock University. 5. Southwest Research Institute. (*[email protected]) Abstract. We present a new database of >300 layer-orientations from sedimentary mounds on Mars. These layer orientations, together with draped landslides, and draping of rocks over differentially- eroded paleo-domes, indicate that for the stratigraphically-uppermost ~1 km, the mounds formed by the accretion of draping strata in a mound-shape. The layer-orientation data further suggest that layers lower down in the stratigraphy also formed by the accretion of draping strata in a mound-shape. The data are consistent with terrain-influenced wind erosion, but inconsistent with tilting by flexure, differential compaction over basement, or viscoelastic rebound. We use a simple landscape evolution model to show how the erosion and deposition of mound strata can be modulated by shifts in obliquity. The model is driven by multi-Gyr calculations of Mars' chaotic obliquity and a parameterization of terrain-influenced wind erosion that is derived from mesoscale modeling. Our results suggest that mound-spanning unconformities with kilometers of relief emerge as the result of chaotic obliquity shifts. Our results support the interpretation that Mars' rocks record intermittent liquid-water runoff during a (cid:1233)108-yr interval of sedimentary rock emplacement. 1. Introduction. Understanding how sediment accumulated is central to interpreting the Earth's geologic records (Allen & Allen 2013, Miall 2010). Mars is the only other planet known to host an extensive sedimentary record. Gale crater and the Valles Marineris (VM) canyon system contain some of Mars' thickest (2-8km) and best-exposed sequences of sedimentary rock (Malin & Edgett 2000, Milliken et al. 2010). "The origin of [these sedimentary] mounds is a major unresolved question in Mars geology" (Grotzinger & Milliken 2012). The mounds are thought to have formed <3.7 Ga, relatively late in Mars' aqueous history and many contain sulfates that precipitated from aqueous fluids (Gendrin et al. 2005, Bibring et al. 2006, Mangold et al. 2008, Murchie et al. 2009a). The fluid source could be groundwater, rain, or snowmelt (Andrews-Hanna et al. 2010, Kite et al. 2013b). Proposed depositional scenarios (Nedell et al. 1987, Lucchita 1992) range from primarily aeolian sedimentation in a climate dry enough that aeolian erosion could define moats around the growing mounds (Catling et al. 2006, Michalski & Niles 2012, Kite et al. 2013a), through sand/dust cementation in horizontal playa-lake beds (Andrews-Hanna et al. 2010, Fueten et al. 2008, Murchie et al. 2009b), to fluvial sediment transport from canyon/crater rims into canyon/crater-spanning lakes (Grotzinger et al. 2015). At Gale crater, aeolian processes contributed to the deposition of the mound, evidenced by preserved bedforms within the stratigraphy (Milliken et al. 2014, Banham et al. 2016). Following the depositional era, aeolian erosion cut into the rocks, exposing layers and perhaps deepening moats (Day & Kocurek 2016). These paleo-environmental scenarios make contrasting predictions for the orientations of mound sediment layers and unconformities. If layers dip away from mound crests and layers have been little tilted since the time of deposition, then the mounds formed as mounds (similar to ice 1 mounds within Mars' polar craters; e.g. Brothers & Holt 2016). By contrast, gravity-driven deposition predicts layers that were originally flat-lying or oriented away from crater walls or canyon walls, with the modern topography resulting entirely from later erosion. The full internal architecture of the VM and Gale mounds cannot be directly observed, but can be inferred from outcrop measurements of layer-orientations and unconformities (Okubo et al. 2008). Layer-orientation measurements for Mars are obtained using orbiter image stereopairs to construct digital terrain models (DTMs) that form the basis for fitting planes to traces of stratigraphic surfaces (Lewis et al. 2008). From orbit it is usually not possible (due to limited resolution) to distinguish the traces of beds from the traces of lower-order bounding surfaces, although both should closely correspond to basin topography at around the time of deposition. These fitted planes usually dip in a downslope direction, but may suffer from downslope bias (e.g. Fueten et al. 2006). However, consensus on the interpretation of Mars layer data has been hindered by doubts about the accuracy of layer-orientations measured from orbiter image data, the possibility that layer-orientations do not reflect paleo slopes, and the absence of a physical mechanism that could account both for layer-orientations and for Mars' large unconformities. Corresponding to this lack of consensus in data interpretation, there are two endmember views of how Mars' mounds formed: 1. In one view, craters/canyons were fully filled by flat-lying or shallowly dipping strata, e.g. playa-deposits or fluviodeltaic deposits (Fig. 1a), and later underwent extensive erosion to their present form (Malin & Edgett 2000, Andrews-Hanna et al. 2010), presumably through wind erosion (Kite et al. 2013a, Day et al. 2016). In this view, the primary cause of non-horizontal layer-orientations is downslope measurement bias and/or postdepositional distortion (by flexure, landslides, soft-sediment deformation, tectonics, and differential compaction) (e.g. Nedell et al. 1987, Metz et al. 2010, Grotzinger et al. 2015). Preferential infilling of topographic lows through deposition (compensational stacking) is ubiquitous in well-studied aqueous sedimentary environments on Earth (Straub et al. 2009). Therefore, it is tempting to assume that Earth analogy, which has been used effectively to interpret sedimentary structures viewed by rovers (McLennan & Grotzinger 2008, Grotzinger et al. 2015), also holds at the scale of Mars basins. layers on preexisting in place by net deposition of 2. In another view, the downslope layer-tilts are primary. If this is correct, then mounds grew topographic highs (anticompensational stacking) (e.g. Niles & Michalski 2012, Kite et al. 2013a). This distinctively Martian mechanism is suggested by:- (i) growth of polar ice/dust/sand mounds by anticompensational stacking (Holt et al. 2010, Conway et al. 2012, Brothers et al. 2013, Brothers & Holt 2016); (ii) the importance of aeolian sediment transport and slope-winds on modern Mars (Spiga et al. 2011, Spiga 2011, Kok et al. 2012, Bridges et al. 2013, Silvestro et al. 2013, Kite et al. 2013a); (iii) the strong inference of layered- sediment accumulation via anticompensational stacking for some Mars equatorial layered sediments (the Medusae Fossae Formation; Bradley et al. 2002, Zimbelman & Scheidt 2012, Kite et al. 2015). Dry conditions bring aeolian processes to the fore, whereas vigorous and sustained fluvial erosion would inhibit mound construction. Therefore anticompensational stacking corresponds to a paleoenvironment where fluvial sediment transport is infrequent, consistent with models of Mars paleoclimate (Kite et al. 2013b, Mischna et al. 2013, Segura et al. 2013, Urata and Toon 2013, Halevy and Head 2014, 2 Wordsworth et al. 2013, Ramirez et al. 2014, Kerber et al. 2015, Wordsworth et al. 2015, Wordsworth 2016). 1.1. Outline. Here we construct a new database (section 2) of layer-orientations (section 3) and unconformities (section 4) within Martian mounds, in order to constrain accumulation of sedimentary rocks (section 5). We also present a new model (section 6) of mound emplacement. Implications and tests are discussed in section 7, and conclusions are listed in section 8. Our work has 3 purposes: a) To address concerns to the mounds-grew-as-mounds hypothesis of Kite et al. (2013a). These concerns are:- • That layer orientations "have not been independently confirmed" (Grotzinger et al. 2015); • That layer orientations can be accounted for by differential compaction of originally-horizontal layers over basement relief (basement = rocks that predate sedimentary infill), removing the need for slope- wind erosion during the depositional era (Grotzinger et al. 2015). We resolve these concerns in sections 2-5:- • Exhaustive tests show that layer orientations are accurate and reproducible, and that layer orientations errors (including downslope bias) are insignificant for the purpose of determining mound origin (section 2). • Layer-orientations in VM mounds show an outward dip – a direction opposite that predicted for differential compaction over basement relief (section 3). Layer- orientations in Gale are unlikely to result from differential compaction over a central ring or central peak (Gabasova & Kite 2016) (section 3). Unconformity data and draped-landslide data show that the mounds grew by anticompensational stacking at least for the topmost ~1 km of the mounds (section 4). Below this level, the data suggest two options: (i) accretion of draping strata in a mound shape, or (ii) slope-wind erosion sculpts pre-compacted sedimentary deposits, which subsequently act as a mound-shaped form over which later sediments may be differentially compacted (section 5). b) To expand the database of High Resolution Imaging Science Experiment (HiRISE)-derived layer orientation data for Mars. Using HiRISE (McEwen et al. 2007) data, we gathered 182 new layer-orientations from seven VM mounds, and increased the number of independent layer orientations for the mound in Gale crater from 80 to 126, for a total of 308 (section 3). Together, these mounds make up ~½ of the total volume of canyon/crater-hosted sedimentary mounds on Mars. Our work builds on previous studies (e.g. Fueten et al. 2006), but uses a procedure that is more accurate, includes error bars, and has been validated (section 2). Our database can be applied to many Mars geology problems (Supplementary Table). As one example, we test the prediction of Kite et al. (2013a) that systematically outward-oriented dips should be common in Mars mounds (section 5). 3 c) To propose a new model for the major unconformities in Mars' mounds. Our new analysis of stratigraphic surfaces previously-reported as mound-spanning unconformities show that these commonly have a dome shape (section 4). In order to match these data, we introduce a new model (section 6) that quantitatively integrates temporal variations and spatial variations in Mars sedimentation - for the first time for ancient Mars sedimentary-mound analysis (see also Howard 2007). Our model successfully reproduces the shape and tilt of the observed unconformities (section 6). To support these goals, we improved the SWEET (Slope-Wind Enhanced Erosion and Transport) model of Kite et al. (2013a). Kite et al. (2013a) showed how slope-winds create mounds, provided that the crater/canyon is larger than a critical size, and that long-term-average deposition rate is neither much larger nor much smaller than long-term-average wind-erosion rate (consistent with data: Bridges et al. 2012, Lewis & Aharonson 2014). Two features inherent to the relatively simple SWEET model are the absence of a physically realistic relationship between slope and shear stress, together with the lack of any explanation of mound-spanning unconformities. (Although steady forcing in SWEET can produce autogenic unconformities, the younger layers grow off to one side – rather than building on top of the thickest point of the main mound, as is commonly observed for Mars' mountains.) We solve this problem in our improved model, which we term SOURED (Stratigraphy with Obliquity-triggered Unconformities and Relief-influenced Erosion & Deposition). Specifically, we include realistic multi-Gyr calculations of Mars obliquity, and a more-realistic parameterization of terrain-influenced wind erosion derived from mesoscale modeling (Appendix A and Appendix B). 1.2. Geologic scope and geologic context. Fig. 1. Location of mounds (green triangles) investigated in this work. Background is Mars Orbiter Laser Altimeter shaded relief. 4 For this study we selected sedimentary mounds that are voluminous, light-toned, show well- exposed off-horizontal layering, have good HiRISE stereopair coverage, and either host sulfates or are stratigraphically associated with sulfates. We further selected only mounds that sit within deep, wide and steep-sided craters/canyons, attributes that favor slope winds. In both VM and Gale, the erodible sedimentary mounds are contained within craters/canyon walls made up of much-less-erodible basement materials. The 8 mounds (mensae) that were selected are Nia (7.6°S, 67.2°W), Juventae (8.0°S, 65.6°W), Mt. Sharp / Aeolis Mons (5.1°S, 137.8°E), Ophir (4.0°S, 73.5°W), Ceti (6.1°S, 75.8°W), Melas (10.7°S, 74.1°W), Coprates (12.5°S, 71.5°W), and Ganges (7.2°S, 48.9°W) – all at <15° latitude. We excluded the well-studied mounds Juventae Chasma (Catling et al. 2006, Bishop et al. 2009) and Candor Mensa (Mangold et al. 2008, Ferguson et al. 2015, Fueten et al. 2014). (The criteria exclude a large number of sedimentary accumulations on Mars: e.g., plateau deposits (Mawrth, Loizeau et al. 2015; Meridiani Planum, Hynek & Phillips 2008), the clay-bearing Terby deposits (Ansan et al. 2011), the free-standing Medusae Fossae mounds (Bradley et al. 2002, Zimbelman & Scheidt 2012, Kite et al. 2015), and veneers and smaller mounds in and around VM (e.g. Milliken et al. 2008, Thollot et al. 2012, Weitz & Bishop 2016).) Gale's mound (Aeolis Mons; also known as Mount Sharp) is the largest among 50 documented crater-hosted mounds outside the polar regions (Bennett & Bell 2016), is the primary science target of MSL (MSL Extended Mission Plan 2014), and has the best HiRISE stereopair coverage of any within-crater mound, justifying our emphasis on this within-crater mound. The 8 mounds studied here were among the first Mars sedimentary rock accumulations to be described (Malin & Edgett 2000, Malin et al. 2010). The rocks formed relatively late in Mars' aqueous history and contain hematite and sulfates (Christensen et al. 2001, Gendrin et al. 2005, Bibring et al. 2007, Weitz et al. 2008, Murchie et al. 2009a, Roach et al. 2010, Fassett & Head 2011, Ehlmann et al. 2011, Fergason et al. 2014). Most of our layer-orientation data comes from the VM mounds ("Interior Layered Deposits"; ILD). Crosscutting relationships, and contrasts in texture, thermal inertia, erodibility, and mineralogy between sedimentary-mound rocks and canyon-wall rocks all indicate that the ILD accumulated after the canyons formed (Peterson 1981, Lucchitta 2010, Okubo et al. 2008, Schultz 2002, Andrews-Hanna 2012a; see also Montgomery et al. 2009). Mound stratigraphy (Fig. 2), which usually includes at least one mound-spanning unconformity, is described in a large literature (e.g. Malin & Edgett 2000, Le Deit et al. 2013, Anderson & Bell 2010, Milliken et al. 2010, Thomson et al. 2011, Grotzinger & Milliken 2012). The following trends are a useful guide to correlation: (1) Within-mound materials below the lowest mound- spanning unconformity usually, but not exclusively, correspond to the "Laterally Continuous Sulfate" orbital facies of Grotzinger & Milliken (2012). (2) Materials found above the lowest mound-spanning unconformity within a mound usually, but not exclusively, correspond to the "rhythmite" of Grotzinger & Milliken 2012. (3) Darker-toned indurated materials draping the present topography usually correspond to the widespread "thin mesa" units of Malin & Edgett (2000) (Fig. 2). We did not measure layer-orientations on "thin mesa" units. The VM and Gale mounds are no older than the Noachian/Hesperian boundary, based on the timing of VM formation, and on the crater-retention age of Gale's ejecta (Anderson et al. 2001, Thomson et al. 2011, Le Deit et al. 2013). The topmost sedimentary rocks could be as young as Upper Amazonian (Mangold et al. 2010, Thollot et al. 2012). Therefore, crater chronology 5 permits a (cid:1233)100 Myr interval of sedimentary rock accumulation. This is consistent with other methods (Lewis & Aharonson 2014). Fig. 2. Schematic shows an idealized sedimentary rock mound within an erosion-resistant container (crater or canyon). In this paper, we focus on rocks within the topographically-defined mound (ignoring moat rocks and wall rocks). We neglect <<100m thick "thin mesa" units that drape modern topography (white outline). Layer orientations constrain mode(s) of mound emplacement (section 3). Unconformity data and associated isochores, as well as draped landslides ("ls."), further constrain basin evolution for the upper part of the mounds (section 4). The physical processes and patterns of deposition for all these rock units is uncertain, and in the words of Grotzinger & Milliken (2012), "[m]easurements of the strike, dip, and stratal geometries of layers within these units would help to place further constraints on their mode(s) of emplacement." Such measurements are the focus of the work presented here. Remarkably, despite the size of the mound-spanning unconformities of Mars, we are not aware of any previous physical model for their origin. 1.3. Relation to rover data. Co-analysis of rover data and orbiter data can increase the science value of both (Arvidson et al. 2006, Fraeman et al. 2013, Arvidson et al. 2015, Lapotre et al. 2016, Stack et al. 2016). In 2012 the Mars Science Laboratory (MSL) rover landed successfully in Gale crater, ~6 km away from the layers in Mt. Sharp / Aeolis Mons where layer-orientations are reported (Le Deit et al. 2013, Kite et al. 2013a, Stack et al. 2013). Rover results to date from Gale crater are interpreted as primarily fluviolacustrine deposits (Grotzinger et al. 2014), which are overlain unconformably by later aeolian sands (Lewis et al. 2015). Our unconformity results based on analysis of orbiter data echo recent rover discoveries in the Gale moat (Watkins et al. 2016). As MSL continues its drive (MSL Extended Mission Plan 2014), the rover's instruments may decisively constrain the 6 sediment transport mechanism for the lower layers of Gale crater's mound (section 7). As of mid-2016, the rover is ~5 km from the sulfate-bearing layers where layer-orientations are reported. Throughout the traverse to date, Mastcam rover imagery has resolution at the sulfate- bearing layers where layer-orientations are reported that is inferior to HiRISE. Specifically, the Mastcam M100 has an angular resolution of 74 µrad/pixel (85 cm/pixel for a 25° slope, 37 cm/pixel for a vertical target at 5 km) (Malin et al. 2010). This compares to HiRISE (from 250 km: 28 cm/px for a 25° slope, 25 cm/px for a horizontal target). Due to foreground obstructions, and edge-on views, layers are more easily visualized in orbiter imagery. The ChemCam Remote Micro-Imager (RMI) has a nominal resolution of 20 µrad/pixel and its potential for long-range stereophotogrammetry is exciting (Le Mouelic et al. 2015). However, we are not aware of any suitable Mt. Sharp / Aeolis Mons RMI stereopairs. Because of the (current) superiority of orbiter images compared to rover images for the purposes of stereo determination of layer-orientations within the sulfate-bearing layers, our work is largely based on orbiter data analysis. Rover imagery shows apparent dips in Mt. Sharp / Aeolis Mons layers that are qualitatively consistent with dips obtained from HiRISE DTMs. 2. Data analysis methods. 2.1. DTM production method. HiRISE DTMs and orthoimages were used as the basis for layer tracing (section 1.3). We produced CTX and HiRISE DTMs using the NASA Ames Stereo Pipeline (ASP) (Moratto et al. 2010, Beyer et al. 2014, Shean et al. 2016). As part of this processing, we developed a set of scripts that act as wrappers around the ASP routines, which increase the level of automation and computational efficiency of the DTM production (Mayer & Kite 2016). Initial CTX point clouds were aligned to MOLA shot data using an iterative closest points (ICP) algorithm before being interpolated to DTMs and orthoimages with a grid spacing of 18 m. Initial HiRISE point clouds were then similarly aligned to the CTX DTMs before being interpolated to DTMs and orthoimages with a grid spacing of 1 m (2 m for HiRISE input collected in 2×2 binning mode; Table 1). As an independent check on the quality of our DTM production workflow, we compared 3 of our HiRISE DTMs to DTMs generated from the same HiRISE stereopairs and available from the Planetary Data System (these PDS DTMs were produced using SOCET SET; Kirk et al. 2008). Because we are primarily interested in the vertical differences between DTMs produced using different methods, we coregistered the PDS-released products to our products by using tie points selected manually on the orthoimages and then applying the resulting transform to the DTMs in order to eliminate any horizontal offsets. We then subtracted the elevation values of our DTMs from the PDS-released DTMs to create a series of difference rasters. For the purposes of the layer orientation measurements in this paper, the most important differences were broad tilts across the entire image. We inspected the resulting difference rasters to characterize tilts. These tilts were 0.2°, 0.17°, and 0.09° for the 3 DTMs investigated, which is much smaller than our error bars. In addition to the stereo DTMs, digital models of each mound were extracted from MOLA gridded data. Mound basal surfaces were defined from the MOLA elevation data using cubic polynomial interpolation within mound edges. Mound crest-lines and edges were drawn by visual inspection of THEMIS mosaics. 7 2.2. Layer tracing method. Layer traces were carried out by visual inspection using orthorectified HiRISE images and corresponding DTMs (Fig. 3), following the method of Lewis et al. (2008). Most traces were >150m long. For Gale's mound, we included data from Kite et al. (2013a). Layer orientations were calculated for the best-fit plane for each layer trace. Layers were rejected if their pole error was >2° (calculated following Lewis et al. 2008). Mound DTM location Image 1 Image 2 DTM Posting PSP_006855_1750 PSP_003896_1740 PSP_008893_1760 PSP_010660_1700 PSP_005953_1695 PSP_001377_1685 1 m/pixel 1 m/pixel 2 m/pixel 1 m/pixel 1 m/pixel 1 m/pixel 1 m/pixel 1 m/pixel 1 m/pixel 2 m/pixel 2 m/pixel 1 m/pixel 1 m/pixel 1 m/pixel 1 m/pixel 1 m/pixel 1 m/pixel 1 m/pixel 2 m/pixel 1 m/pixel 1 m/pixel 1 m/pixel 1 m/pixel 1 m/pixel 1 m/pixel Coprates Mensa Coprates Mensa Coprates Mensa Ophir Mensa Ophir Mensa Ophir Mensa Ophir Mensa Melas Mensa Melas Mensa Melas Mensa Melas Mensa Melas Mensa Melas Mensa Ceti Mensa Juventae Mensa Juventae Mensa Nia Mensa Nia Mensa Nia Mensa Mt. Sharp / Aeolis Mons Mt. Sharp / Aeolis Mons Mt. Sharp / Aeolis Mons Mt. Sharp / Aeolis Mons Mt. Sharp / Aeolis Mons Mt. Sharp / Aeolis Mons Crater/ canyon DTMs produced for this study 13S 289E ESP_027723_1670 ESP_027746_1670 SE Melas 13S 290E ESP_035450_1670 ESP_034250_1670 SE Melas 13S 288E ESP_028567_1680 ESP_027657_1680 SE Melas 4S 286E ESP_034949_1760 ESP_034738_1760 Ophir 4S 286E ESP_015974_1760 ESP_020220_1760 Ophir 4S 286E PSP_008458_1760 Ophir 4S 286E ESP_017886_1760 ESP_017675_1760 Ophir PSP_007812_1700 10S 286E SC Melas PSP_002630_1695 10S 286E SC Melas PSP_001852_1685 11S 286E SC Melas 11S 286E ESP_012361_1685 ESP_012572_1685 SC Melas 11S 285E ESP_033169_1690 ESP_032747_1690 SC Melas 10S 285E ESP_028633_1695 ESP_034382_1695 SC Melas 6S 283E PSP_002841_1740 W Candor 8S 294E ESP_017411_1715 ESP_017266_1715 E Candor 7S 294E ESP_037586_1725 ESP_037731_1725 E Candor 8S 293E ESP_034896_1725 ESP_036452_1725 E Candor 7S 292E ESP_031982_1730 ESP_031916_1730 E Candor 7S 292E ESP_014154_1730 ESP_014431_1730 E Candor 5S 137E PSP_007501_1750 Gale 5S 137E ESP_012195_1750 ESP_012340_1750 Gale 6S 138E PSP_002464_1745 Gale 5S 138E ESP_016375_1750 ESP_016520_1750 Gale 5S 138E ESP_030880_1750 ESP_030102_1750 Gale 5S 137E ESP_012907_1745 ESP_013540_1745 Gale Additional DTMs from Kite et al. (2013a), produced and analyzed by K.W. Lewis. 2° (worst-case) error assumed. Gale Gale Gale Gale Gale Additional DTM produced and traced by Okubo (2014), traces not included in the main database. W Candor Additional DTMs produced and analyzed by Alicia Hore (Hore 2015), summarized in Fig. 8f but not included in the main database Ganges Ganges Ganges Ganges Ganges Ganges 5S 138E /PSP_008938_1750 5S 137E ESP_023957_1755 ESP_024023_1755 PSP_001752_1750 5S 137E 5S 137E PSP_009294_1750 6S 138E ESP_014186_1745 ESP_020410_1745 Mt. Sharp / Aeolis Mons Mt. Sharp / Aeolis Mons Mt. Sharp / Aeolis Mons Mt. Sharp / Aeolis Mons Mt. Sharp / Aeolis Mons 1 m/pixel 1 m/pixel 1 m/pixel 1 m/pixel 1 m/pixel PSP_002550_1725 PSP_006519_1730 7S 311E PSP_007020_1730 7S 311E ESP_013059_1725 ESP_012993_1725 7S 311E PSP_003618_1725 7S 312E ESP_011648_1730 ESP_011582_1730 7S 311E ESP_018162_1730 ESP_018633_1730 PSP_007521_1725 7S 311E Ganges Mensa Ganges Mensa Ganges Mensa Ganges Mensa Ganges Mensa Ganges Mensa Table 1. Table of DTMs. Layers were traced on the HiRISE DTMs listed in Table 1. Linear subhorizontal features observed in Mars outcrops from orbit might correspond to depositional beds, first-order bounding surfaces, deflation surfaces, diagenetic bands, or even buttress unconformities or wave runup features (Rubin & Hunter 1982, Kocurek 1988, Edgar et al. 2012, Parker et al. 2014). Where rovers have explored sulfate-rich rocks on Mars, shallow/early diagenesis blurs the distinction between diagenetic bands and depositional beds. (Later diagenetic fronts need not be parallel to depositional beds; Davies & Cartwright 2002, Borlina et al. 2015). Therefore, we n.a. n.a. n.a. n.a. n.a. n.a. PSP_007877_1725 PSP_001488_1750 PSP_009149_1750 PSP_003176_1745 7S 284E PSP_001641_1735 PSP_002063_1735 1 m/pixel PSP_008437_1750 Ceti Mensa 8 aimed to trace stratigraphic surfaces that closely corresponded to basin-scale topography at the time of deposition (we refer to these stratigraphic surfaces as "layers"). To maximize the likelihood of tracing layers, we followed Lewis (2009) and avoided drawing traces that crossed faults in the rocks where displacement may have occurred, and areas adjacent to faults where folding can distort layers into non-planar surfaces. We avoided tracing on landslides, convolute folding (Metz et al. 2010), superscoops, zones of apparent soft-sediment deformation, and 'thin mesa' materials (Malin & Edgett 2000). Examples of the trace locations and corresponding results are shown in Fig. 3. The fine scale and high degree of lateral continuity of layers (e.g. Fig. 3) is strong evidence that the observed layering represents true depositional bedding and not, for instance, diachronous facies boundaries or late-diagenetic alteration horizons (Le Deit et al. 2013, Stack et al. 2013, Milliken et al. 2014). (a) (b) 9 dips (°) for and 3. (a) Traces corresponding the Fig. ESP_017411_1711/ESP_017266_1715 stereopair. (b) Detailed trace identification for part of the reentrant canyon shown in Fig. 4a (ESP_012907_1745/ESP_013540_1745 stereopair.) Errors in tracing a layer on a slope on an orthorectified image will produce a downslope bias in plane-fits to the trace using the corresponding DTM. Four tests show that downslope bias in our dataset does not affect our conclusions:- part of (1) Reentrant-canyon test (Kite et al. 2013a). For the reentrant canyon at 137.2°E 5.3°S (Fig. 4a), a dominant direction of layer azimuth contrasts with a nearly complete radial rotation in dominant downslope direction. We found that layers dip in a systematic direction, typically perpendicular to local downslope. This rules out severe downslope bias. (2) Resolution-sensitivity test. We compared the traces of identical layers at different image grid spacings (Fig. 5). If downslope bias affects the HiRISE layer- orientations (1m/pixel elevation model), then the same layers traced on CTX (18m/pixel DTM) will suffer a bias that is more severe. For layers in the canyon at 137.2°E 5.3°S (Fig. 5), we obtained two metrics of downslope bias (Fig. 5): (a) the angle between the best-fit plane and local topography projected onto the vertical plane parallel to steepest topographic slope, and (b) the map-plane angle between the best-fit plane and the topographic downslope. We do not find any systematic tendency for the CTX layer-orientations to be rotated downslope relative to the HiRISE layer-orientations, suggesting that the HiRISE bias is itself small. Full Dip Direction (° CCW from E) Cut Dip (°) Dip Direction (° CCW from E) Topo. Dip (°) Topographic Dip Direction (° CCW from E) Difference from Topo. (°) Full Cut Proximity Mesa Mesa HiRISE Latitude Longitude Image No. (°) (°) Full Dip (°) ESP _012551 -4.852 137.255 3.03 -150.44 _1750 ESP _012551 -4.948 137.242 3.53 149.11 _1750 ESP _012551 -4.97 137.271 2.62 117.80 _1750 ESP _016375 -5.346 138.528 2.50 39.00 _1750 ESP _016375 -5.335 138.533 0.62 89.42 _1750 ESP _012361 -11.290 -74.68173 5.21 172.61 _1685 ESP _012361 -11.220 -74.69151 3.13 -108.08 _1685 1 2 1 2 1 2 1 2 1 2 1 2 1 2 1.09 -105.03 21.67 21.01 1.77 -151.54 23.40 -163.19 8.51 166.50 28.19 178.68 2.04 44.01 29.52 10.42 174.76 33.70 6.61 10.01 2.66 98.69 16.19 178.11 0.16 179.08 26.73 34.38 4.11 36.29 18.46 -153.49 0.73 127.35 26.73 34.38 4.11 36.29 18.46 -153.49 7.07 169.71 12.65 -171.69 5.21 -169.60 8.10 3.32 -78.37 6.09 25.84 -27.64 3.87 -144.70 10.65 -175.29 24.67 20.45 25.17 32.39 34.58 15.06 24.23 20.91 26.38 18.75 7.75 12.78 6.36 9.86 Difference from Topo. (°) 22.32 21.66 19.94 27.93 43.82 15.91 26.86 22.52 26.78 22.52 6.35 13.20 4.74 7.57 to Topo.vs. Full (°) 2.34 -1.21 5.22 4.46 -9.24 -0.85 -2.62 -1.61 -0.40 -3.77 1.40 -0.41 1.62 2.29 Table 2. Downslope bias is shown to be small by elliptical-mesa check (Fig. 4b). "Full" refers to the entire elliptical trace. "Cut" refers to an arcuate subset of the elliptical layer, chosen to be oriented along the long-axis of the elliptical mesa (this is the worst case). The final column shows the rotation of the 10 pole to the best-fit plane into the downslope direction, which is positive when data are consistent with downslope rotation, and negative when the data show uplslope rotation. The highlighted rows correspond to traces that are shown in Figure 4b. (3) Circular-mesa test. (Fig. 4b, Table 2). In rare cases, conical topographic features show layers that can be traced in a closed loop, rather than an open curve. Because there is no obvious 'downslope direction' for closed-loop traces, the topography-induced measurement error (downslope bias) of closed-loop traces is close to zero. After tracing 7 such mesa-encircling layers, we split each elliptical trace along its ~200m-long major axis (the worst-case for downslope bias). This creates 14 test traces with a clear downslope direction and a high aspect ratio (Table 2) – again, the worst-case for downslope bias. The DTM under each trace was clipped by minimum bounding rectangle, and best-fit planes were fit to each of the traces and to the elevation data contained within the minimum bounding rectangle for that trace. The best-fit poles to the halved test cuts are consistent with zero downslope rotation. For n = 14, test cuts are -0.2° closer to topography on average (i.e. we find the unexpected result of upslope rotation), with a standard deviation of 3.7°, minimum of -9.2°, and maximum of 5.2°. (4) Geologic-control test. HiRISE DTM layer-dip measurements made using the same technique show near-horizontal layers in areas where near-horizontal layers are expected from geological context. Specifically, near-horizontal layers have been measured from Eberswalde's delta topsets, Holden's delta topsets, and the Juventae plateau layered deposits (Irwin et al. 2015, Stack et al. 2013). These near-horizontal measurements are reported from places where the present-day erosional surface slopes steeply and so might be expected to produce large downslope bias. This geologic "control case" strongly suggests that off-horizontal Mars layer orientations are not artifacts of downslope bias, but rather geological. (b) (a) Fig. 4. To show that downslope bias does not affect our conclusions. (a) Examples of layers (yellow) showing similar dips whether traced on CTX DTMs (white) or HiRISE DTMs (black) in a reentrant canyon at 137.2°E 5.3°S. Gray contours show 100m topographic intervals. Brown is high. Backdrop is HiRISE DTM shaded relief. (b) Example of arcuate subsets of a layer (yellow) on a circular mesa. Mesa slope is ~23°, yet plane-fits to arcuate subsets of the layer (black symbols) show no downslope bias relative to the plane-fit to the entire elliptical trace (red symbol). 11 Although our checks indicate that downslope bias does not affect our conclusions, our database includes a small number (<5) of measurements where downslope bias may set the dip azimuth. These measurements all have a minimum bounding rectangle that has an aspect ratio greater than 8:1, i.e. small curvature. a) b) c) Fig. 5. CTX-vs.-HiRISE layer orientation test using layer traces from the area of Fig.4a. (a) Cartoon showing how CTX-derived and HiRISE-derived layer orientations are projected onto the plane containing the downslope (topography) vector. (b) Results of HiRISE-CTX comparison. (c) Quantifying geologic noise: showing divergence between pole-fits to layers as a function of separation. Orange line shows 2° threshold. Red circles mark the mean of angular differences, binned by seperation. Red whiskers correspond to the standard deviation of the logarithms of the binned data. 12 We found same-worker reproducibility within error. Formal DTM precision makes a negligibly small contribution to the error. Consistency between measurements by workers in the University of Chicago and Johns Hopkins University labs using the same procedure was demonstrated. Between-lab reproducibility for poles-to-layer-planes in NW Gale (JHU vs. Chicago) was 1.3° on average (standard deviation 1.0°, worst-case 3.7°, n = 17), which is less than our error bars. Outward dips at Mt. Sharp / Aeolis Mons have been independently confirmed by Fraeman et al. (2013), Le Deit et al. (2013), and Stack et al. (2013). Within-measurement errors (the residuals of measured points around the best-fit plane) were quantified using the method of Lewis (2009), which is a conservative approximation to a 95% regression-error estimate. Points with >2° error were rejected. The mean pole error in the whole database is 0.98°. Same-worker reproducibility averaged 1.3°, with a standard deviation of 1.0°. The same-worker reproducibility check layers were chosen to systematically span a range from smallest to largest ΔZ, where ΔZ is the absolute range of elevation values. We did not find any tendency for reproducibility to get worse with decreasing ΔZ. However, small-ΔZ traces remain sensitive to small-scale geologic variation (e.g. fractures, boulders), so caution is warranted in interpretation of individual traces with ΔZ < 3.5 m. To quantify between-measurement variations (geologic noise), we plotted (for each DTM) pairwise angular differences between the poles-to-layer-planes as a function of the pairwise separation between median {x,y} positions of individual traces (Fig. 5c). We found that the pairwise differences are well-fit by a line that increases log-linearly with separation, and intersects 180 m at ~4° (Fig. 5c). Within-DTM differences in layer orientation can greatly exceed our error bars, and so are likely real (geological). These layer-orientation differences could be primary depositional features, or the result of short-wavelength postdepositional tilting. Together, these tests show that our measurements are accurate and reproducible and that downslope bias does not affect our conclusions. We cannot exclude a selection bias (layers that dip close to slope will have corrugated outcrops that are easier to measure). However, our measurements cover many mounds and a broad range of stratigraphic elevation, minimizing this effect. For the purpose of understanding mound build-up, within-DTM scatter in the measurements (km-wavelength geologic noise) sets the practical limit on interpretation - not measurement precision or accuracy. Results are given in Figs. 6-9 and section 3. 2.3. Fitting of stratigraphic surfaces interpreted as erosional unconformities. We traced stratigraphic surfaces (interpreted by previous workers as erosional unconformities) in W. Candor, Ophir, and Gale (Anderson & Bell 2010, Thomson et al. 2011, Le Deit et al. 2013, Lucchitta 2015). We interpret the traces as unconformities on the basis of a sharp break in tone, erosional or layering style, crater density, or slope, at a stratigraphic level that, in at least one location, corresponds to an unconformity (shown by buried craters, or by truncated layers) (e.g. Fig. 9). In none of these cases is definitive unconformity mapping possible using CTX data alone, and complete HiRISE coverage is not available. In W. Candor and Gale, we believe that the traces do correspond to major unconformities (e.g. Fig. 9), and that our traces follow a stratigraphic surface sufficiently closely to determine the qualitative paleotopography (dome, trough, saddle, or roughly flat), and to put lower bounds on isochore measurements. Next, we 13 made use of DTMs constructed using CTX stereo data (for Gale) or using MOLA data (for W. Candor). For segments of the trace where we were confident about the location of the unconformity, we calculated the total relief (max. elevation – min. elevation) of the unconformity trace. Next, the digitized points were interpolated to form unconformity surfaces using (i) inverse distance weighting, (ii) planar interpolation, and (iii) quadratic global polynomial interpolation. In principle, the interpolation procedure is subject to a dome bias that is analogous to the downslope bias in layer-orientation fits. In practice however, the >km total relief of the unconformity traces means that any such bias is unimportant for the purpose of determining the best-fit shape of the stratigraphic surface. Following interpolation, we subtracted these surfaces from present-day topography to create thickness contours (isochores) for the material above the within-mound unconformities. Results are given in Figs. 10-12 and section 4.1. 2.4. Identification of draped landslides. We identified mass-wasting units (flows, slides, spreads, falls and topples), which we refer to as "landslides", using THEMIS and CTX images. Comparison with a preliminary U.S. Geological Survey geologic map of the Central Valles Marineris (Fortezzo et al. 2016) shows that our identifications of mass-wasting zones agree. We additionally looked for locations where undeformed layered materials superposed the source zones of the landslides, indicating layered material deposition after moat formation (Anderson & Bell 2010, Okubo 2014, Neuffer & Schultz 2006). Results are given in Figs. 13-14 and section 4.2. 3. Layer-orientation results. 3.1. Overview. Among our measurements (308 layer dips extracted from 30 DTMs), most strata within VM and Gale's mound were found to dip away from mound crests (Fig. 6b). For layers above the mound base, the dip azimuth of 87% of the measured layers falls within 90° of the vector directly away from the nearest mound crest (mound centroid for Gale); 57% are aligned within 45°. This tendency is equally strong in Gale's mound (n = 126) and VM (n = 182) (Fig. 7d). The median dip of the measurements in our database (5°) corresponds (for an 80-km wide mound) to 3.5 km of relief on a stratigraphic surface. Indeed, canyons carved into Gale's mound show easily- observable relief of 500m on individual layers. Lowermost strata (≤0.5 km above the interpolated basal surface), which will soon be visited by the MSL rover, still dip preferentially away from mound crests (Fig. 6c). This structural consistency with elevation contrasts with the mineralogical variability observed at Gale and elsewhere (Milliken 2010). Sulfate detections specifically correspond to outward-dipping layers at Ganges Mensa, Melas Mensa, and Gale's mound (Chojnacki & Hynek 2008, Fueten et al. 2014; Fig. 8), as well as for Hebes and Candor Mensae (Schmidt 2016, Jackson et al. 2011, Fueten et al. 2014). Furthermore, draping layers in SW Melas Chasma show sulfate signatures (Weitz et al. 2015), suggesting that sulfate-bearing rocks on Mars can form at primary depositional angles that are far from horizontal. Therefore, preferentially-outward dips are not restricted to the spectrally-bland, capping 'rhythmite' facies identified in many sedimentary deposits on Mars, including the uppermost Gale strata (Grotzinger & Milliken 2012, Lewis & Aharonson 2014). Although the rhythmite facies lies topographically above the northern rim of Gale crater, its induration still suggests cementation involving liquid water (Lewis et al. 2008). 14 Just as the mineralogical transitions up-section do not correspond to the end of surface liquid water on Mars, our measurements further suggest that they need not be accompanied by a change in the physical process of deposition. (a) (b) (d) (c) Fig. 6. Layer-orientations summary: (a) Expectations for primary depositional orientation (Leeder et al. 2011, Moore & Howard 2005, Davis 2007, Kite et al. 2013a, Grotzinger et al. 2015). "Delta foresets" refers to a container-wall sediment source. (b) Results, for measurements above the interpolated basal surfaces of the mounds. Solid and dashed contours enclose 50% and 68% of data, respectively, after accounting for heteroskedastic error. Marginalizing over dip, 87% of the azimuth data lie within 90° of a line directed away from mountain crest. (c) Distribution of layer orientations relative to elevation above interpolated basal surface of mound (lower strata are ≤0.5 km above mound base; upper strata are >0.5 km above mound base; 22.5° bins). 15 (a) (b) (c) (d) (e) (f) Fig. 7. Layer-orientation details: (a) Data from Gale (black) compared to data from VM (red). Solid and dashed contours enclose 50% and 68% of measurements, respectively. (b) Layer- orientations ≤0.5 km elevation above interpolated basal surface (purple) compared to layer- orientations >0.5 km above interpolated basal surface (green). Solid and dashed contours enclose 50% and 68% of measurements, respectively (Compare Fig. 6c). (c) Dip azimuth of all layers in Gale compared to dip azimuth of all layers in VM. (d) Distribution of layer orientations relative to elevation above interpolated basal surface of mound (22.5° bins). (e) Distribution of layer dip 16 with vertical offset below mound summit (filled symbols for Gale data, open symbols for VM data). Gray line shows maximum error of included data points, and black line shows mean error of included data points. Solid and dashed contours enclose 50% and 68% of measurements, respectively. (f) Cumulative probability distribution function for dip amplitude, categorized by distance below mound summit. Median dip >2 km below mound summit is 4.7° deg (n = 216), less than median dip ≤2 km below mound summit (7.0°, n = 92) (Fig. 7). Similarly, layers >1.5 km above the interpolated basal surface (n = 99) dip more steeply (median 7.5°) than layers ≤1.5 km above the interpolated basal surface (n = 209, median dip 4.5°). Gale data are shallow-dipping and >3km below mound summit, and removal of Gale data (or removal of VM data) would remove the dip-versus- elevation correlation in our database. The tendency for layers above mound base surface to dip away from the center of the mounds is insensitive to the error threshold (cutoff) beyond which data are discarded. Our nominal cutoff of 2° gives 87% of layers dipping away from the mound center. A cutoff of 1° gives 84% of layers dipping away from the mound center. Accepting all measurements, with no cutoff, yields ~10% more layer-traces but no change in the percentage of layers that dip away from the mound center (87%). The data indicate a strong preference for layers to be oriented away from mound centerlines. 17 3.2. Seven of the eight mounds investigated individually exhibit the outward dips predicted by Kite et al. (2013a). Mound-by-mound analysis shows that seven of these eight mounds studied in this paper individually exhibit the outward dips predicted by Kite et al. (2013a) – Gale's mound, plus Ceti, Ophir, Melas, Ganges, Nia, and Juventae Mensae (Fig. 8). For each mound, we visually inspected the intersections of layers in our CTX orthophotos with contour lines generated using our mound-spanning CTX DTM mosaics and confirmed that these structure contours are qualitatively consistent with the patterns described below using HiRISE data. (a) (b) 18 (c) (d) 19 (e) (f) 20 (g) (h) 21 Fig. 8. Mound-by-mound layer-dip data (dips in degrees). Color scale is clipped at high and low elevations in order to emphasize mound topography. Light gray shows the outlines of the HiRISE orthoimages/DTMs. Dark gray lines enclose the topographically-defined mounds. Red lines show the mound crest-lines. Strike-dip symbols and labels indicate average orientations of all layers traced on the corresponding orthoimage/DTM. (a) East Candor Chasma contains Nia Mensa (west) and Juventae Mensa (east). White rectangle shows the location of the draped landslide shown in Fig. 13. (b). West Candor Chasma contains Ceti Mensa (drawn to include Nia Tholus). Crest-line is drawn across a late-stage erosional window in central Ceti Mensa. See also Fueten et al. (2006) and Murchie et al. (2009a, 2009b). (c). Ophir Chasma contains Ophir Mensa. Crest-line is drawn to cut across a late-stage erosional window in the east of the mound. See also Wendt et al. (2011). (d). South-Central Melas Chasma contains Melas Mensa. Crest-line is drawn to crosscut a topographic low that is interpreted as an erosional trough. An alternative scenario, in which Melas Mensa is in fact two mensae, is indicated by the dashed line. (e). South East Melas Chasma contains Coprates Mensa. Crestline is drawn to cut across some small troughs. (f). Ganges Mensa (data from Hore 2015). For DTMs that straddle the mound centerline, we plot the average for data north of the centerline separately from the average for data south of the centerline. (g). Gale crater contains Mt. Sharp / Aeolis Mons. Gray star shows the centroid of Gale's mound. Main canyon of Gale's mound is to the W. White rectangles show location of draped landslide and draped canyon in Fig. 14a and 14b, respectively. (h). Showing Gale data in relation to Gale crater. Range rings (white dashed lines) show distance of candidate peak-ring (Allen et al. 2014) from Gale's central peak (red star). East Candor Chasma (Fig. 8a). E Candor contains the tallest sedimentary rock mounds on Mars, Juventae Mensa and Nia Mensa. Five HiRISE DTMs were obtained, whose mean dips (n = 46) systematically point away from the present-day mound crests (Fig. 8a). One DTM (ESP_034896_1725/ESP_036542_1725) has only 2 traces within the 2° error threshold. Near the base of Nia Mensa, an arcuate feature has been interpreted as a delta (Le Deit et al. 2008): in our layer-trace database, this feature shows northward dips that lack the fanning-out dip-directions expected of a delta. Nia Mensa and Juventae Mensa are dusty, and we are not aware of published sulfate detections there (Roach 2009). West Candor Chasma (Fig. 8b). Measured dips within Ceti Mensa are outwards. Our Ceti Mensa observations support Okubo's (Okubo et al. 2008, Okubo 2010, Okubo 2014) interpretation of a paleo-moat. The red point in Fig. 8b was calculated by taking the average of the 210 dips reported by Okubo (2014) from the northernmost (highest in elevation) outcrop of the CeMk unit as defined in Okubo (2014). Sulfate minerals are found at levels stratigraphically equivalent to many of the outward dips (Gendrin et al. 2005, Mangold et al. 2008, Murchie et al. 2009). Our one additional DTM W of the mound centerline has 3 good traces, showing generally W-directed dips. Ophir Chasma (Fig. 8c). Four HiRISE DTMs (n = 48), all from the western end of the Ophir Mensa mound, show dips that are directed away from the mound crest except for the lowest- elevation DTM, which shows dip directions that parallel the mound crest. The DTM marked "2.8°" shows layers that drape the lowest part of the canyon wall in the W of the DTM, and layers that dip steeply towards the canyon wall in the E of the DTM. Dust on Ophir Mensa complicates spectroscopy. We know of one kieserite detection on Ophir Mensa at the level of our measurements (Gendrin et al. 2005, Chojnacki & Hynek 2008). 22 affected (ESP_012361_1685_ESP_012572_1685 South Central Melas Chasma (Fig. 8d). SC Melas Chasma hosts Melas Mensa, an ~3km-high mound with 6 HiRISE DTMs (n = 60). Layers dip away from the mound centerline around the mound. Sulfate minerals are found at levels stratigraphically equivalent to many of the outward dips (Gendrin et al. 2005, Chojnacki & Hynek 2008). However, DTMs of sedimentary rock layers at the base of the mound show more variable layer-orientations, including some layers dipping back toward the mound. Melas Mensa has a N-S aligned medial trough. If this trough is used to divide the mound into two mounds, then traces from two elevation maps are most strongly and PSP_005953_1695_PSP_002630_1695), totaling 21 traced layers. Under likely measures of two central ridge features for each mound considered separately, 9 of these traces would be oriented more directly away from central ridges, and 13 would be oriented less directly. Therefore, our conclusion is not affected by whether Melas Mensa is considered as 1 or 2 mountains. South-East Melas Chasma (Fig 8e). SE Melas Chasma hosts Coprates Mensa, a relatively low ~2km mound with layer orientations that do not match those predicted by Kite et al. (2013a). Two of our three DTMs (n = 23) show layer orientations that are variable, but average out to dips that are parallel to the mound-crest; the remaining DTM shows dips that slope back toward the mound crest. Sulfates exist at the level of some of our measurements in the E of Coprates Mensa (Gendrin et al. 2005, Chojnacki & Hynek, 2008). Ganges Chasma (Fig. 8f). A comprehensive (6 DTMs) study of Ganges Mensa (Hore 2015) shows systematic outwards dips (Fig. 8f). Hore (2015) does not provide error bars, so we do not include Ganges data in Figs. 6-7. Sulfates are common in Ganges Mensa (Chojnacki & Hynek 2008), including at the stratigraphic level of our measurements. Gale crater (Fig. 8g, 8h). Our database for Gale's mound, Mt. Sharp / Aeolis Mons (n = 126) includes new measurements for 6 DTMs. We combine these with layer orientations from 5 DTMs presented in Kite et al. (2013a). Layers dip systematically away from the mound center, including up the main canyon of Mt. Sharp / Aeolis Mons and close to the center of the mound. However, we did not find many traceable layers in HiRISE stereopairs close to the center of the mound, and the 2 DTMs closest to the center of the mound each have only 3 traces within the 2° error threshold. The observed persistence of outward layer dips up the main canyon of Gale's mound rules out the hypothesis that a peak ring is solely responsible for the layer-orientations. Arcuate mounds between 40 km and 50 km from Gale's central peak may be eroded remnants of a peak ring (Allen et al. 2014). If this central ring persists underneath Gale's mound, then it might affect layer orientations locally. We did not find clear evidence for a peak ring effect on layer orientations; it is possible that further analysis of HiRISE DTMs might turn up such evidence. Whether or not a "peak ring effect" is detectable in the orientation of some the layers of Mt. Sharp / Aeolis Mons, the outward layer dips we observe occur at a wide range of distances from Gale's central peak – too wide a range for a peak ring to explain the outward dips. Because Gale's central peak is volumetrically negligible compared to the volume of Mt. Sharp's lower unit, and is visibly intact, erosion of Gale' central peak cannot account for the deposits contained within Aeolis Mons / Mt. Sharp's lower unit. Hebes Mensa (not shown) shows systematic outward dips, and sulfate detections, but a flat unconformity (Jackson et al. 2011, Schmidt et al. 2015, Schmidt 2016). 23 4. Stratigraphic unconformities and draped landslides. a) b) Fig. 9. Examples of unconformities. (a) HiRISE snapshot of a mound-spanning unconformity in a within-crater mound (Gale's mound, near 4.83S 137.41E). Note embedded crater in bottom left. (b) HiRISE snapshot of a large unconformity in a within-canyon mound (Ceti Mensa, near 5.80S 76.47W). 4.1. Stratigraphic unconformities. We analyzed the major unconformities reported at Gale and W Candor (Anderson & Bell 2010, Lucchitta 2015, Thomson et al. 2011, Le Deit et al. 2013) (Figs. 9-12). In every case (Table 3), present-day exposures of these surfaces show large (1-4 km) relief (Malin & Edgett 2000, Fueten et al. 2014). Using procedures described in Section 2.3, we identify and interpolate these stratigraphic horizons across each deposit. The interpolated unconformity surfaces dip steeply (Thomson et al. 2011), analogous to the modern mound forms, and consistent with past wind erosion (Heermance et al. 2013). Interpolated surfaces typically define paleo-domes within the interior of each mound (Table 3). Paleo-dome summits are usually close to modern topographic highs (Fig. 9). Furthermore, isochores show preferential deposition near paleo-dome summits (Figs. 9-12). These paleo-domes, defined by unconformable surfaces deep within the stratigraphy, strongly suggest the existence of moats during the interval of deposition (i.e., anti- compensational stacking). Our data suggest that that the mound-spanning unconformities truncate underlying layers, usually dip toward the canyon edge or crater rim, and are draped by parallel layers (Anderson & Bell 2010, Holt et al. 2010, Okubo 2014). Draping implies that post-unconformity sediments were wind-transported (similar to Holt et al. 2010). Water-transported sediments would onlap the paleo-dome. We looked for, but did not find, evidence for onlap. We are not aware of basin-scale unconformities of this type on Earth. 24 (a) (d) (b) (e) (f) (c) Fig. 10. Paleo-domes within Mars mounds. (a) Topography (blue contours) for Gale's mound (Gale is 155 km diameter). Red lines show the trace of stratigraphic surface interpreted as unconformity by Anderson & Bell (2010), which has >1 km of relief. ê=sedimentary-mound summit. (b, c) Colors show paleo-topography of Gale's mound, interpolated using (b) inverse- distance weighting and (c) quadratic polynomial interpolation. Black contours show isochores for late-deposited material. p,r=high points of unconformity surfaces (filled for IDW- interpolated, open for quadratic-interpolated). ×=locations of maximum thickness for upper units. (d, e, f): As (a-c), but for stratigraphic surface interpreted as unconformity from Thomson et al. (2011). Background is shaded relief of CTX DTM mosaic. 25 (a) (b) (c) (d) (f) (e) Fig. 11. Paleo-domes within Mars mounds. (a) Topography (blue contours) for Gale's mound (Gale is 155 km diameter). Red lines show the trace of the base Syu surface interpreted as unconformity by Le Deit et al. (2013), which has >1 km of relief. ê=sedimentary-mound summit. (b, c) Colors show paleo-topography of Gale's mound, interpolated using (b) inverse- distance weighting and (c) quadratic polynomial interpolation. Black contours show isochores for late-deposited material. p,r=high points of unconformity surfaces (filled for IDW- interpolated, open for quadratic-interpolated). ×=locations of maximum thickness for upper units. (d, e, f): As (a-c), but for base Bu surface interpreted as unconformity from Le Deit et al. (2013) (this surface is best-fit by a saddle: see Table 3). 26 (a) (b) (c) (d) (f) (e) Fig. 12. Paleo-domes within Mars mounds. (a) Topography (blue contours) for Ceti Mensa (Candor Chasma is ~120 km wide). Red lines show the trace of base Caprock surface interpreted as unconformity by Lucchitta (2015), which has >1 km of relief. ê=sedimentary-mound summit. (b, c) Colors show paleo-topography of Ceti Mensa, interpolated using (b) inverse-distance weighting and (c) quadratic polynomial interpolation. Black contours show isochores for late- deposited material. p,r=high points of unconformity surfaces (filled for IDW-interpolated, open for quadratic-interpolated). ×=locations of maximum thickness for upper units. (d, e, f): As (a-c), but for base Rimrock surface interpreted as unconformity by Lucchitta (2015). Background is THEMIS VIS mosaic. 27 L o c a t i o n M o u n d u n c o n f o r m i t y ) ( i n t e r p r e t e d a s S t r a t i g r a p h i c s u r f a c e R M S e r r o r o n p l a n a r f i t R M S e r r o r o n q u a d r a t i c f i t Q u a d r a t i c f i t u n c o n f o r m i t y ) s u r f a c e ( i n t e r p r e t e d a s R e l i e f o n s t r a t i g r a p h i c u n c o n f o r m i t y ( I D W i n t e r p o l a t i o n ) 1 M a x i m u m t h i c k n e s s a b o v e u n c o n f o r m i t y ( q u a d r a t i c i n t e r p o l a t i o n ) M a x i m u m t h i c k n e s s a b o v e 0.7 km 0.3 km 1.5 km 0.9 km 2.1 km 2.1 km W Candor Ceti Mensa W Candor Ceti Mensa W Candor Ceti Gale Gale Gale Gale Ophir* Mensa Mt. Sharp / Aeolis Mons Mt. Sharp / Aeolis Mons Mt. Sharp / Aeolis Mons Mt. Sharp / Aeolis Mons Ophir Mensa * base caprock2 (Lucchitta 2015) base caprock (Lucchitta 2015), high confidence region base Rimrock2 (Lucchitta 2015) base Bu (Le Deit et al. 2013) base (Syu2 + Cyu) (Le Deit et al. 2013) base 'Upper mound (Um) formation' (Thompson et al. 2011) base 'Upper unit' (Anderson & Bell 2010) "Marker horizon" (Wendt et al. 2011 / Peralta et al. 2015) 500 m 204 m Saddle, high axis oriented 95° CW from N, saddle located near 76.3° W 5.8S. 301 m 77 m Dome, ellipticity 1.4, centered at 76.35°W 5.87°S, long axis 6° CW from N. 475 m 154 m 168 m Dome, ellipticity 3.2, centered 76.16°W 5.71°S, long axis 32° CCW from N. Saddle, high axis oriented 13° CCW from N, saddle located near 137.8°E 4.7°S. 86 m 220 m 132 m Dome, ellipticity 1.6, center 137.86°E 5.08°S, long axis 96° CW from N. 826 m 426 m Dome, ellipticity 1.3, center 137.80°E 5.25°S, long axis 41°CW from N. 3.2 km 1.8 km 2.7 km 2.4 km 3.3 km 4.0 km 1.4 km 0.4 km 1.4 km 1.5 km 2.7 km 2.8 km 219 m 114 m Dome, ellipticity 1.7, center 137.88°E 4.77°S, long axis oriented 13°CW from N. 1.4 km 1.6 km 0.6 km. 850 m* 339 m* Dome, center 73.54°W 3.99°S, ellipticity 1.8, long axis 86° CW from N. * 5.2 km* 1.2 km* 1.1 km* Table 3. Table of mound-spanning stratigraphic surfaces interpreted as unconformities. Notes: 1. 100 nearest-neighbouring points (i.e., vertices on unconformity trace), 100km search radius, quadratic weighting. 2. The 'ildu' of Fortezzo et al. (2016) corresponds to base rimrock in places, and to base caprock in other places. However, fitting a quadratic surface to the Fortezzo et al. (2016) 'ildu' would also produce a dome. *Doubtful. The "marker horizon" trace follows the top of a cliff near 74°E 4°S, in an area without stratigraphic cues; all measurements associated with this trace are lower confidence. 28 4.2. Draped landslides. Fig. 13. Draped landslides in E. Candor. (a) Left panel: landslide on S side of Juventae Mensa (65.9W 8.1S). Range of elevations ~3 km. Right panel: Sketch interpretation (200m topographic contours). Crosshatching denotes post-slide sedimentary rock. All parts of figure use CTX DTMs. Location indicated in Figure 8a. (a) (b) Fig. 14. Draped landslides and draped canyons in Gale. (a) Left panel: landslide on N side of Gale's mound (137.9E 4.8S). Range of elevations ~4 km. Right panel: Sketch interpretation (200 m topographic contours). (b) Left panel: Draped landslide on W side of Gale's mound (137.5E 5.1S). Range of elevations ~2 km. Right panel: Sketch interpretation (200 m topographic contours). Diamonds denote post-canyon sedimentary rock. Stippling denotes mobile cover. All parts of figure use CTX DTMs. Locations indicated in Figure 8f. 29 Gravity-slide deposits, when interstratified with sedimentary rocks, point away from paleo-highs on unconformity surfaces (Sharp 1940). Landslides encircling Ceti, Coprates, and Juventae Mensae, Gale's mound, and possibly Melas Mensa, flowed away from mound crests and are overlain by sedimentary rocks, especially at locally high elevations (Lucchitta 1990, Neuffer & Schultz 2006, Okubo 2014) (Figs. 13-14). (A moatward-draining canyon at Gale's mound is also draped by sedimentary rocks; Fig. 12). Therefore, sedimentary rock emplacement on topographic highs continued after moats were defined. Therefore, draped landslides record are diagnostic for paleo-moats. Draped landslides exclude a scenario in which paleo-domes result from rapid differential compaction of initially-horizontal layers, because that scenario does not permit dome-shaped syndepositional paleotopography. To the contrary, draped landslides suggest that the paleo-dome unconformities had an erosional origin. 5. Assessment of mound emplacement hypotheses, emphasizing Valles Marineris. The VM mound-emplacement hypotheses that are most frequently discussed are the compensational-stacking (playa/lake or fluviodeltaic infill) and anticompensational stacking (e.g. slope winds) models (Section 1). Other VM ILD formation models (Nedell et al. 1987, Lucchitta 1992) include nunutaks (Gourronc et al. 2014), tuyas (Chapman & Tanaka 2001), spring mounds (Rossi et al. 2008), salt- sheet outliers (Montgomery et al. 2009), salt tectonics (Jackson et al. 1991, 2011; Baioni 2013), and carbonate deposits (McKay & Nedell 1988) (Fig. 15). The tuya and carbonate mound hypotheses fail to match post-2004 spectroscopic data. We cannot logically exclude a scenario in which the VM mounds are volcaniclastic/ash/pyroclastic deposits emplaced on the flanks of a dyke or a central volcano. However, this possibility is disfavored by (i) the tendency of fissure eruptions to evolve to pipe-eruptions geologically quickly (Wylie et al. 1999), in contrast with the shapes of the VM mounds; (ii) the regular layering of the rhythmite, suggesting quasi- periodic deposition as opposed to the power-law behavior exhibited by volcanic eruptions (Lewis et al. 2008, Pyle 1998). The salt-sheet outliers hypothesis invokes a laterally continuous salt layer (extending under the plateaus encircling VM). This hypothesized layer is hard to reconcile with VM wallrock observations that do not show salt layers, or that show salt layers which drape onto wallrock. The nunataks hypothesis invokes wet-based glaciers for which there is little uncontested evidence. The spring mounds hypothesis has difficulty explaining the great lateral continuity of observed layers. We cannot exclude the possibility that the VM mounds are giant salt domes. However, salt movement (Jackson et al. 1991, Jackson et al. 2011) after moat formation would be sideways, not upwards (as a salt glacier). Salt diapirism before moat erosion would not lead to systematic outward dips in outcrop. Where diapirism is inferred on Mars, it has a horizontal length scale that is comparable to the thickness of the sedimentary layer, and so much less than the ~102 km length of the VM mounds (Bernhardt et al. 2016). Faulting can and does tilt layers (Lewis & Aharonson 2014), but syndepositional basement uplifts beneath (and only beneath) mounds are unlikely. In particular, we disagree with the syndepositional-tilting proposal of Fueten et al. (2008) because the upper materials – the "caprock" and "rimrock" of Lucchitta (2015) – lack obvious major faults. Predepositional tectonic uplifts might nucleate draping deposition, but 30 draping deposition on highs is an example of anticompensational stacking, not an alternative. Landslides are common, but are easy to identify and are excluded from our layer orientation measurements (Fig. 11). Differential compaction of sedimentary layers over basement relief has been proposed to reconcile deposition of flat-lying strata with observed layer-orientations (Grotzinger et al. 2015). This is inconsistent with VM data. In VM, the basement surface of canyon floors that are not covered by thick sedimentary deposits is observed to be flat (e.g., E Coprates Chasma, Ganges Chasma, Noctis Labyrinthis). The implication that the central Valles Marineris canyons formed via near-vertical tectonic subsidence is strongly supported by independent tectonics data and modeling (Andrews-Hanna 2012b). If the basement of the VM canyon floors is flat before sedimentary loading, then after flexural subsidence the basement will dip inwards. This inward dip should set the sign of differential compaction tilts for initially flat-lying strata with uniform grain-size. If grain-size is not uniform, then differential compaction can tilt layers away from coarse-grained deposits. However, fluviolacustrine processes will preferentially deposit coarse grains near the margins of the canyon, again leading to inwards dips. Therefore, if differential compaction caused layer-dips in VM, we should not see outward dips on both sides of a mound. However, we do observe outward dips on both sides of mounds (Fig. 6, Fig. 8), contradicting the hypothesis of differential compaction for layer-orientations in VM mounds. A hybrid hypothesis could reconcile VM layer orientations with initially horizontal deposition. In this hypothesis, early-deposited sediments were first pre-compacted by thick overburden, and then eroded into wedge-shaped outliers. A later generation of sediments was differentially compacted over these wedge-shaped outliers, leading to the observed outward-dipping layers (Fig. 16). This hypothesis is a hybrid because terrain-influenced winds are required to erode the pre-compacted sediments into a correctly-shaped wedge prior to further sediment deposition. Even if a buried wedge of ancient sediments exists and was pre-compacted sufficiently to act as a rigid floor for later differential compaction, compaction is at best marginally sufficient to explain the large amplitude of observed dips (Gabasova & Kite 2016). Furthermore, if hypothetical wedge-shaped remnant deposits exist, then they exist mainly in subcrop, because inspection of HiRISE images does not show the large-scale within-mound onlap predicted by this scenario. Other mechanisms that rotate layers outwards during mound construction are quantitatively insufficient to explain the data, require fine-tuning, or both. For example, flexural tilting due to late-stage volcanism is <0.1° (Isherwood et al. 2013); outward tilting by the flexural response to erosional unroofing is <0.2° (Davis 2007) and can only partly recover inward tilting during sedimentary rock loading; and post-Noachian crustal-flow is minor (Karimi et al. 2016). These considerations favor the interpretation that the dip directions are primary, i.e. that the mounds grew as mounds, and that present-day mound crests are close to the crests of the growing mounds (Anderson & Bell 2010) (Figs. 2-3). In combination with the paleo-dome and draped-landslide evidence, the layer-dips suggest anticompensational stacking. One mechanism that predicts anticompensational stacking is slope-wind intensification of erosion on steep topographic slopes (Kite et al. 2013a, Day & Kocurek 2016). In this model, terrain-induced winds inhibit sedimentary rock emplacement on crater/canyon walls, creating paleo-moats. These paleo-moats serve as the basal surface for subsequent deposition. Slope-wind controlled sedimentary-basin build-up combines processes that individually have a well- 31 understood terrestrial analog but which rarely occur in combination on Earth. For example, katabatic winds drain the Antarctic plateau (Parish & Bromwich 1991); deep incision into rock by wind erosion has been reported from the Atacama (Perkins et al. 2015); and the Qaidam basin is being exhumed by wind erosion (Heermance et al. 2013). Slope-wind dynamics are not the only means of producing anticompensational-stacking kinematics: snow destabilization by föhn winds (Brothers et al. 2013), reduced saltation- transport to higher elevations due to lower pressure, and greater availability of abrasive sand at lower elevations, could all cause preferential net erosion of sediments at lower elevations – and thus favor anticompensational stacking. Sediment can be delivered by suspension transport ("airfall") and also by saltation transport. Saltation transport to paleohighs need not be prevented by moats; sand dunes flow uphill in modern VM and on polar mounds (Chojnacki et al. 2010, Conway et al. 2012; another example is visible in ESP_029504_1745). These slope-dependent models have the common advantage that they all predict that outward- directed dips should be ubiquitous, provided that craters/canyons have long, steep walls (Figure DR2 in Kite et al. 2013a). This matches our observations – outward dips are very common (section 3). Fig. 15. (Modified from Kite et al. 2013a). Comparison of the layer orientations predicted by different mound growth hypotheses, for an idealized cross section of a mound-bearing crater. Inverted triangle marks past water table. 32 Fig. 16. Summaries of 'prevailing view,' 'hybrid scenario,' and 'preferred view' of mound formation. 6. Model: anticompensational stacking and climate change. Anticompensational stacking implies that layers steepen over time. Steepening could occur via layer truncation at unconformities, mound-scale layer pinch-out, or both. We did not find evidence for mound-scale layer pinch-out. Instead, we found layer truncation at a small number of large unconformities. We interpret these unconformities as paleomoat bounding surfaces (Figs. 8-9, Table 3) (Okubo et al. 2014). We infer that anticompensational stacking arises from long depositional intervals separated by major erosive intervals (Fig. 16) – a drape-and-scrape cycle. Wind erosion can form paleomoats. Wind-induced saltation-abrasion is widely accepted to erode mound material in the present epoch (Grotzinger 2014), to have formed present-day moats (Kite et al. 2013a, Day & Kocurek 2016), and to have had greater erosive power in Mars' past. To parameterize moat and paleomoat formation, we used Mars Regional Atmospheric Modeling System (MRAMS) simulations (Rafkin et al. 2001), a realistic day-night cycle, and idealized mound-and-moat topography (Appendix B). MRAMS results indicate that slope effects are crucial to moat formation at low atmospheric pressure, with wind stresses ~5× greater on steep slopes relative to flat floors within craters/canyons (Fig. B3). Higher wind stresses are likely correlated with faster long-term wind-erosion, because wind stress is observed to exert strong 33 control on aeolian sediment transport rates on Mars (Ayoub et al. 2014), and because aeolian sediment transport is required to provide abrasive particles for sandblasting and/or to remove debris. Mars' obliquity (φ) varies quasi-periodically at 105-yr timescales, but chaotically at longer timescales, ranging from 0°-70° (Laskar et al. 2004), with significant effects on climate. At low φ, models indicate that water is less available at the low latitude of VM and Gale (e.g. Jakosky & Carr 1985, Mischna et al. 2003, Madeleine et al. 2009, Andrews-Hanna & Lewis 2011, Mischna et al. 2013, Wordsworth et al. 2013, Kite et al. 2013b). Without water for cementation, sediment does not get preserved in the sedimentary rock record. Sedimentary rock formation is also disfavored by surface condensation of atmospheric CO2 at low-φ (Forget et al. 2013, Soto et al. 2015); atmospheric collapse suppresses aeolian-sediment supply and surface liquid water. At high φ, by contrast, water is progressively driven to lower latitudes as polar regions receive greater insolation (e.g. Mischna et al. 2013). Additionally, sediment deposition rates may be enhanced by globe-spanning storms expected at high φ (Haberle et al. 2003, Armstrong & Leovy 2005, Newman et al. 2005). These considerations suggest an important role for φ in modulating sedimentary rock build-up (Lewis et al. 2008). In the words of Metz et al. (2009), "Obliquity- driven climate […] may be a more significant factor in the development of the stratigraphic record of Mars as compared to Earth." To model mound build-up including chaotic φ forcing and paleomoat formation, we carried out >100 simulations of Mars φ history. Each simulation combines 3-Gyr long 8-planet mercury6 (Chambers 1999) simulations and an obliquity model (Armstrong et al. 2004, 2014). For each simulation, we assume sedimentary rock accumulation (assumed, for simplicity, to occur at a spatially uniform rate) competes with terrain-influenced erosion at VM and Gale when Mars' obliquity (φ) > 40°, but that erosion alone operates when φ < 40°. The critical obliquity value is somewhat arbitrary, although all low-atmospheric-pressure models predict a nonlinear increase in the abundance of surface water ice at the latitude of VM and Gale at φ > (40+5 -8)°. We do not model the between-basin variation in availability of liquid water needed for cementation; previous work shows (Andrews-Hanna & Lewis 2011, Kite et al. 2013b) that between-basin variability can match the scenario presented here. We also do not model the <105 yr-timescale cycles that are responsible for the development of the layers whose orientation we measure, because these cycles occur at much shorter timescales than the overall mound construction modeled here (Lewis & Aharonson 2014). Possible causes of layering are discussed in (e.g.) Kite et al. (2013b) and Andrews-Hanna & Lewis (2011). These simplifications ensure that the details of the model do not obscure the processes modeled by SOURED (slope-wind control of within-basin spatial variations, and nonlinear obliquity control of mound-spanning unconformities variations). To combine obliquity forcing and wind-terrain feedback, we use a 2D (horizontal-vertical) landscape evolution model. Sediment is supplied from distant sources, and eroded material is removed to distant sinks. Consistent with CTX-scale morphology, thermal inertia, and the paucity of craters on sedimentary mounds (Malin et al. 2007), we assume that sedimentary rocks are much more erodible than igneous "basement". We adjust accumulation- rate so that modeled mounds are ~3 km tall. The model produces mounds of the correct height with a mound-sediment deposition rate D≈25 µm/yr, a rate that is independently suggested by the thicknesses of orbitally-paced layers (Lewis & Aharonson 2014). Model maximum deposition- rates are similar to maximum erosion-rates. Higher wind-erosion rates 3 Gya are consistent with O(1) µm/year modern-era wind-erosion rates (Grindrod & Warner 2014, Farley et al. 2014) if the 34 supply of abrading particles is not limiting, wet-era atmospheric pressure was ~60 mbar (Catling 2009, Brain & Jakosky 1998), and sandblasting rate increases faster than linearly with atmospheric pressure (Kok et al. 2012). (a) (b) Fig. 17. Model of how Mars mound stratigraphy might encode chaotic climate change. (a) Square-wave demonstration of how obliquity forcing and slope-winds combine to explain the basin-scale stratigraphy of the largest sedimentary rock mounds on Mars. Right panel: Alternations every 300 Ma between high mean obliquity (deposition) and low mean obliquity (no deposition, erosion only). Critical obliquity shown by horizontal blue line. Left panel: Sedimentary stratigraphy shown by colored line. Black line is nonerodible container. Lines are drawn at 20 Myr intervals. Colors change every 300 Myr. The numbered properties of the model 35 output are consistent with the data: (cid:1393) First-deposited stratigraphic package has layers that gently dip away from the mound center (Fig. 6c). (cid:1394) Unconformities slope away from mound center, defining paleo-domes (Fig. 9-10). (cid:1395) Unconformities steepen up-mound. (cid:1396) Dips steepen up- mound (Fig. 7). (cid:1397) Unconformity-bounded stratigraphic packages thin moving up-mound (Table 3). (b) Right panel: One possible history of orbital forcing. Horizontal line shows critical obliquity above which sedimentary rock emplacement is permitted. Left panel: Stratigraphy of mound formation for this orbital forcing (late-stage erosion is not shown). Colored lines show stratigraphy. Black line is nonerodible container. Model output (Figs. 17-18) shows that aeolian sedimentary rock emplacement forced by chaotic φ change, and including the wind-terrain feedback effect, can produce free-standing mounds within a crater/canyon (Kite et al. 2013a). The basic implications of obliquity forcing for slope winds are illustrated in Fig. 17. Fig 17a uses square-wave deposition forcing, and Fig. 17b uses an example realistic forcing. In reality, obliquity is chaotic; many simulations are needed to bracket the range of possible behavior (e.g. Fig. 18). As expected, the dominant behavior is anti- compensational stacking. A key attribute of modeled sedimentary deposits (Figs. 17-18) is that both layers and internal unconformities dip away from mound crests, consistent with data (Fig. 6-8). Though the mound topography and pattern of outward dip directions observed within Mars' sedimentary rock mounds are the most prominent features explained by this mechanism, the predicted stratigraphy simultaneously matches a range of observed physical attributes. These include the average dip magnitudes (which cluster at the mound height:width ratio), the thinning-upwards of unconformity-bounded stratigraphic packages (Malin & Edgett 2000), the layer thicknesses, and the outward dip of unconformities (Figs. 2-3) (Banham et al. 2016). The results explain why layer orientations frequently conform to modern topographic slope (Fueten et al. 2008). Because deposition occurs by progressive draping on pre-existing mound topography, only strongly nonuniform erosion (e.g. the canyons incised into Gale's mound) can create slopes that greatly differ from layer orientations. In Mars' mounds, layers are predicted to steepen upward in the stratigraphy, as subsequent layers jacket a more-gently-dipping mound core; the opposite of the geometry encountered in mountains on Earth. Modeled dips tend to steepen up-mound. The dips of exposed layers can either steepen up-mound or remain constant, depending on the depth of late-stage erosion. Because chaotic shifts in mean obliquity are infrequent (Lissauer et al. 2012, Li & Batygin 2014), the 1-2 large unconformities observed in some mounds suggest a (discontinuous) span of liquid water ≫100 Myr long. This is consistent with the ~100 Myr lower bound estimated by rhythmic layering using only the thickness of the preserved sedimentary rock, and not accounting for unconformities (Lewis & Aharonson 2014). 36 a) b) 37 c) d) Fig. 18. Additional examples of mound stratigraphies with their corresponding obliquity forcing, chosen to illustrate a range of interesting behavior. Layers drawn every 20 Myr of simulated time, color change every 300 Myr. Notice the within-moat depositional package in (a), and the "scabbed" depositional packages on the mound flank in (c) and (d). 38 7. Discussion. 7.1. Limitations of data interpretation. Anticompensational stacking can explain most of the layer orientations of most >1 km-thick sulfate-dominated stratigraphies within deep and steep-sided craters/canyons. (The residuals might be due to gravity-driven slumping, or to deposition onto a paleo-surface that had been wind-eroded into a non-axisymmetric shape - e.g., the present topography of the mound in Nicholson crater). However, many Mars mound stratigraphies do not fall into this category. For example, sedimentary mounds in Terby crater (Ansan et al. 2011) show a complicated 3D stratal architecture that cannot be reproduced by the 2D slope-winds model used here (Wilson et al. 2007, Ansan et al. 2011). Mounds within craters in W. Arabia Terra have been argued to be outliers of a formerly more extensive deposit on the basis of geographic continuity (Bennett & Bell 2016). Geographic continuity makes predictions regarding the ice mounds encircling Mars' North Polar Layered Deposits that are known to be incorrect (Conway et al. 2012, 2013; Brothers et al. 2013; Brothers & Holt 2016); therefore, geographic continuity is inconclusive. Layer dip data are unavailable for these Arabia mounds. A second key limitation of the anticompensational-stacking interpretation is that it does not work for small deposits. For example, small catenae contain layered deposits that dip inward (e.g. Weitz & Bishop 2016), and small craters in Arabia show inward dipping layers in anaglyph (e.g. HiRISE PSP_001981_1825/PSP_0012258_1825). This proves that for small container size ((cid:1232) 100 km), anticompensational stacking is not effective. In turn, this suggests a critical length/depth scale above which slope winds are most effective (§2.5). This means that our MRAMS mesoscale results need not contradict the Day et al. (2016) Large Eddy Simulation study (which emphasizes the role of unidirectional winds), but could simply refer to a different (> 100 km) scale of Mars crater/canyon. 7.2. Assumptions and limitations of model. The biggest uncertainty in our landscape evolution model is sediment availability. Sediment is assumed to be available for sedimentary rock emplacement during depositional intervals, and sand is also assumed to be available for sandblasting. This assumption of "sufficient" sand/dust/ash in Mars' past is motivated by modern data. Today, sand is present almost everywhere, but (except in a few places) is probably not pervasive and persistent enough to armor steeply-sloping bedrock over geologic time (Hayward et al. 2014). Present-day gross dust accumulation rates are not much less than inferred ancient sediment accumulation rates (Kinch et al. 2007, Lewis & Aharonson 2014). The rate of production of fine-grained material would be greater in the past because the rates of volcanism, impacts, physical erosion, and chemical weathering were all greater in the past (e.g. Golombek et al. 2006, Levy et al. 2016, Carter et al. 2013). This motivates the assumption that over long timescales, sediment is not limiting. On shorter timescales that are not resolved by the landscape evolution model, peaks in both erosion and deposition will probably be tied to the passage of supplies of abundant sand (so sediment starvation might control mound build-up at <1 km stratigraphic scales). Even if sediment is available, it will not stay in one place for Gyr unless liquid water is available to indurate it. Because our data indicate that mound build-up continued after a topographic moat was defined, regional groundwater flow is implausible as a water source for those upper layers and so the water needed for cementation must be from a top-down water source such as rain or snowmelt (Clow 1987, Niles et al. 2009, Kite et al. 2013b, Fairén et al. 2014). 39 The most important assumption in the wind erosion model is that output from 6 mbar simulations is relevant to the times when most erosion (and sedimentation) occurred, when the atmospheric pressure was likely higher (e.g. Catling et al. 2009). Since the absolute erosion rate is nondimensionalized in our model, only the pattern of wind erosion matters. Strong slope winds are expected on long steep slopes provided that the atmosphere is thin enough to permit large day-night swings in temperature (Zardi & Whiteman 2013). Thus we expect that terrain strongly influenced wind-erosion patterns in Mars' past. Patches of layered deposits veneer the slopes of some of the VM canyons (e.g. Fueten et al. 2010, 2011). Pasted-on wall-slope deposits can form in our model, but tend to be removed by late-stage erosion. The observed persistence of these outliers highlights the limitations of our 2D modeling approach. To investigate these outliers would require a fully coupled 3D model of landscape-wind coevolution. Although anticompensational stacking is the dominant behavior in our model, we did find cases where the slope winds model places lenses of sedimentary rock low down on the mound or in the moat (e.g. Fig. 18a). These packages correspond to late-stage materials that are on close-to- modern topography. Possible real-world examples are (1) young materials in the moat SW of Ceti Mensa (Okubo 2010), (2) the light-toned yardang-forming unit towards which the Mars Science Laboratory rover is driving, (3) the Siccar Point group in Gale including the Stimson formation (Fraeman et al. 2016). In our model there is no secular climate change. This is unrealistic; secular climate change clearly occurred on Mars (Jakosky & Phillips 2001). Our calculations assume that climate change driven by chaotic alternations in mean obliquity introduces a large-amplitude overprint on secular change, and we focus on those alternations. 7.3. Geological implications and tests. Obliquity strongly influences the three limiting factors for sedimentary rock build-up on Mars: sediment supply, water supply, and erosion intensity. Orbitally-forced drape-and-scrape cycles produce a good match to observations (Fig. 17). However, alternatives to φ-modulated accumulation exist. Secular variations in sediment supply, induced for example by regionally- coordinated volcanism, could explain the unconformities. This could be tested by mapping longitudinal trends in unconformity patterns. Alternatively, volcanism might globally coordinate wet episodes via greenhouse forcing. However, volcanic greenhouse gases are either too long- lived (CO2) or too short-lived (SO2) to easily explain the modulations (Kerber et al. 2015). Ice/dust cover might intermittently shield rocks from abrasion, but latitudinal shifts of cover materials are likely to be themselves φ-paced. If the great unconformities are obliquity paced, then the time gaps at unconformities should be >100 Ma. This can be tested via counts of embedded craters. Only one time gap at a Mars unconformity has been constrained so far (Kite et al. 2015), and the time gap is found to be >100 Ma, as predicted. Latitudinal variations offer clues to mound origin. The biggest sedimentary mounds on Mars lie near the equator. These mounds have few obvious mound-spanning angular unconformities. By contrast, mounds poleward of ±25° (e.g. Galle, Terby) show numerous unconformities. This is expected for deposits forming at the margins of the latitudinal belt that permitted sedimentary rock formation (Kite et al. 2013b). The variation in mound height between canyons (the thickest 40 deposits are in Northern VM, i.e. closer to Mars' Equator) could be due to a preference for sedimentary rock emplacement near the equator (Kite et al. 2013b). Alternatively, greater erosion in the canyons that now have thinner deposits might explain the latitudinal trend. Tests include measuring layer thicknesses (e.g. Lewis & Aharonson 2014, Cadieux & Kah 2015) and unconformity spacings. Obliquity-modulated build-up can be tested by the Curiosity rover's climb through sulfate- bearing layers toward the major unconformity at Gale's mound identified by Malin & Edgett (2000). An origin via chaotic shifts in mean obliquity predicts that sedimentation episodes are long and few in number. φ-control predicts long time gaps at unconformities (Fig. 3), with gently dipping layers erosionally truncated and overlain by more-steeply-dipping layers draped over preexisting stratigraphy. Detection of gravels sourced from Gale's rim within strata high in Mt. Sharp / Aeolis Mons would disprove our model. Instead, aeolian (and reworked-aeolian) deposits should dominate. Evidence for paleo-erosion by wind should be common close to unconformities. Onlap at unconformities would support the hybrid hypothesis in Fig. 16, whereas draping at unconformities would support the preferred interpretation in Fig 16. Mound formation processes are tightly linked to early-Mars runoff intermittency. Even small seasonal streams would suppress the sand migration that is required for saltation-driven erosion (Krapf 2003), and gravity-driven stream erosion would also suppress the anticompensational growth of mounds. Aeolian sediment supply can be reconciled with lakes in VM (Harrison & Chapman 2008) if climate permitted lakes for only a small percentage of years (Palucis et al. 2016, Buhler et al. 2014, Irwin et al. 2015). Wet-dry alternations during Mars' era of sedimentary-rock accumulation, including long dry periods, are predicted by our preferred scenario. Intermittent habitability is consistent with the persistence of surface olivine on Mars, and the detection in Gale mudstones of chemical markers for extreme aridity (Farley et al. 2016). Our data disfavor the long-standing hypothesis (McCauley 1978) that the VM outcrops are lake deposits, but are consistent with a lacustrine origin for outcrops below the base of the topographically defined mounds in VM and below the clay/sulfate transition at Gale (Grotzinger et al. 2015). 8. Conclusions. We introduce new data and a new model for the evolution of eight major sedimentary mounds in Valles Marineris and Gale crater. Data: • Seven out of eight mounds investigated show layer orientations that dip systematically away from the mound centerline, with median dip 5° (n = 308). • Layer-orientation data have a precision and accuracy that are sufficient for the purpose of constraining mound origin. • Stratigraphic surfaces interpreted as major mound-spanning unconformities are well-fit by a dome-shape in 6 out of 8 cases. Interpretation: • When combined, the layer orientation data, draped landslides, and our interpretation of stratigraphic surfaces interpreted as unconformities, require primary deposition of layers on outward-tilted slopes for the topmost ~1 km of the mounds. 41 • Lower in the stratigraphy, the layer orientation data are consistent with either (i) primary deposition of layers on outward-tilted slopes (Kite et al. 2013a) or (ii) a hybrid hypothesis in which slope-wind erosion sculpts pre-compacted sediments that later act as wedge- shaped indentors for differential compaction of later-deposited sediments. Model: • We present a model that combines spatially-resolved forcing (from mesoscale meteorological simulations) and time-variable forcing (realistic orbital integrations) to make quantitative predictions for the evolution of the major sedimentary basins of Mars. The meteorological simulations confirm a strong trend of increasing wind stress with topographic slope within both craters and canyons. • The model predicts that Mars mound stratigraphy emerges from a drape-and-scrape cycle. • The model simultaneously matches the following mound attributes: (i) layers dip away from mound crests; (ii) internal unconformities have a dome shape; (iii) average dip magnitudes cluster at the mound height:width ratio; (iv) unconformity-bounded stratigraphic packages thin upwards; (v) layer orientations frequently conform to modern topographic slope. • We propose that major mound-spanning unconformities within Mars mountains correspond to periods of low mean obliquity (Mischna et al. 2013, Kite et al. 2015). Because chaotic shifts in mean obliquity are infrequent, the 1-2 large unconformities observed in some mounds suggest a (discontinuous) span of liquid water ≫100 Myr long. In our model, the major mound-spanning unconformities (once correctly ordinated) can be used for planetwide correlation. • On the Earth, first-order erosion-deposition alternations (Sloss 1963) are driven at a global scale by the Wilson cycle (via orogeny and eustasy). On Mars, climate changes driven by infrequent chaotic shifts in mean obliquity may play an analogous role in shaping the planet's sedimentary record. Acknowledgements. We thank Michael Lamb, Frank Fueten, Gene Schmidt, Chris Okubo, Leila Gabasova, John Grotzinger, Mackenzie Day, Jeff Barnes, Daniel Tyler, Claire Newman, Mark Richardson, John Armstrong, Bill Dietrich, Jasper Kok, Corey Fortezzo, Baerbel Lucchitta, Nathan Bridges, Brad Thomson, and David Rowley, for enlightening comments and discussions, and for sharing unpublished data. We thank four reviewers for timely, thorough, and thoroughly useful reviews. We thank the HiRISE team for maintaining the HiWish program, which provided multiple images that were valuable for this work. We thank UChicago's Research Computing Center. This work was financially supported by the U.S. taxpayer via NASA grant NNX15AH98G. Layer-orientation data are available as a supplementary table. DTM scripts are written in Bash https://psd-repo.uchicago.edu/kite- and lab/uchicago_asp_scripts/. DTMs and other data produced for this study may be obtained for unrestricted further use by contacting the lead author ([email protected]). download from are available for 42 Appendix A. Stratigraphic Model. A.1. Overview and physical basis of stratigraphic model. The central element of our SOURED (Fig. A1) is the forward model of landscape evolution and stratigraphy (section A.2), which incorporates time-varying sedimentary rock emplacement (assumed uniform within craters/canyons for simplicity) and spatially-varying feedback from slope-winds. Time-varying sedimentary rock emplacement is forced by 3-Gyr long integrations of the orbit and spin-pole orientation of Mars (section A.3). Spatially-varying feedback from slope-winds is forced by a mesoscale wind model (Appendix B). Essentially, SOURED = an upgraded version of SWEET + MRAMS + (mercury6 + oblique) (Fig. A1). Slope winds are important on Mars. These diurnally-reversing winds result from the combination of high relief and day-night temperature swings of up to 130K (e.g. Kass et al. 2003). Slope winds are particularly strong within the equatorial craters and canyons that host sedimentary rock mounds, where Coriolis effects are weak and relief can approach 10 km. The coupling between long, steep slopes and strong winds on Mars emerges from basic physical principles and is model-independent (Spiga et al. 2011, Kite et al. 2013a, Zardi & Whiteman 2013, Moreau et al. 2014, Tyler & Barnes 2015, Rafkin et al. 2016). Fig. A1. Sketch of SOURED model. Combining the MRAMS output with the obliquity forcing, we use the stratigraphic forward model to predict the structure of the mounds. A.2. Stratigraphic forward model. The purpose of our forward stratigraphic model is generate basin stratigraphies for comparison with observations. Earth models with the same purpose (but different physics) include SedSim (Griffiths et al. 2001) and Dionisos (Csato et al. 2014). Our forward stratigraphic model is 2D (one horizontal dimension and one vertical dimension), with a nominal resolution of ~1 km in the horizontal dimension and 1 Myr in time. Our model does not attempt to resolve processes operating at shorter scales of space and/or time. 43 The model is modified after the Slope-Wind Enhanced Erosion and Transport (SWEET) model of Kite et al. (2013a), with significant enhancements to incorporate parameterized erosion estimators obtained from mesoscale models and time-varying climate forcing (Fig. A1). In SWEET, (1) where D is deposition rate and eE is erosion rate, (2) where kE is an erodibility parameter, U is wind shear-stress, and β is in the range 2-4 for wind- erosion processes (Kok et al. 2012). The threshold for sediment mobilization is omitted, which is a large simplification. Since the gap between the fluid threshold for saltation initiation and the impact threshold for saltation cessation is so large on Mars, the threshold is very uncertain. Large values of β produce a similar pattern of normalized wind erosion to large values of the mobilization threshold. Therefore, combining the threshold with β is a reasonable simplification. In earlier work (Kite et al. 2013a) we treated the relative importance of slope winds (Us) and background or "synoptic" winds U0 as a free parameter, (3) where U0 could be varied. Here we remove the free parameter U0 by calculating erosion estimators directly from cell-by-cell mesoscale model output, (4) where Ns is the number of grid cells with slopes in the range of interest, t is total elapsed time, tsu is spin-up time (t - tsu is always an integer number of sols), ∆t is timestep, τs is the instantaneous surface shear stress (in Pa), and β is from equation (2). In practice we use a log-linear fit to the cell-by-cell data to get a smooth relationship between slope and erosion (Appendix B). kE is adjusted to match the height of observed mounds. In the limit where erosion depends only on local slope (modeled here), and where eE ~ 0 for s = 0, the model will tend to produce a cone (or triangular prism) of sedimentary rocks whose side-slope is dz/dx = (D/kE)(1/β). SWEET does not conserve mass locally. Instead, material is added from distant sources (e.g. by airfall), and eroded material is removed to a distant sink (e.g. the Martian lowlands; Grotzinger & Milliken 2012). Layers in the model are assumed to be indurated (mobile sand is assumed to be topographically superficial or to have a geologically short residence time). Because induration probably involves cementation by mineral precipitation from aqueous fluids, long-term secular decline in Mars' ability to form sedimentary rocks (due to, for example, water loss and CO2 loss) means that the model is most applicable to Early Mars. 44 A.3. Orbital dynamics model and obliquity model. The purpose of our orbital dynamics model and obliquity model is to generate an ensemble of realistic 3.1-Gyr-long obliquity tracks for Mars (Kite et al. 2015). We generated φ tracks using mercury6 (the N-body code of Chambers 1999), and the obliquity code of Armstrong et al. (2004, 2014). For each >3 Gyr long 8-planet solar system integration (n = 37) (the combined eccentricity pdf from these integrations is very similar to that of Laskar et al. 2004), we seeded 24 Mars φ-tracks drawing the initial φ from the long term distribution of Laskar et al. 2004. From the ensemble, we selected those φ-tracks which ended (after 3.1 Gyr) in the range 20°-35° (consistent with present-day Mars φ). The figures in this paper show a subset of the stratigraphic output forced by those φ-tracks, chosen to illustrate a range of common stratigraphic outcomes. Appendix B. Mesoscale model. B.1. Mesoscale model input. The purpose of our mesoscale modeling work is to verify that wind stress increases with topographic slope. We also seek the 'slope enhancement factor' – to what extent is erosion rate (assumed to scale as wind stress to some power β) faster on steep slopes than on flat slopes within craters/canyons? We have already verified (Kite et al. 2013a), using the MarsWRF model (Toigo et al. 2012, Richardson et al. 2007), that the strongest winds are on the steepest slopes for a simulation of one year's winds at Gale crater. Here we use the Mars Regional Atmospheric Modeling System (Rafkin et al. 2001) to extend our earlier results through exploring a range of idealized topographies (Tyler & Barnes 2015, Day et al. 2016). MRAMS is derived from the terrestrial RAMS model (Mahre & Pielke 1976). MRAMS has been used to model the entry and descent of all NASA Mars landers subsequent to Pathfinder (Michaels & Rafkin 2008). We use a horizontal resolution of 4.4 km, and a vertical resolution varying from 15 m near the surface to >1 km at high altitude. A realistic diurnal cycle in insolation is imposed (including planetary- scale thermal tides). Our runs are carried out at 6 mbar; the pattern of wind forcing should be similar for other atmospheres that are thin enough for a large day-night cycle in surface temperature. The orbital parameters are for modern Mars, but the diurnally-reversing mesoscale circulation should operate similarly at high obliquity (the background winds may be stronger; Haberle et al. 2003, Newman et al. 2005). Boundary conditions are supplied by the NASA Ames Mars General Circulation Model (MGCM) (Haberle et al. 1993). For this project, we modify MRAMS to simulate idealized craters and idealized canyons. We use smoothed background topography, and insert oblong canyons of width ~130 km and length ~350 km and depth ~4.5 km. We run the model both without mounds, and for canyons containing mounds of 100% of the full height of the canyon (Fig. B1). These runs correspond to idealized topography for a large canyon hosting a large mound (e.g. Candor, Hebes, Ophir). Separately, we insert 4.5 km-deep, 155 km-diameter axisymmetric craters (Fig. B1), with their corresponding mounds. These runs correspond to idealized topography for a large crater hosting a large mound (e.g. Gale crater, Nicholson crater). We ran for 5.7 day-night cycles for solar longitude Ls = {30°, 90°, 150°, 180, 210°, 270°, 330°}. The first 1.7 sols are discarded as spinup. Simulated crater/canyon latitude is ~5°S. 45 a) c) b) d) Fig. B1. Topographies investigated using MRAMS simulations. Contours at 500m intervals. a) Rectangular canyon, width ~130 km and length ~350 km, with full-height mound. b) Crater, 155 km diameter, with full-height mound. c) Flat-floored canyon without sediment infill. d). Flat- floored crater without sediment infill. B.2. Mesoscale model output. Our MRAMS runs confirm that the strongest winds within craters/canyons are associated with diurnally-reversing (anabatic/katabatic) flows, and are located on the steepest slopes (Fig. B2). Terrain-controlled circulation dominates the overall circulation inside our idealized craters and canyons (consistent with Tyler & Barnes 2015) (Fig. B3). The importance of slope winds in our idealized-topography runs is somewhat offset for real craters and canyons by regional effects (e.g. the planetary topographic dichotomy boundary, Rafkin et al. 2016). To simplify the analysis, we assume cell-scale (4km-scale) control of terrain on wind stress. Gridcells inside a canyon or crater are generally less windy than on the plateau surrounding the depression (Fig. B3). This is partly because the plateau is subject to the morning "surge" of air moving away from the canyon (Tyler & Barnes 2015). However, within the crater/canyon, wind stress is about 5× greater for 15° slopes than for flat surfaces. Points just below the rim of the crater/canyon have stronger wind stress than expected for their slope, because they participate in the morning "surge" of air moving away from the crater/canyon. The scatter of mean wind stress is about a factor of 2. We use the crater output; the same trends were found for canyons as for craters. 46 How we get from mesoscale model output to erosion estimators: Even if our wind models perfectly represented wind stresses inside Mars craters/canyons, that would not be enough to correctly diagnose the rate of aeolian erosion of bedrock. Aeolian erosion of rock is a multi-step process (Shao et al. 2008, Kok et al. 2012), and it is difficult to determine the rate-limiting step from orbit. Possibilities include breakdown of sedimentary layers to wind-transportable fragments by weathering and/or volume changes associated with hydration state changes (e.g., Chipera & Vaniman 2007); physical degradation by mass wasting, combined with aeolian removal of talus (Kok et al. 2012); aeolian erosion of weakly salt-cemented sediments (Shao, 2008); and aeolian abrasion of bedrock (Wang et al., 2011). Rather than attempting to directly predict erosion rate, we use the strong evidence for geologically-recent wind erosion of the mounds (e.g. Day et al. 2016) to establish the feasibility of aeolian sculpting of the mounds, and we use the wind models to get the pattern of past wind erosion. This requires us to accept two limitations: 1. The relationship between wind stress and erosion rate will vary depending on both past atmospheric pressure and the details of the erosion process. We parameterize this uncertainty by a power-law exponent, β. 2. Wind erosion is not carried out by the wind directly, but by sand grains carried by the wind (at least for most erosion processes), and sand grains are not tracked by the model This is directly analogous to the tools-and-cover problem in modeling the evolution of Earth's mountains (section 7). To get the relationship between wind stress and slope, we tried fitting the unbinned data with various functions (exponential, two-exponential, power law, polynomial, e.t.c). The most visually satisfying fit is a log-linear function. The fit suffices to capture the basic tendency for wind erosion within craters and canyons to be stronger on steep slopes than on gentle slopes, by a factor of between ~4 (if erosion is proportional to mean wind stress) and >10 (if erosion is proportional to wind stress raised to the fourth power). We defined erosion estimators from the MRAMS model output on a per-gridcell basis as follows (5a) (5b) We obtained eE by regression using a log-linear fit where eE = 10( k1 s + k2), where s is slope. We did this for the 100%-mound simulation (mound-in-crater) (Fig. B3). The erosion estimators are (for β =1, i.e. erosion proportional to mean wind speed) k1 = 3.08(2.84,3.31), k2 = -3.37(-3.41,- 3.33), (for β =2) k1 = 5.34(4.87,5.82), k2 = -6.34(-6.42,-6.26), (for β =3) k1 = 7.44(6.69,818), k2 = -9.13(-9.26,-9.01) and (for β =4) k1=9.42(8.39,10.44), k2 = -11.81(-11.98,-11.64). Here, the brackets give the formal confidence interval of the fits. The choice of erosion estimator depends on the paleoatmospheric pressure and on the mechanism of erosion. For low atmospheric pressure, u*cr (the surface-stress threshold for sand motion) approaches the maximum wind speed, and β à ∞ (i.e. erosion only responds to the very strongest gusts). Sand dunes on Mars today are in active motion (Bridges et al. 2012a), so β < ∞. When atmospheric pressure was 47 higher earlier in Mars history, u*cr would become small compared to frequently-encountered wind speeds. Under those circumstances, 2 < β < 4 is appropriate. We used β = 3 to make the plots shown in this paper. We carried out sensitivity tests changing the β parameter, finding no qualitative difference for 2 < β < 4 (after adjustment for each β of the dimensionless deposition rate in order to match observed mound heights). The past terrain-averaged erosion rate is effectively a free parameter in our model. We find good results with maximum past rates that are comparable to Earth wind erosion rates, that agree with previous calculations of peak present-day Mars wind erosion rates (Bridges et al. 2012b), and that are 1 order of magnitude greater than typical present-day Mars sedimentary rock wind erosion rates (Golombek et al. 2014, Kite & Mayer submitted), consistent with higher atmospheric pressure (or weaker rocks) in the past. 48 a) b) d) c) Fig. B2. Winds are strong on crater/canyon walls and on mound flanks, but weak in the moat. (a) Snapshot of daytime flow in idealized-topography crater, dominated by upslope (anabatic) winds. Contours (m) are elevation. Colors are wind speed (m/s) at 15m elevation. Topographic contour interval 500m. (b) Snapshot of nighttime flow in idealized-topography crater, dominated by downslope (katabatic) winds. Topographic contour interval 500m. (c) Snapshot of daytime flow in idealized-topography canyon, dominated by upslope (anabatic) winds. Topographic contour interval 1000m. (d) Snapshot of nighttime flow in idealized-topography canyon, dominated by downslope (katabatic) winds. Topographic contour interval 1000m. (This figure was produced using the Grid Analysis and Display System, GrADS: http://cola.gmu.edu/grads/). 49 (a) (c) (b) Fig. B3. MRAMS run-integrated output. (a) Maximum wind stress from the last 4 sols of model runs, for topographic boundary conditions featuring large central mounds. Model output is binned according to grid cell slope, and the median for each bin is shown (stars). Colors correspond to distance from mound center in kilometers. The black line (and gray error bars) corresponds to the best log-linear fit to all data. The red line (and red error bars) correspond to the best log-linear fit to the binned data (asterisk symbols). (b) Mean wind stress from the last 4 sols of model runs at different seasons, for topographic boundary conditions corresponding to a full-height central mound inside a crater. The numbers in the legend correspond to the Ls (Martian season) of each run. (c) The overall best fits to the slope-effect multiplier (offset for clarity; for slope = 0, the slope-multiplier effect is 1 by definition). 50 References. Allen, C. C.; Dapremont, A. M.; Oehler, D. Z. (2014), The Complex, Multi-Stage History of Mt. Sharp, 45th Lunar and Planetary Science Conference, The Woodlands, Texas. LPI Contribution No. 1777, p.1402 Allen, P.A. & J.R. Allen (2013), Basin Analysis, 3rd edition, Wiley-Blackwell. Anderson, Ryan B.; Bell, James F., III (2010), Geologic mapping and characterization of Gale Crater and implications for its potential as a Mars Science Laboratory landing site, International Journal of Mars Science and Exploration, vol. 4, p.76-128. Anderson, Robert C.; Dohm, James M.; Golombek, Matthew P.; Haldemann, Albert F. C.; Franklin, Brenda J.; Tanaka, Kenneth L.; Lias, Juan; Peer, Brian (2001), Primary centers and secondary concentrations of tectonic activity through time in the western hemisphere of Mars, J. Geophys. Res., 106, E9, p. 20563-20586. Andrews-Hanna, J.C.; Zuber, M.T.; Arvidson, R.E.; Wiseman, S.M. (2010), Early Mars hydrology: Meridiani playa deposits and the sedimentary record of Arabia Terra, J. Geophys. Res., 115, E06002 . Andrews-Hanna, J.C.; Lewis, K.W. (2011), Early Mars hydrology: 2. Hydrological evolution in the Noachian and Hesperian epochs, J. Geophys. Res., 116, E02007. Andrews-Hanna, Jeffrey C. (2012a), The formation of Valles Marineris: 3. Trough formation through super- isostasy, stress, sedimentation, and subsidence, J. Geophys. Res., 117, E6, CiteID E06002 Andrews-Hanna, Jeffrey C. (2012b), The formation of Valles Marineris: 1. Tectonic architecture and the relative roles of extension and subsidence, J. Geophys. Res., 117, E3, doi:10.1029/2011JE003953. Ansan, V.; Loizeau, D.; Mangold, N.; Le Mouélic, S.; Carter, J.; Poulet, F.; Dromart, G.; Lucas, A.; Bibring, J.-P.; Gendrin, A.; Gondet, B.; Langevin, Y.; Masson, Ph.; Murchie, S.; Mustard, J. F.; Neukum, G. (2011), Stratigraphy, mineralogy, and origin of layered deposits inside Terby crater, Mars, Icarus 211, 273-304. Armstrong, John C.; Leovy, Conway B.; Quinn, Thomas (2004), A 1 Gyr climate model for Mars: new orbital statistics and the importance of seasonally resolved polar processes, Icarus, 171, 255-271. Armstrong, John C.; Leovy, Conway B. (2005), Long term wind erosion on Mars, Icarus, 176, 1, 57-74. Armstrong, John C., et al. (2014), Effects of extreme obliquity variations on the habitability of exoplanets, Astrobiology, 14, doi:10.1089/ast.2013.1129. Arvidson, R. E.; Poulet, F.; Morris, R. V.; Bibring, J.-P.; Bell, J. F.; Squyres, S. W.; Christensen, P. R.; et al. (2006), Nature and origin of the hematite-bearing plains of Terra Meridiani based on analyses of orbital and Mars Exploration rover data sets, J. Geophys. Res., 111, E12, CiteID E12S08. Arvidson, R. E.; Bell, J. F., III; Catalano, J. G.; Clark, B. C.; Fox, V. K.; Gellert, R.; Grotzinger, J. P.; Guinness, E. A.; Herkenhoff, K. E.; Knoll, A. H.; Lapotre, M. G. A.; McLennan, S. M.; Ming, D. W.; Morris, R. V.; Murchie, S. L.; Powell, K. E.; Smith, M. D.; Squyres, S. W.; Wolff, M. J.; Wray, J. J. (2015), Mars Reconnaissance Orbiter and Opportunity observations of the Burns formation: Crater hopping at Meridiani Planum, J. Geophys. Res.: Planets, 120, 3, 429-451. Ayoub, F., Avouac, J.-P., Newman, C.E., Richardson, M.I., Lucas, A., Leprince, S., Bridges, N.T. 2014. Threshold for sand mobility on Mars calibrated from seasonal variations of sand flux. Nature Communications 5, 5096. Baioni, Davide (2013), Morphology and geology of an interior layered deposit in the western Tithonium Chasma, Mars, Planetary and Space Science, 89, 140-150. Banham, S. G.; Gupta, S.; Rubin, D. M.; Watkins, J. A.; Sumner, D. Y.; Grotzinger, J. P.; Lewis, K. W.; Edgett, K. S.; Edgar, L. A.; Stack, K. M. (2016), Reconstruction of an Ancient Eolian Dune Field at Gale Crater, Mars: Sedimentary Analysis of the Stimson Formation, 47th Lunar and Planetary Science Conference, held March 21-25, 2016 at The Woodlands, Texas. LPI Contribution No. 1903, 2346 Bennett, K.A., Bell, J.F. 2016. A global survey of martian central mounds: Central mounds as remnants of previously more extensive large-scale sedimentary deposits. Icarus 264, 331-341. Bernhardt, H.; Reiss, D.; Hiesinger, H.; Ivanov, M. A. (2016), The honeycomb terrain on the Hellas basin floor, Mars: A case for salt or ice diapirism, J. Geophys. Res.: Planets, 121, 4, 714-738. Beyer, R. A.; Alexandrov, O.; Moratto, Z. M. (2014), Aligning Terrain Model and Laser Altimeter Point Clouds with the Ames Stereo Pipeline, 45th Lunar and Planetary Science Conference, LPI Contribution No. 1777, 2902. Bibring, Jean-Pierre; Langevin, Yves; Mustard, John F.; Poulet, François; Arvidson, Raymond; Gendrin, Aline; et al. (2006), Global Mineralogical and Aqueous Mars History Derived from OMEGA/Mars Express Data, Science, 312, 5772, 400-404 (2006). 51 Bibring, J.-P.; Arvidson, R. E.; Gendrin, A.; Gondet, B.; Langevin, Y.; Le Mouelic, S.; Mangold, N.; Morris, R. V.; Mustard, J. F.; Poulet, F.; Quantin, C.; Sotin, C. (2007), Coupled Ferric Oxides and Sulfates on the Martian Surface, Science, 317, 5842, 1206-. Bishop, Janice L.; Parente, Mario; Weitz, Catherine M.; Noe Dobrea, Eldar Z.; Roach, Leah H.; Murchie, Scott L.; McGuire, Patrick C.; McKeown, Nancy K.; Rossi, Christopher M.; Brown, Adrian J.; Calvin, Wendy M.; Milliken, Ralph; Mustard, John F. (2009), Mineralogy of Juventae Chasma: Sulfates in the light-toned mounds, mafic minerals in the bedrock, and hydrated silica and hydroxylated ferric sulfate on the plateau, J. Geophys. Res., 114, 1, CiteID E00D09. Borlina, Cauê S.; Ehlmann, Bethany L.; Kite, Edwin S. (2015), Modeling the thermal and physical evolution of Mount Sharp's sedimentary rocks, Gale Crater, Mars: Implications for diagenesis on the MSL Curiosity rover traverse, J. Geophys. Res.: Planets, 120, 8, 1396-1414. Bradley, Bethany A.; Sakimoto, Susan E. H.; Frey, Herbert; Zimbelman, James R. (2002), Medusae Fossae Formation: New perspectives from Mars Global Surveyor, J. Geophys. Res. (Planets), 107, E8, 2-1, CiteID 5058, DOI 10.1029/2001JE001537. Brain, David A.; Jakosky, Bruce M. (1998), Atmospheric loss since the onset of the Martian geologic record: Combined role of impact erosion and sputtering, J. Geophys. Res., 103, E10, 22689-22694. Brain, D. A.; McFadden, J. P.; Halekas, J. S.; Connerney, J. E. P.; Bougher, S. W.; Curry, S.; Dong, C. F.; Dong, Y.; Eparvier, F.; Fang, X, et al. (2015), The spatial distribution of planetary ion fluxes near Mars observed by MAVEN, Geophys. Res. Lett., 42, 21, 9142-9148 Bridges, N. T.; Bourke, M. C.; Geissler, P. E.; Banks, M. E.; Colon, C.; Diniega, S.; Golombek, M. P.; Hansen, C. J.; Mattson, S.; McEwen, A. S.; Mellon, M. T.; Stantzos, N.; Thomson, B. J. (2012a), Planet-wide sand motion on Mars, Geology, vol. 40, issue 1, 31-34 Bridges, N. T.; Ayoub, F.; Avouac, J.-P.; Leprince, S.; Lucas, A.; Mattson, S. (2012b), Earth-like sand fluxes on Mars, Nature, 485, 7398, 339-342. Bridges, Nathan; Geissler, Paul; Silvestro, Simone; Banks, Maria (2013), Bedform migration on Mars: Current results and future plans, Aeolian Research, 9, 133-151. Brothers, T. C.; Holt, J. W.; Spiga, A. (2013), Orbital radar, imagery, and atmospheric modeling reveal an aeolian origin for Abalos Mensa, Mars, Geophys. Res. Lett., 40, 1334-1339. Brothers, T.C., and J.W. Holt (2016), 3-dimensional structure and origin of a 1.8-km-thick ice dome within Korolev Crater, Mars, Geophys. Res. Lett. 43, 1443–1449 Buhler, Peter B.; Fassett, Caleb I.; Head, James W.; Lamb, Michael P. (2014), Timescales of fluvial activity and intermittency in Milna Crater, Mars, Icarus, 241, 130-147. Burr, Devon M.; Enga, Marie-Therese; Williams, Rebecca M. E.; Zimbelman, James R.; Howard, Alan D.; Brennand, Tracy A. (2009), Pervasive aqueous paleoflow features in the Aeolis/Zephyria Plana region, Mars, Icarus, 200, 52-76. Carter, Lynn M.; Campbell, Bruce A.; Watters, Thomas R.; Phillips, Roger J.; Putzig, Nathaniel E.; Safaeinili, Ali; Plaut, Jeffrey J.; Okubo, Chris H.; Egan, Anthony F.; Seu, Roberto; Biccari, Daniela; Orosei, Roberto (2009), Shallow radar (SHARAD) sounding observations of the Medusae Fossae Formation, Mars, Icarus, 199, 295-302. Carter, J.; Poulet, F.; Bibring, J.-P.; Mangold, N.; Murchie, S. (2013), Hydrous minerals on Mars as seen by the CRISM and OMEGA imaging spectrometers: Updated global view, J. Geophys. Res.: Planets, 118, 4, 831-858. Catling, David C.; Wood, Stephen E.; Leovy, Conway; Montgomery, David R.; Greenberg, Harvey M.; Glein, Christopher R.; Moore, Jeffrey M. (2006), Light-toned layered deposits in Juventae Chasma, Mars, Icarus, 181, 26- 51. Catling, D.C. (2009), Atmospheric Evolution of Mars. In: V. Gornitz (ed.) Encyclopedia of Paleoclimatology and Ancient Environments, Springer, Dordrecht, 66-75. Chambers, J. E. (1999), A hybrid symplectic integrator that permits close encounters between massive bodies, Monthly Notices of the Royal Astronomical Society, 304, 793-799. Cadieux, Sarah B.; Kah, Linda C. (2015), To what extent can intracrater layered deposits that lack clear sedimentary textures be used to infer depositional environments?, Icarus, 248, 526-538. Chapman, Mary G.; Tanaka, Kenneth L. (2001), Interior trough deposits on Mars: Subice volcanoes?, J. Geophys. Res., 106, E5, 10087-10100. Chipera, S.J., & Vaniman, D.T., 2007, Experimental stability of magnesium sulfate hydrates that may be present on Mars, Geochimica et Cosmochimica Acta, 71, 241-250. 52 Chojnacki, Matthew; Moersch, Jeffrey E.; Burr, Devon M. (2010), Climbing and falling dunes in Valles Marineris, Mars, Geophysical Research Letters, 37, 8, CiteID L08201. Christensen, P. R.; Morris, R. V.; Lane, M. D.; Bandfield, J. L.; Malin, M. C. (2001), Global mapping of Martian hematite mineral deposits: Remnants of water-driven processes on early Mars, J. Geophys. Res., 106, 23873-23886 Chojnacki, Matthew; Hynek, Brian M. (2008), Geological context of water-altered minerals in Valles Marineris, Mars, J. Geophys. Res., 113, CiteID E12005. Clow, G. D. (1987), Generation of liquid water on Mars through the melting of a dusty snowpack, Icarus (ISSN 0019-1035), vol. 72, Oct. 1987, 95-127. Conway, S.J.; Hovius, N., Barnie, T.; Besserer, J.; Le Mouélic, S.; Orosei, R.; Read, N. (2012), Climate-driven deposition of water ice and the formation of mounds in craters in Mars' north polar region, Icarus, 220, 174-193. Csato, I.; Catuneanu, O.; Granjeon, D. (2014), Millennial-Scale Sequence Stratigraphy: Numerical Simulation With Dionisos, Journal of Sedimentary Research, 84, 394-406. Davies, R.J., and Cartwright, J., (2002), A fossilized Opal A to Opal C/T transformation on the northeast Atlantic margin, Basin Research 14, 467-486. Davis, B.J. (2007), Erosion-driven uplift and tectonics at Valles Marineris, Mars, Colorado School of Mines, MSci thesis. Day, M., & Kocurek, G., (2016), Observations of an aeolian landscape: From surface to orbit in Gale Crater, Icarus 2016 doi:10.1016/j.icarus.2015.09.042. Day, M., Anderson, W., Kocurek, G., Mohrig D. (2016), Carving intracrater layered deposits with wind on Mars, Geophysical Research Letters, DOI: 10.1002/2016GL068011. Edgar, Lauren, et al. (2012), Stratigraphic architecture of bedrock reference section, Victoria Crater, Meridiani Planum, Mars, in Grotzinger, J., & Milliken, R., Sedimentary Geology of Mars, SEPM Special Publication 102, 195-209. Ehlmann, Bethany L.; Mustard, John F.; Murchie, Scott L.; Bibring, Jean-Pierre; Meunier, Alain; Fraeman, Abigail A.; Langevin, Yves, (2011), Subsurface water and clay mineral formation during the early history of Mars, Nature, 479, 7371, 53-60. Ehlmann, Bethany L.; Mustard, John F. (2012), An in-situ record of major environmental transitions on early Mars at Northeast Syrtis Major, Geophys. Res. Lett., 39, L11202 Fairén, Alberto G.; Davila, Alfonso F.; Gago-Duport, Luis; Haqq-Misra, Jacob D.; Gil, Carolina; McKay, Christopher P.; Kasting, James F. (2011), Cold glacial oceans would have inhibited phyllosilicate sedimentation on early Mars, Nature Geoscience, 4, 10, 667-670. Fairén, Alberto G.; Stokes, Chris R.; Davies, Neil S.; Schulze-Makuch, Dirk; Rodríguez, J. Alexis P.; et al. (2014), A cold hydrological system in Gale crater, Mars, Planetary and Space Science, Volume 93, p. 101-118. Farley, K. A, et al., (2014), In Situ Radiometric and Exposure Age Dating of the Martian Surface, Science, 343, id. 1247166 (2014). Farley, K. A.; Martin, P.; Archer, P. D.; Atreya, S. K.; Conrad, P. G.; et al. (2016), Light and variable 37Cl/35Cl ratios in rocks from Gale Crater, Mars: Possible signature of perchlorate, Earth and Planetary Science Letters, 438, 14-24. Fassett, Caleb I.; Head, James W. (2011), Sequence and timing of conditions on early Mars, Icarus, 211, 1204-1214. Fastook, James L.; Head, James W.; Marchant, David R.; Forget, Francois (2008), Tropical mound glaciers on Mars: Altitude-dependence of ice accumulation, accumulation conditions, formation times, glacier dynamics, and implications for planetary spin-axis/orbital history, Icarus, 198, 2, 305-317. Fergason, R. L.; Gaddis, L. R.; Rogers, A. D. (2014), Hematite-bearing materials surrounding Candor Mensa in Candor Chasma, Mars: Implications for hematite origin and post-emplacement modification, Icarus, 237, 350-365. Forget, F., Wordsworth, R., Millour, E., Madeleine, J.-B., Kerber, L., et al. (2013), 3D modelling of the early martian climate under a denser CO2 atmosphere: Temperatures and CO2 ice clouds. Icarus 222, 81-99. Fraeman, A. A.; Arvidson, R. E.; Catalano, J. G.; Grotzinger, J. P.; Morris, R. V.; et al. (2013), A hematite-bearing layer in Gale Crater, Mars: Mapping and implications for past aqueous conditions, Geology, 41, 1103-1106. 53 A. A. Fraeman et al. (2016), The stratigraphy and evolution of lower Mount Sharp from spectral, morphological, and thermophysical orbital data sets, J. Geophys. Res., DOI: 10.1002/2016JE00509 Fueten, F.; Stesky, R.; MacKinnon, P.; Hauber, E.; Gwinner, K.; Scholten, F.; Zegers, T.; Neukum, G. (2006), A structural study of an interior layered deposit in southwestern Candor Chasma, Valles Marineris, Mars, using high resolution stereo camera data from Mars Express, Geophys. Res. Lett., 33, CiteID L07202 Fueten, F.; Stesky, R.; MacKinnon, P.; Hauber, E.; Zegers, T.; Gwinner, K.; Scholten, F.; Neukum, G. (2008), Stratigraphy and structure of interior layered deposits in west Candor Chasma, Mars, from High Resolution Stereo Camera (HRSC) stereo imagery and derived elevations, J. Geophys. Res., 113, E10, CiteID E10008 . Fueten, F.; Racher, H.; Stesky, R.; MacKinnon, P.; Hauber, E.; McGuire, P. C.; Zegers, T.; Gwinner, K. (2010), Structural analysis of interior layered deposits in Northern Coprates Chasma, Mars, Earth and Planetary Science Letters, 294, 3-4, 343-356. Fueten, F.; Flahaut, J.; Le Deit, L.; Stesky, R.; Hauber, E.; Gwinner, K. (2011), Interior layered deposits within a perched basin, southern Coprates Chasma, Mars: Evidence for their formation, alteration, and erosion, J. Geophys. Res., 116, E2, CiteID E02003 Fueten, F.; Flahaut, J.; Stesky, R.; Hauber, E.; Rossi, A. P. (2014), Stratigraphy and mineralogy of Candor Mensa, West Candor Chasma, Mars: Insights into the geologic history of Valles Marineris, J. Geophys. Res.: Planets, 119, 331-354 Gabasova, L. R.; Kite, E. S., Sediment Compaction on Mars and Its Effect on Layer Orientation, 47th Lunar and Planetary Science Conference, held March 21-25, 2016 at The Woodlands, Texas. LPI Contribution No. 1903, 1209 Gendrin, Aline; Mangold, Nicolas; Bibring, Jean-Pierre; Langevin, Yves; Gondet, Brigitte; Poulet, François; Bonello, Guillaume; Quantin, Cathy; Mustard, John; Arvidson, Ray; Le Mouélic, Stéphane (2005), Sulfates in Martian Layered Terrains: The OMEGA/Mars Express View, Science, 307, 5715, 1587-1591. Gillon, M.; Demory, B.-O.; Barman, T.; Bonfils, X.; Mazeh, T.; Pont, F.; Udry, S.; Mayor, M.; Queloz, D. (2007), Accurate Spitzer infrared radius measurement for the hot Neptune GJ 436b, Astronomy and Astrophysics, 471, L51- L54. Golombek, M. P.; Grant, J. A.; Crumpler, L. S.; Greeley, R.; Arvidson, R. E.; Bell, J. F.; Weitz, C. M.; Sullivan, R.; Christensen, P. R.; Soderblom, L. A.; Squyres, S. W. (2006), Erosion rates at the Mars Exploration Rover landing sites and long-term climate change on Mars, J. Geophys. Res., 111, E12, CiteID E12S10 Golombek, M. P.; Warner, N. H.; Ganti, V.; Lamb, M. P.; Parker, T. J.; Fergason, R. L.; Sullivan, R. (2014), Small crater modification on Meridiani Planum and implications for erosion rates and climate change on Mars, Journal of Geophysical Research: Planets, Volume 119, Issue 12, pp. 2522-2547. Gourronc, Marine; Bourgeois, Olivier; Mège, Daniel; Pochat, Stéphane; Bultel, Benjamin; Massé, Marion; Le Deit, Laetitia; Le Mouélic, Stéphane; Mercier, Denis (2014), One million cubic kilometers of fossil ice in Valles Marineris: Relicts of a 3.5 Gy old glacial landsystem along the Martian equator, Geomorphology, 204, 235-255. Griffiths, Cedric M., Chris Dyt, Evelina Paraschivoiu, Keyu Liu (2001), Sedsim in Hydrocarbon Exploration, pp 71- 97 in Geologic Modeling and Simulation, Part of the series Computer Applications in the Earth Sciences. Grindrod, P. M.; Warner, N. H. (2014), Erosion rate and previous extent of interior layered deposits on Mars revealed by obstructed landslides, Geology, 42, 795-798. Grotzinger, J.P. and R.E. Milliken (2012), The Sedimentary Rock Record of Mars: Distribution, Origins, and Global Stratigraphy, 1-48, in Sedimentary Geology of Mars SEPM Special Publication, Vol. 102, 2012, Edited by John P. Grotzinger and Ralph E. Milliken. Grotzinger, J.P., (2014), Habitability, organic taphonomy, and the search for organic carbon on Mars, Science, 343, 386-7. Grotzinger, J. P.; et al. (2014), A Habitable Fluvio-Lacustrine Environment at Yellowknife Bay, Gale Crater, Mars, Science, 343, 6169, id. 1242777. Grotzinger, J. P.; Gupta, S.; Malin, M. C.; Rubin, D. M.; Schieber, J.; Siebach, K.; Sumner, D. Y.; Stack, K. M.; Vasavada, A. R.; Arvidson, R. E.; Calef, F.; Edgar, L.; Fischer, W. F.; et al. (2015), Deposition, exhumation, and paleoclimate of an ancient lake deposit, Gale crater, Mars, Science, 350, DOI: 10.1126/science.aac7575. Haberle, R. M., J. B. Pollack, J. R. Barnes, R. W. Zurek, C. B. Leovy, J. R. Murphy, H. Lee, and J. Schaeffer (1993), Mars atmospheric dynamics as simulated by the NASA Ames General Circulation Model: 1. The zonal‐ mean circulation, J. Geophys. Res., 98, 3093–3123, doi:10.1029/92JE02946. 54 Haberle, Robert M.; Murphy, James R.; Schaeffer, James (2003), Orbital change experiments with a Mars general circulation model, Icarus, 161, 1, 66-89. Halevy, Itay; Head, James W., III (2014), Episodic warming of early Mars by punctuated volcanism, Nature Geoscience, 7, 12, 865-868. Hayward, R.K., L.K. Fenton, T.N. Titus (2014), Mars Global Digital Dune Database (MGD3): Global dune distribution and wind pattern observations, Icarus 230, 38–46, doi:10.1016/j.icarus.2013.04.011 Harrison, Keith P.; Chapman, Mary G. (2008), Evidence for ponding and catastrophic floods in central Valles Marineris, Mars, Icarus, 198, 351-364. Heermance, R. V.; Pullen, A.; Kapp, P.; Garzione, C. N.; Bogue, S.; Ding, L.; Song, P. (2013), Climatic and tectonic controls on sedimentation and erosion during the Pliocene-Quaternary in the Qaidam Basin (China), Geological Society of America Bulletin, 125, 833-856. Holt, J. W.; Fishbaugh, K. E.; Byrne, S.; Christian, S.; Tanaka, K.; Russell, P. S.; Herkenhoff, K. E.; Safaeinili, A.; Putzig, N. E.; Phillips, R. J. (2010), The construction of Chasma Boreale on Mars, Nature, 465, 446-449 (2010). Hore, Alicia (2014), Structural analysis, layer thickness measurements and mineralogic investigation of the large interior layered deposit within Ganges Chasma, Valles Marineris, Mars, MSci Thesis, Brock University (Canada). Howard, Alan D. (2007), Simulating the development of Martian highland landscapes through the interaction of impact cratering, fluvial erosion, and variable hydrologic forcing, Geomorphology, 91, 3-4, 332-363. Hynek, Brian M.; Phillips, Roger J. (2008), The stratigraphy of Meridiani Planum, Mars, and implications for the layered deposits' origin, Earth and Planetary Science Letters, 274, 1-2, 214-220. Irwin, R.P.; Lewis, K.W.; Howard, A.D.; and Grant, J.A., (2015), Paleohydrology of Eberswalde crater, Mars, Geomorphology 240, 83-101. Isherwood, Ryan J.; Jozwiak, Lauren M.; Jansen, Johanna C.; Andrews-Hanna, Jeffrey C. (2013), The volcanic history of Olympus Mons from paleo-topography and flexural modeling, Earth and Planetary Science Letters, 363, 88-96 Jackson, M.P.A., et al. (1991), Salt Diapirs of the Great Kavir, Central Iran, Geol. Soc. Am. Memoir 177. Jackson, M. P. A.; Adams, J. B.; Dooley, T. P.; Gillespie, A. R.; Montgomery, D. R. (2011), Modeling the collapse of Hebes Chasma, Valles Marineris, Mars, Geological Society of America Bulletin, vol. 123, issue 7-8, 1596- Jakosky, B. M.; Carr, M. H. (1985), Possible precipitation of ice at low latitudes of Mars during periods of high obliquity, Nature, 315, 559-561. Jakosky, Bruce M.; Phillips, Roger J. (2001), Mars' volatile and climate history, Nature, 412, 6843, 237-244. Karimi, Saman, Andrew J. Dombard, Debra L. Buczkowski, Stuart J. Robbins, Rebecca M. Williams (2016), Using the Viscoelastic Relaxation of Large Impact Craters to Study the Thermal History of Mars, Icarus, 272, 102-113. Kass, D. M.; Schofield, J. T.; Michaels, T. I.; Rafkin, S. C. R.; Richardson, M. I.; Toigo, A. D. (2003), Analysis of atmospheric mesoscale models for entry, descent, and landing, J. Geophys. Res., 108, E12, ROV 31-1, CiteID 8090, DOI 10.1029/2003JE002065. Kerber, Laura; Forget, François; Wordsworth, Robin (2015), Sulfur in the early martian atmosphere revisited: Experiments with a 3-D Global Climate Model, Icarus, 261, 133-148. Kinch, Kjartan M.; Sohl-Dickstein, Jascha; Bell, James F.; Johnson, Jeffrey R.; Goetz, Walter; Landis, Geoffrey A. (2007), Dust deposition on the Mars Exploration Rover Panoramic Camera (Pancam) calibration targets, J. Geophys. Res., 112, E6, CiteID E06S03. Kite, E.S., Lewis, K.W., Lamb, M.P., Newman, C.E., & Richardson, M.I., (2013a), Growth and form of the mound in Gale Crater, Mars: Slope-wind enhanced erosion and transport, Geology, 41, 543-546, doi:10.1130/G3309.1 Kite, E.S., Halevy, I., Kahre, M.A., Manga, M., & Wolff, M., (2013b), Seasonal melting and the formation of sedimentary rocks on Mars, Icarus, 223, 181-210. Kite, E.S., Howard, A., Lucas, A., Armstrong, J.C., Aharonson, O., & Lamb, M.P., (2015), Stratigraphy of Aeolis Dorsa, Mars: stratigraphic context of the great river deposits, Icarus, 253, 223-242. Kite, E.S., and Mayer, D.P., Mars erosion rates constrained using crater counts, with applications to organic-matter preservation and to the global dust cycle, submitted to Icarus. Kirk, R. L.; Howington-Kraus, E.; Rosiek, M. R.; Anderson, J. A.; Archinal, B. A.; Becker, K. J.; Cook, D. A.; Galuszka, D. M.; Geissler, P. E.; Hare, T. M.; Holmberg, I. M.; Keszthelyi, L. P.; Redding, B. L.; Delamere, W. A.; 55 Gallagher, D.; Chapel, J. D.; Eliason, E. M.; King, R.; McEwen, A. S. (2008), Ultrahigh resolution topographic mapping of Mars with MRO HiRISE stereo images: Meter-scale slopes of candidate Phoenix landing sites, J. Geophys. Res., 113, E12, CiteID E00A24. Kneissl, T.; van Gasselt, S.; Neukum, G. (2010), Map-projection-independent crater size-frequency determination in GIS environments-New software tool for ArcGIS, Planetary and Space Science, 59, 1243-1254. Kocurek, G. (1988), First-order & super bounding surface in eolian sequences, Sedimentary Geology 56, 193-206. Kok, Jasper F.; Parteli, Eric J. R.; Michaels, Timothy I.; Karam, Diana Bou (2012), The physics of wind- blown sand and dust, Reports on Progress in Physics, 75, id. 106901 (2012). Krapf, C. (2003), Contrasting styles of ephemeral river systems and their interaction with dunes of the Skeleton Coast erg (Namibia), Quaternary International, 104, 41-52 Lapotre, M. G. A.; Ewing, R. C.; Lamb, M. P.; Fischer, W. W.; Grotzinger, J. P.; Rubin, D. M.; Lewis, K. W.; Ballard, M. J.; Day, M.; Gupta, S.; Banham, S. G.; Bridges, N. T.; Des Marais, D. J.; Fraeman, A. A.; Grant, J. A.; Herkenhoff, K. E.; Ming, D. W.; Mischna, M. A.; Rice, M. S.; Sumner, D. A.; Vasavada, A. R.; Yingst, R. A. (2016), Large wind ripples on Mars: A record of atmospheric evolution, Science, 353, 55-58. Laskar, J.; Correia, A. C. M.; Gastineau, M.; Joutel, F.; Levrard, B.; Robutel, P. (2004), Long term evolution and chaotic diffusion of the insolation quantities of Mars, Icarus, 170, 343-364. Le Deit, Laetitia; Hauber, Ernst; Fueten, Frank; Pondrelli, Monica; Rossi, Angelo Pio; Jaumann, Ralf (2013), Sequence of infilling events in Gale Crater, Mars: Results from morphology, stratigraphy, and mineralogy, J. Geophys. Res.: Planets, 118, 2439-2473 Le Mouélic, S.; Gasnault, O.; Herkenhoff, K. E.; Bridges, N. T.; Langevin, Y.; Mangold, N.; Maurice, S.; Wiens, R. C.; Pinet, P.; Newsom, H. E.; Deen, R. G.; Bell, J. F.; Johnson, J. R.; Rapin, W.; Barraclough, B.; Blaney, D. L.; Deflores, L.; Maki, J.; Malin, M. C.; Pérez, R.; Saccoccio, M., (2015), The ChemCam Remote Micro-Imager at Gale crater: Review of the first year of operations on Mars, Icarus, 249, 93-107. Leeder, M.R. (2011), Sedimentology and Sedimentary Basins: From Turbulence to Tectonics, 2nd Edition, Wiley, ISBN: 978-1-4051-7783-2, 784 pages Levy, Joseph S.; Fassett, Caleb I.; Head, James W. (2016), Enhanced erosion rates on Mars during Amazonian glaciation, Icarus, 264, 213-219. Lewis, Kevin W.; Aharonson, Oded; Grotzinger, John P.; Kirk, Randolph L.; McEwen, Alfred S.; Suer, Terry-Ann (2008), Quasi-Periodic Bedding in the Sedimentary Rock Record of Mars, Science, 322, 1532-. Lewis, Kevin W. (2009), The Rock Record of Mars: Structure, Sedimentology and Stratigraphy, PhD Thesis (California Institute of Technology). Lewis, Kevin W.; Aharonson, Oded (2014), Occurrence and origin of rhythmic sedimentary rocks on Mars, J. Geophys. Res.: Planets, 119, 1432-1457 Lewis, K. W.; Grotzinger, J. P.; Gupta, S.; Rubin, D. M. (2015), Investigation of a Major Stratigraphic Unconformity with the Curiosity Rover, American Geophysical Union, Fall Meeting 2015, abstract #P43B-2117. Li, Gongjie; Batygin, Konstantin (2014), On the Spin-axis Dynamics of a Moonless Earth, The Astrophysical Journal, 790, 1, article id. 69, 7 (2014). Lissauer, Jack J.; Barnes, Jason W.; Chambers, John E. (2012), Obliquity variations of a moonless Earth, Icarus, 217, 1, 77-87. Loizeau, D.; Mangold, N.; Poulet, F.; Bibring, J.-P.; Bishop, J. L.; Michalski, J.; Quantin, C. (2015), History of the clay-rich unit at Mawrth Vallis, Mars: High-resolution mapping of a candidate landing site, J. Geophys. Res.: Planets, 120, 11, 1820-1846. Lucchitta, B.K., (1990), Young volcanic deposits in the Valles Marineris, Mars. Icarus 86, 476–509. Lucchitta, B. K.; McEwen, A. S.; Clow, G. D.; Geissler, P. E.; Singer, R. B.; Schultz, R. A.; Squyres, S. W. (1992), The canyon system on Mars In: Mars (University of Arizona Press), Mildred Shapley Matthews; Hugh H. Kieffer; Bruce M. Jakosky; Conway Snyder (editors), 453-492. Lucchitta, B. K. (2015), Geologic Map of West Candor Chasma, Mars: A Progress Report, 46th Lunar and Planetary Science Conference, Woodlands, Texas. LPI Contribution No. 1832, 1550. Madeleine, J.-B.; Forget, F.; Head, James W.; Levrard, B.; Montmessin, F.; Millour, E. (2009), Amazonian northern mid-latitude glaciation on Mars: A proposed climate scenario, Icarus, 203, 2, 390-405. Malin, Michael C.; Edgett, Kenneth S. (2000), Sedimentary Rocks of Early Mars, Science, 290, 1927-1937. 56 Malin, Michael C.; Bell, James F.; Cantor, Bruce A.; Caplinger, Michael A.; Calvin, Wendy M.; Clancy, R. Todd; Edgett, Kenneth S.; Edwards, Lawrence; Haberle, Robert M.; James, Philip B.; Lee, Steven W.; Ravine, Michael A.; Thomas, Peter C.; Wolff, Michael J. (2007), Context Camera Investigation on board the Mars Reconnaissance Orbiter, J. Geophys. Res., 112, E5, CiteID E05S04. Malin, M.C., M. A. Caplinger, K. S. Edgett, F. T. Ghaemi, M. A. Ravine, J. A. Schaffner, J. M. Baker, J. D. Bardis, D. R. DiBiase, J. N. Maki, R. G. Willson, J. F. Bell III, W. E. Dietrich, L. J. Edwards, B. Hallet, K. E. Herkenhoff, E. Heydari, L. C. Kah, M. T. Lemmon, M. E. Minitti, T. S. Olson, T. J. Parker, S. K. Rowland, J. Schieber, R. J. Sullivan, D. Y. Sumner, P. C. Thomas, and R. A. Yingst (2010), The Mars Science Laboratory (MSL) Mast- mounted cameras (Mastcams) flight instruments, 41st Lunar and Planetary Science Conference, Abstract 1123. Mangold, Nicolas; Gendrin, Aline; Gondet, Brigitte; Le Mouelic, Stephane; Quantin, Cathy; Ansan, Véronique; Bibring, Jean-Pierre; Langevin, Yves; Masson, Philippe; Neukum, Gerhard, Spectral and geological study of the sulfate-rich region of West Candor Chasma, Mars, Icarus, 194, 2, 519-543. Mangold, N.; Roach, L.; Milliken, R.; Le Mouélic, S.; Ansan, V.; Bibring, J. P.; Masson, Ph.; Mustard, J. F.; Murchie, S.; Neukum, G. (2010), A Late Amazonian alteration layer related to local volcanism on Mars, Icarus, 207, 1, 265-276. Mayer, D. P.; Kite, E. S. (2016), An Integrated Workflow for Producing Digital Terrain Models of Mars from CTX and HiRISE Stereo Data Using the NASA Ames Stereo Pipeline, 47th Lunar and Planetary Science Conference, held March 21–25, 2016 at The Woodlands, Texas. LPI Contribution No. 1903, 1241. McCauley, J. F. (1978), Geologic Map of the Coprates Quadrangle of Mars. U.S. Geol. Survey Misc. Inv. Series Map 1-897, scale 1 : 5,000,000. McEwen, Alfred S.; Eliason, Eric M.; Bergstrom, James W.; Bridges, Nathan T.; Hansen, Candice J.; Delamere, W. Alan; Grant, John A.; Gulick, Virginia C.; Herkenhoff, Kenneth E.; Keszthelyi, Laszlo; Kirk, Randolph L.; Mellon, Michael T.; Squyres, Steven W.; Thomas, Nicolas; Weitz, Catherine M. (2007), Mars Reconnaissance Orbiter's High Resolution Imaging Science Experiment (HiRISE), J. Geophys. Res., 112, E5, CiteID E05S02. McKay, C. P.; Nedell, S. S. (1988), Are there carbonate deposits in the Valles Marineris, Mars?, vol. 73, Jan. 1988, 142-148. McLennan, S. M.; Grotzinger, J. P. (2008), The sedimentary rock cycle of Mars, 541- in The Martian Surface - Composition, Mineralogy, and Physical Properties. Edited by Jim Bell, Cambridge University Press. Metz, J. M.; Grotzinger, J. P.; Rubin, D. M.; Lewis, K. W.; Squyres, S. W.; Bell, J. F. (2009), Sulfate-Rich Eolian and Wet Interdune Deposits, Erebus Crater, Meridiani Planum, Mars, Journal of Sedimentary Research, vol. 79, issue 5, 247-264. Metz, Joannah; Grotzinger, John; Okubo, Chris; Milliken, Ralph (2010), Thin-skinned deformation of sedimentary rocks in Valles Marineris, Mars, J. Geophys. Res., 115, E11004, doi:10.1029/2010JE003593. Miall, A. (2010), The geology of stratigraphic sequences, Springer. Michalski, J.; Niles, P. B. (2012), Atmospheric origin of Martian interior layered deposits: Links to climate change and the global sulfur cycle, Geology, 40, 419-422. Michaels, T. I., and S. C. R. Rafkin (2008a), Meteorological predictions for candidate 2007 Phoenix Mars Lander sites using the Mars Regional Atmospheric Modeling System (MRAMS), J. Geophys. Res., 113, E00A07, doi:10.1029/2007JE003013 Milliken, R. E.; Swayze, G. A.; Arvidson, R. E.; Bishop, J. L.; Clark, R. N.; Ehlmann, B. L.; Green, R. O.; Grotzinger, J. P.; Morris, R. V.; Murchie, S. L.; Mustard, J. F.; Weitz, C. (2008), Opaline silica in young deposits on Mars, Geology, vol. 36, 11, 847-850. Milliken, R. E.; Grotzinger, J. P.; Thomson, B. J. (2010), Paleoclimate of Mars as captured by the stratigraphic record in Gale Crater, Geophys. Res. Lett., 37, CiteID L04201. Milliken, R. E.; Ewing, R. C.; Fischer, W. W.; Hurowitz, J. (2014), Wind-blown sandstones cemented by sulfate and clay minerals in Gale Crater, Mars, Geophys. Res. Lett., 41, 1149-1154. Mischna, Michael A.; Richardson, Mark I.; Wilson, R. John; McCleese, Daniel J., (2003), On the orbital forcing of Martian water and CO2 cycles: A general circulation model study with simplified volatile schemes, J. Geophys. Res. Planets, 108, E6, 16-1, CiteID 5062, DOI 10.1029/2003JE002051 (JGRE Homepage). Mischna, Michael A.; Baker, Victor; Milliken, Ralph; Richardson, Mark; Lee, Christopher (2013), Effects of obliquity and water vapor/trace gas greenhouses in the early martian climate, J. Geophys. Res.: Planets, 118, 560- 576. 57 Montgomery, D.R., et al. (2009), Continental-scale salt tectonics on Mars and the origin of Valles Marineris and associated outflow channels, GSA Bulletin; January/February 2009; v. 121; no. 1/2; 117–133; doi: 10.1130/B26307.1 Moore, Jeffrey M. (1990), Nature of the mantling deposit in the heavily cratered terrain of northeastern Arabia, Mars, J. Geophys. Res., 95, 14279-14289. Moore, Jeffrey M.; Howard, Alan D. (2005), Large alluvial fans on Mars, J. Geophys. Res., 110, E4, CiteID E04005. Moratto, Z. M.; Broxton, M. J.; Beyer, R. A.; Lundy, M.; Husmann, K. (2010), Ames Stereo Pipeline, NASA's Open Source Automated Stereogrammetry Software, 41st Lunar and Planetary Science Conference, held March 1-5, 2010 in The Woodlands, Texas. LPI Contribution No. 1533, 2364. Moreau, A.J.M., R.M. Haberle, S.C.R Rafkin, M.A. Kahre, and J.L. Hollingsworth, 2014, Dust erosion and sedimentation patterns in Gale Crater as simulated by the Mars Regional Atmospheric Modeling System (MRAMS), Eighth International Conference on Mars, abstract #1430. MSL Extended Mission Plan (2014), (led by Erickson, J. & Grotzinger, J.P), Mars Science Laboratory Senior Review Proposal / Extended Missions Plan, Mission to Mt. Sharp: Habitability, Preservation of Organics, and Environmental Transitions, NASA/JPL, downloaded from http://mars.nasa.gov/files/msl/2014-MSL-extended- mission-plan.pdf Murchie, Scott L.; Mustard, John F.; Ehlmann, Bethany L.; Milliken, Ralph E.; Bishop, Janice L.; McKeown, Nancy K.; Noe Dobrea, Eldar Z.; Seelos, Frank P.; Buczkowski, Debra L.; Wiseman, Sandra M.; Arvidson, Raymond E.; Wray, James J.; Swayze, Gregg; Clark, Roger N.; Des Marais, David J.; McEwen, Alfred S.; Bibring, Jean-Pierre (2009a), A synthesis of Martian aqueous mineralogy after 1 Mars year of observations from the Mars Reconnaissance Orbiter, J. Geophys. Res.: Planets, 114, CiteID E00D06. Murchie, Scott; Roach, Leah; Seelos, Frank; Milliken, Ralph; Mustard, John; Arvidson, Raymond; Wiseman, Sandra; Lichtenberg, Kimberly; Andrews-Hanna, Jeffrey; Bishop, Janice; Bibring, Jean-Pierre; Parente, Mario; Morris, Richard (2009b), Evidence for the origin of layered deposits in Candor Chasma, Mars, from mineral composition and hydrologic modeling, J. Geophys. Res., 114, E12, CiteID E00D05. Nedell, S. S.; Squyres, S. W.; Andersen, D. W. (1987), Origin and evolution of the layered deposits in the Valles Marineris, Mars, Icarus, 70, 409-441. Neuffer, D.P., and R.A. Schultz (2006), Mechanisms of slope failure in Valles Marineris, Mars, Quarterly Journal of Engineering Geology and Hydrogeology, 39, 227–240. Newman, Claire E.; Lewis, Stephen R.; Read, Peter L. (2005), The atmospheric circulation and dust activity in different orbital epochs on Mars, Icarus, 174, 1, 135-160. Okubo, Chris H.; Lewis, Kevin W.; McEwen, Alfred S.; Kirk, Randolph L. (2008), Relative age of interior layered deposits in southwest Candor Chasma based on high-resolution structural mapping, J. Geophys. Res., 113, CiteID E12002. Okubo, Chris H. (2010), Structural geology of Amazonian-aged layered sedimentary deposits in southwest Candor Chasma, Mars, Icarus, 207, 1, 210-225. Okubo, C.H. (2014), Bedrock geologic and structural map through the western Candor Colles region of Mars: U.S. Scientific Geological 1:18,000, http://dx.doi.org/10.3133/sim3309. Palucis, M., William E Dietrich, Rebecca M.E. Williams, Alexander G. Hayes, Tim Parker, Dawn Y. Sumner, Nicolas Mangold, Kevin Lewis, Horton Newsom (2016), Sequence and relative timing of large lakes in Gale crater (Mars) after the formation of Mt. Sharp, J. Geophys. Res. – Planets, doi: 10.1002/2015JE004905 Parish, Thomas R.; Bromwich, David H. (1991), Continental-Scale Simulation of the Antarctic Katabatic Wind Regime, Journal of Climate, vol. 4, 2, 135-146. Parker, T. J.; Dietrich, W. E.; Palucis, M. C.; Calef, F. J., III; Newsom, H. E. (2014), Banding in Mount Sharp, Gale Crater: Stratigraphy, Strandlines, or Buttress Unconformities? American Geophysical Union, Fall Meeting 2014, abstract #P43D-4013 Peralta, J.; Fueten, F.; Cheel, R.; Stesky, R.; Flahaut, J.; Hauber, E. (2015), Layer Attitude and Thickness Measurements in Western Portion of the Ophir Chasma Interior Layered Deposit, Valles Marineris, Mars, 46th Lunar and Planetary Science Conference, held March 16-20, 2015 in The Woodlands, Texas. LPI Contribution No. 1832, 1153 Perkins, J.P., Finnegan, N.J., and de Silva, S.L. (2015), Amplification of bedrock canyon incision by wind, Nature Geoscience 8, 305-310 (2015). Investigations Map Survey 3309, pamphlet 8 , scale 58 Peterson, C.M. (1981). Hebes Chasma - Martian Pyroclastic Sink. Lunar and Planetary Science Conference 12, 828- 829. Pyle, David M. (1998), Forecasting sizes and repose times of future extreme volcanic events, Geology, vol. 26, 4, 367-dyk. Rafkin, S.C.R., Haberle, R. M., and T. I. Michaels (2001), The Mars Regional Atmospheric Modeling System (MRAMS): Model description and selected simulations. Icarus, 151, 228-256. Rafkin, Scot C. R.; Michaels, T. I. (2003), Meteorological predictions for 2003 Mars Exploration Rover high- priority landing sites, J. Geophys. Res., 108, E12, ROV 32-1, CiteID 8091, DOI 10.1029/2002JE002027. Rafkin, S.C.R., J. Pla-Garcia, M. Kahre, J. Gomez-Elvira, V.E. Hamilton, et al. (2016), The meteorology of Gale Crater as determined from Rover Environmental Monitoring Station observations and numerical modeling. Part II: Interpretation, Icarus, in press, doi:10.1016/j.icarus.2016.01.031 Ramirez, R., et al., 2014. Warming early Mars with CO2 and H2. Nat. Geosci. 7, 59–63. Richardson, Mark I.; Mischna, Michael A. (2005), Long-term evolution of transient liquid water on Mars, J. Geophys. Res., 110, E3, CiteID E03003. Richardson, M.I., Toigo, A.D., and Newman, C.E. (2007), PlanetWRF: A general purpose, local to global numerical model for planetary atmospheric and climate dynamics, J. Geophys. Res. 112, 150 E09001 Roach, Leah H., (2009), Sulfates in Valles Marineris as Indicators of the Aqueous Evolution of Mars, PhD thesis, Brown University. Roach, Leah H.; Mustard, John F.; Lane, Melissa D.; Bishop, Janice L.; Murchie, Scott L. (2010), Diagenetic haematite and sulfate assemblages in Valles Marineris, Icarus, 207, 659-674. Rossi, Angelo Pio; Neukum, Gerhard; Pondrelli, Monica; van Gasselt, Stephan; Zegers, Tanja; Hauber, Ernst; Chicarro, Agustin; Foing, Bernard (2008), Large-scale spring deposits on Mars?, J. Geophys. Res., 113, E8, CiteID E08016. Rubin, D.M., Hunter, R.E. (1982), Bedform climbing in theory and nature, Sedimentology 29, 121-138. Schmidt, G. (2016), Geology of Hebes Chasma, Valles Marineris, Mars, MSci Thesis, Brock University. Schmidt, G.; Flahaut, J.; Fueten, F.; Hauber, E.; Stesky, R. (2015), Evidence for an Unconformity Within the Interior Layered Deposit of Hebes Chasma, Valles Marineris, Mars, 46th Lunar and Planetary Science Conference, held March 16-20, 2015 in The Woodlands, Texas. LPI Contribution No. 1832, 1237. Schultz, R.A. (1998), Multiple-process origin of Valles Marineris basins and troughs, Mars, Planetary and Space Science 46(6), 827-834. Schultz, Richard A. (2002), Stability of rock slopes in Valles Marineris, Mars, Geophys. Res. Lett., 29, CiteID 1932, DOI 10.1029/2002GL015728 Segura, T.L., Zahnle, K., Toon, O.B., McKay, C.P., 2013. The effects of impacts on the climates of terrestrial planets. In: Mackwell, S.J., et al. (Eds.), Comparative Climatology of Terrestrial Planets. U. Arizona Press, 417– 438. Shao, Y. (2008), Physics and modeling of wind erosion, 2nd revised and expanded edition, Springer. Silvestro, S., Vaz, D.A., Ewing, R.C., Rossi, A.P., Fenton, L.K., Michaels, T.I., Flahaut, J., Geissler, P.E. 2013. Pervasive aeolian activity along rover Curiosity's traverse in Gale Crater, Mars. Geology 41, 483-486. Sharp, R. P. (1940), Ep-Archean and Ep-Algonkian erosion surfaces, Grand Canyon, Arizona, Geological Society of America Bulletin, 51, 1235-1269 Shean, D.E.; Alexandrov, O., Moratto, Z.M.; Smith, B.E.; Joughin, I.R.; Porter, C.; Morin, P. (2016), An automated, open-source pipeline for mass production of digital elevation models (DEMs) from very-high-resolution commercial stereo satellite imagery, ISPRS Journal of Photogrammetry and Remote Sensing, vol. 116, 101-117. Sloss, L. L. (1963), Sequences in the Cratonic Interior of North America, Geological Society of America Bulletin, 74, 93 Soto, Alejandro; Mischna, Michael; Schneider, Tapio; Lee, Christopher; Richardson, Mark (2015), Martian atmospheric collapse: Idealized GCM studies, Icarus, 250, 553-569. 59 Spiga, A. (2011), Elements of comparison between Martian and terrestrial mesoscale meteorological phenomena: Katabatic winds and boundary layer convection, Planetary and Space Science, 59, 10, 915-922. Spiga, A., F. Forget, J.-B. Madeleine, L. Montabone, S.R. Lewis, E. Millour (2011), The impact of martian mesoscale winds on surface temperature and on the determination of thermal inertia, Icarus, 212, 504-519 Stack, K. M.; Grotzinger, J. P.; Milliken, R. E. (2013), Bed thickness distributions on Mars: An orbital perspective, J. Geophys. Res.: Planets, 118, 1323-1349 Stack, K.M., C.S. Edwards, J.P. Grotzinger, S. Gupta, D.Y. Sumner, F.J. Calef III, et al. (2016), Comparing orbiter and rover image-based mapping of an ancient sedimentary environment, Aeolis Palus, Gale crater, Mars, Icarus, Available online 27 February 2016, doi:10.1016/j.icarus.2016.02.024 Straub, K. M.; Paola, C.; Mohrig, D.; Wolinsky, M. A.; George, T. (2009), Compensational Stacking of Channelized Sedimentary Deposits, Journal of Sedimentary Research, vol. 79, issue 9, 673-688 Summons, R.E., J.P. Amend, D. Bish, R. Buick, G.D. Cody, D.J. Des Marais, G. Dromart, J.L. Eigenbrode, A.H. Knoll, and D.Y. Sumner (2011) Preservation of martian organic and environmental records: Final report of the Mars biosignature working group, Astrobiology, 11(2): doi:10.1089/ast.2010.0506. Thollot, P.; Mangold, N.; Ansan, V.; Le Mouélic, S.; et al. (2012), Most Mars minerals in a nutshell: Various alteration phases formed in a single environment in Noctis Labyrinthus, J. Geophys. Res., 117, CiteID E00J06 Thomson, B. J.; Bridges, N. T.; Milliken, R.; Baldridge, A.; Hook, S. J.; Crowley, J. K.; Marion, G. M.; de Souza Filho, C. R.; Brown, A. J.; Weitz, C. M. (2011), Constraints on the origin and evolution of the layered mound in Gale Crater, Mars using Mars Reconnaissance Orbiter data, Icarus, 214, 413-432. Toigo, A.D.; Lee, C.; Newman, C. E.; Richardson, M. I. (2012), The impact of resolution on the dynamics of the martian global atmosphere: Varying resolution studies with the MarsWRF GCM, Icarus, 221, 1, 276-288. Tyler, Daniel, & Barnes, Jeffrey R. (2015), Convergent crater circulations on Mars: Influence on the surface pressure cycle and the depth of the convective boundary layer, Geophys. Res. Lett., 42, 18, 7343-7350. Wang, Z.-T., Wang, H.-T., Niu, Q.-H., Dong, Z.-B. and Wang, T., (2011), Abrasion of yardangs. Physical Review E, 84, 031304. Warner, N. H.; Sowe, M.; Gupta, S.; Dumke, A.; Goddard, K. (2013), Fill and spill of giant lakes in the eastern Valles Marineris region of Mars, Geology, 41, 675-678 Watkins, J. A.; Grotzinger, J.; Stein, N.; Banham, S. G.; Gupta, S.; Rubin, D.; Stack, K. M.; Edgett, K. S. (2016), 47th Lunar and Planetary Science Conference, held March 21–25, 2016 at The Woodlands, Texas. LPI Contribution No. 1903, 2939 Wendt, L.; Gross, C.; Kneissl, T.; Sowe, M.; Combe, J.-P.; et al. (2011), Sulfates and iron oxides in Ophir Chasma, Mars, based on OMEGA and CRISM observations, Icarus, 213, 86-103. Williams, R.M. E.; Weitz, C.M. (2014), Reconstructing the aqueous history within the southwestern Melas basin, Mars: Clues from stratigraphic and morphometric analyses of fans, Icarus, 242, 19-37. Wilson, S.A.; Howard, A.D.; Moore, J.M.; Grant, John A. (2007), Geomorphic and stratigraphic analysis of Crater Terby and layered deposits north of Hellas basin, Mars, J. Geophys. Res., 112, E8, CiteID E08009. Weitz, Catherine M.; Lane, Melissa D.; Staid, Matthew; Dobrea, Eldar Noe (2008), Gray hematite distribution and formation in Ophir and Candor chasmata, J. Geophys. Res., 113, E2, CiteID E02016. Weitz, C.M., Noe Dobrea, E., Wray, J.J. (2015) Mixtures of clays and sulfates within deposits in western Melas Chasma, Mars. Icarus 251, 291-314. Weitz, Catherine M.; Bishop, Janice L. (2016), Stratigraphy and formation of clays, sulfates, and hydrated silica within a depression in Coprates Catena, Mars, J. Geophys. Res.: Planets, 121, 5, 805-835. Wordsworth, R., et al., (2013). Global modelling of the early martian climate under a denser CO2 atmosphere: water cycle and ice evolution. Icarus 222, 1–19. Wordsworth, R.D.; Kerber, L.; Pierrehumbert, R.T.; Forget, F.; Head, J.W. (2015), Comparison of "warm and wet" and "cold and icy" scenarios for early Mars in a 3-D climate model, J. Geophys. Res.: Planets, 120, 6, 1201-1219 Wordsworth, Robin D. (2016), The Climate of Early Mars, Annual Review of Earth and Planetary Sciences, vol. 44, 381-408. 60 Wylie, Jonathan J.; Helfrich, Karl R.; Dade, Brian; Lister, John R.; Salzig, John F. (1999), Flow localization in fissure eruptions, Bulletin of Volcanology, 60, 6, 432-440. Zardi, D., CD Whiteman (2013), Diurnal mountain wind systems, in Mountain Weather Research and Forecasting, Recent Progress and Current Challenges, edited by Fotini K. Chow, Stephan F.J. De Wekker, Bradley J. Snyder ISBN: 978-94-007-4097-6 (Print) 978-94-007-4098-3 (Online) Zimbelman, J.; Scheidt, S. (2012), Hesperian Age for Western Medusae Fossae Formation, Mars, Science, 336, 1683. 61
0912.2773
1
0912
2009-12-14T23:14:15
Evidence of a massive planet candidate orbiting the young active K5V star BD+20 1790
[ "astro-ph.EP", "astro-ph.SR" ]
BD+20 1790 is a young active, metal-rich, late-type K5Ve star. We have undertaken a study of stellar activity and kinematics for this star over the past few years. Previous results show a high level of stellar activity, with the presence of prominence-like structures, spots on the surface and strong flare events, despite the moderate rotational velocity of the star. In addition, radial velocity variations with a semi-amplitude of up to 1 km/s were detected. We investigated the nature of these radial velocity variations, in order to determine whether they are due to stellar activity or the reflex motion of the star induced by a companion. We have analysed high-resolution echelle spectra and also two-band photometry was obtained to produce the light curve and determine the photometric period. Based upon the analysis of the bisector velocity span, as well as spectroscopic indices of chromospheric indicators and taking into account the photometric analysis, we report that the best explanation for the RV variation is the presence of a sub-stellar companion. The Keplerian fit of the RV data yields a solution for a close-in massive planet with an orbital period of 7.78 days. The presence of the close-in massive planet could also be an interpretation for the high level of stellar activity detected. Since the RV data are not part of a planet search program, we can consider our results as a serendipitous evidence of a planetary companion. To date, this is the youngest main sequence star for which a planetary candidate has been reported.
astro-ph.EP
astro-ph
Astronomy&Astrophysicsmanuscript no. 11000 January 3, 2018 c(cid:13) ESO 2018 Evidence of a massive planet candidate orbiting the young active K5V star BD+20 1790 ⋆ M. Hern´an-Obispo1, M.C. G´alvez-Ortiz2, G. Anglada-Escud´e3,4, S.R. Kane5, J.R. Barnes2, E. de Castro1, and M. Cornide1 1 Dpto. de Astrof´ısica y Ciencias de la Atm´osfera, Facultad de F´ısica, Universidad Complutense de Madrid, Avda. Complutense s/n, E-28040, Madrid, Spain e-mail: [email protected] 2 Centre for Astrophysics Research, Science & Technology Research Institute, University of Hertfordshire, College Lane, Hatfield, Hertfordshire AL10 9AB, UK 3 Department of Terrestrial Magnetism, Carnegie Institution of Washington, 5241 Broad Branch Road, NW, Washington, DC 20015- 1305, USA 4 Departament d ′Astronomia i Meteorologia. Universitat de Barcelona, Mart´ı i Franqu´es 1, Barcelona, 08028. Spain 5 NASA Exoplanet Science Institute, Caltech, MS 100-22, 770 South Wilson Avenue, Pasadena, CA 91125, USA recieved – ; accepted – ABSTRACT Context.BD+20 1790 is a young active, metal-rich, late-type K5Ve star. We have undertaken a study of stellar activity and kinematics for this star over the past few years. Previous results show a high level of stellar activity, with the presence of prominence-like structures, spots on the surface and strong flare events, despite the moderate rotational velocity of the star. In addition, radial velocity variations with a semi-amplitude of up to 1 km s−1 were detected. Aims. We investigated the nature of these radial velocity variations, in order to determine whether they are due to stellar activity or the reflex motion of the star induced by a companion. Methods. We have analysed high-resolution echelle spectra, by measuring stellar activity indicators, and computing radial velocity (RV) and bisector velocity spans. Also two-band photometry was obtained to produce the light curve and determine the photometric period. Results. Based upon the analysis of the bisector velocity span, as well as spectroscopic indices of chromospheric indicators, like e.g. Ca II H & K, Hα, and taking into account the photometric analysis, we report that the best explanation for the RV variation is the presence of a sub-stellar companion. The Keplerian fit of the RV data yields a solution for a close-in massive planet with an orbital period of 7.78 days. The presence of the close-in massive planet could also be an interpretation for the high level of stellar activity detected. Since the RV data are not part of a planet search program, we can consider our results as a serendipitous evidence of a planetary companion. To date, this is the youngest main sequence star for which a planetary candidate has been reported. Key words. stars: activity - stars: late-type - stars: individual (BD+20 1790) - stars: planetary systems 1. Introduction Since the detection of the first planet orbiting a main sequence star, 51 Peg (Mayor & Queloz 1995), the radial velocity (RV) method has become the most successful technique for detecting exoplanets as the vast majority have thus far been discovered in this way (Udry & Santos 2007). This method is especially efficient for giant planets in close-in orbits owing to the large radial velocities they induce in the host star. The use of the RV ⋆ Based collected at on observations the German-Spanish jointly operated by the Max- Astronomical Center, Calar Alto, Planck-Institut fur Astronomie Heidelberg and the Instituto de Astrof´ısica de Andaluc´ıa (CSIC). Based on observations made with the Italian Telescopio Nazionale Galileo (TNG) operated on the island of La Palma by the Fundaci´on Galileo Galilei of the INAF (Istituto Nazionale di Astrofisica) at the Spanish Observatorio del Roque de los Muchachos of the Instituto de Astrof´ısica de Canarias. Based on observations made with the Liverpool Telescope operated on the island of La Palma by Liverpool John Moores University in the Spanish Observatorio del Roque de los Muchachos of the Instituto de Astrof´ısica de Canarias with financial support from the UK Science and Technology Facilities Council. technique to detect exoplanets around young and active stars re- quires, in addition, a careful characterization of stellar activity. An active region on the stellar surface can produce changes in the shape of the spectral lines, thus inducing a subsequent tem- poral variation of the RVs that may mimic a planetary reflex mo- tion with a period equal to the rotational period of the star (Saar & Donahue 1997). Some cases of false planetary detections are provided by Queloz et al. (2001), Bouvier et al. (2007), Huerta et al. (2008) and Hu´elamo et al. (2008). Thus the challenge in using the RV technique to detect young planets lies in disentan- gling the increased levels of stellar activity of young stars from the RV signals of the planets. There is an absence of planets detected around stars younger than 100 Myr (Setiawan et al. 2007, Setiawan et al. 2008). Most RV searches for planetary companions have focussed mainly on stars older than 1 Gyr. Young stars were omitted from RV surveys until recently. Nevertheless, great effort has been made by several groups that have targeted young objects in their RV searches of planetary companions. For example, surveys are being carried out which focus on both nearby associations of young stars and moving groups with ages ranging 10–500 Myr; 2 M. Hern´an-Obispo et al.: Massive candidate orbiting young K5 star examples of which include β Pic (12 Myr), UMa association (300 Myr), Pleiades (100 Myr), IC 2391 (35 Myr), Hyades (700 Myr), Taurus association (2 Myr), ChaI (2 Myr), TWA (10 Myr) (Paulson et al. 2004, Paulson & Yelda 2007, Esposito et al. 2006, Huerta et. al 2007, Setiawan et al. 2007, Setiawan et al. 2008, Prato et al. 2008). Positive identification of planetary signatures from these efforts are few, with only two candidates to date: HD 70573 (Setiawan et al. 2007) and the controversial TW Hya (Setiawan et al. 2008). Planets orbiting around young stars are particularly valuable as they enable us to investigate some of the critical questions about the formation of both stellar and plane- tary systems. How, and at what stage planets form, what is the planet formation mechanism, and how they evolve are important questions which the study of young planetary systems will help to answer. In this paper we report strong evidence of a plane- tary candidate orbiting the young and active K5V star BD+20 1790. Sect. 2, is an overview of the properties and our previous studies of this star. The observational strategy and data analysis are presented in Sect. 3. In Sect. 4 the nature of RV variations is investigated. An orbital solution for the data is presented in Sect. 5, and in Sect. 6 a discussion about planetary parameters, orbital solution, and how stellar activity and the planet are re- lated is shown. Finally, we summarize and offer some conclud- ing remarks in Sect. 7. 2. BD+20 1790: An overview BD+20 1790 was classified by Jeffries (1995) as a K5Ve star, with a magnitude of V = 9.9. Mason et al. (1995) identified this star as the optical counterpart for the 2RE J072343.6+202500 EUV source, located in the ROSAT All-Sky Survey. L´opez- Santiago et al. (2006) proposed its membership in the AB Dor kinematic moving group which has an estimated average age of 50 Myr. By comparing the equivalent width of Li λ 6708 Å with the spectral type, L´opez-Santiago et al. (2006) derived an age estimate of 35–80 Myr. The main stellar parameters for BD+20 1790 are compiled in Table 1. We obtained a value for the stellar radius from the measured rotational velocity and photometric pe- riod. Our estimated radius agrees with the previous K5V spectral classification (from Carrol & Ostlie (2007) tables). Adopting this spectral type, we used the K5V temperature from the Carrol & Ostlie (2007) tables. In conjunction with the photometric param- eters, this enabled us to derive the luminosity, mass and surface gravity. Errors in the parameters were estimated by following the method of propagation of errors, i.e., the uncertainties were calculated from the errors in the variables involved in the deter- mination of each parameter. It has been assumed null correlation between the different variables, in principle independent of each other. In order to test if assuming a fixed value for Teff has a non-negligible effect in the errors computation, we investigated whether an error in Teff could translate into uncertainties of de- rived parameters. We have considered an input error in Teff ∼10 K and conducted an analysis in the propagation of Teff error. Based on this analysis, we have noticed that not consider the er- ror in the Teff leads to an underestimation in the errors in mass and log g, providing unreliable error bars for these parameters. The X-ray luminosity was calculated using the count rates and HR1 hardness ratios from the ROSAT All-Sky Survey. By combining the conversion factor Cx, computed by the formula from Fleming (1995), and the distance estimated by Reid et al. (2004), the stellar X-ray luminosity was calculated as LX = 1.6± 0.5 1029 erg s−1. We compute a preliminary value of metallicity by using a grid of Kurucz et al. (1993) ATLAS9 atmospheres and the 2002 version of MOOG1 synthesis code (Sneden 1973). Atmospheric mod- els were constructed with the data given in Table 1. We used 12 Fe I lines selected from Gonz´alez et al. (2001). We also calcu- lated a value of metallicity by using 7 Fe I lines in the MOOG Abfind routine. We find an average value of A[Fe] = 7.82±0.20 which, when assuming a solar value of A[Fe] = 7.52, results in a [Fe/H] = 0.30±0.20. As mentioned, this is a preliminary value, although compared with the average metallicity of stars of so- lar neighbourhood, we still could consider the star as metal-rich within the error bars. In a recent paper Carpenter et al. (2008) derived the tempera- ture, gravity and metallicity for BD+20 1790, being their values of Teff = 4408 K and log g = 4.50, very close to the correspond- ing values presented in Table 1. We also pointed out that the difference between metallicity values may be explained by the fact that Carpenter et al. (2008) only assumed a fixed metallicity of [Fe/H] = 0.0, but not actually compute it. The Li I abundances analysis was also done in standard local thermodynamic equilibrium (ETL) using MOOG and ATLAS9 in the same way as with the metallicities. Abundances were derived by fitting synthetic spectra to the data. To determine Li abundances we perform a spectral synthesis around the Li I 6707 Å resonance doublet, fitting all spectra between 6702 and 6712 Å, taking into account the relation between Li6 and Li7 isotopes. We determine an average value of lithium abundance of logN(Li) = 1.03±0.04 (where logN(Li) = log (Li/H)+12). In order to study the stellar activity and the kinematics, we have carried out both spectroscopic and photometric monitoring over the past few years: high temporal and spectroscopic resolu- tion and two band photometry. The simultaneous study of pho- tospheric and chromospheric active regions is a powerful tool that allow us to trace, reconstruct and model the puzzle of the magnetic field topology, since these active regions are the finger- prints of magnetic fields (Collier Cameron 2001, Catalano et al. 2002, Frasca et al. 2005, Collier Cameron et al. 2002). Strong chromospheric activity was detected in several observing runs, described by Hern´an-Obispo et al. (2005, 2007). In spite of the fact that the rotational velocity is not very high, vsini ∼10 km s−1 (L´opez-Santiago et al. 2006), all activity indicators are in emis- sion above continuum, from Ca ii H & K, to Ca ii IRT lines (see Fig. 1). Through the study of profile line asymmetries of Hα and Hβ lines, prominence-like structures have been detected in the chro- mosphere of the star (Hern´an-Obispo 2005, 2007). These can be observationally detected as transient absorption features super- imposed on the line profile that are interpreted as the presence of cool material embedded in the surrounding hotter corona and co- rotating with star (Collier Cameron & Robinson 1989a,b, Collier Cameron & Woods 1992, Jeffries et al. 1993, Byrne et al. 1996, Eibe et al. 1998, Barnes et al. 2000, Donati et al. 2001). Several completed prominence-like transients have been detected with durations of orders of a few hours (see Hern´an-Obispo 2005 for details). Modeling these chromospheric phenomenae is an important challenge in this case, due to the detection of these prominence-like structures in unstable positions, far from equa- torial regions (Ferreira 2000, Jardine et al. 2001, Jardine & van Balegooijen 2006). In addition, strong large optical flare events were observed. The gradual decay of the flares was observed for up to 5 hours. Fig. 1 compares the activity indicators for the quiescent state and flare state. The energy released is on the order of ∼1037 erg, 1 The source code of MOOG 2002 can be downloaded at http://verdi.as.utexas.edu/moog.html Table 1. Stellar Parameters of BD+20 1790 M. Hern´an-Obispo et al.: Massive candidate orbiting young K5 star 3 Parameter Spectral Type B − V Ma T b eff log ga EW(Li)a Distancee Agec vsinid Pa ia Ra [Fe/H]a logN(Li)a La X La phot Value K5 V 1.15 0.63 ± 0.09 M⊙ 4410 K 4.53 ± 0.17 110 ± 3 mÅ 25.4 ± 4 pc 35 - 80 Myr 10.03 ± 0.47 km s−1 2.801 ± 0.001 days 50.41 degrees 0.71 ± 0.03 R⊙ 0.30 ± 0.20 1.03 ± 0.04 1.6 ± 0.5 1029 erg s−1 0.17 ± 0.04 L⊙ erg s−1 a This paper b From Carrol & Ostlie, 2007 c From L´opez Santiago et al. 2006 d From L´opez Santiago 2005 e From Reid et al. 2004 while for largest solar flares the released energy is about ∼1029– 1032 erg, thus ranging the flares of BD+20 1790 on the so-called superflare regime (Rubenstein & Schaefer 2000). The photometric observations yielded a light curve with ev- idence of rotational modulation, the semi-amplitude of which is up to ∆V∼ 0.m06 and indicates the presence of spots on the sur- face. The period analysis of the entire set of observations reveals a photometric period of 2.801 (± 0.001) days, in agreement with the period given by the SuperWASP photometric survey (Norton et al. 2007). A detailed and completed study of the chromospheric and photospheric activity characterization will be published in a forthcoming paper (Hern´an-Obispo et al. 2009b, in prep.). 3. Observations and Data Analysis In order to study and characterize active regions at photospheric and chromospheric levels, we carried out photometric and spec- troscopic observations of the target. 3.1. Spectroscopicdata The observational strategy was designed to spectroscopically monitor chromospheric activity indicators with high temporal and spectral resolution. High resolution echelle spectra were ob- tained during four observing runs, from 2004 to 2007, detailed in Table 2. The exposure times ranged from 900 s to 1200 s, depending on weather conditions, in order to obtain a S/N typ- ically greater than 140 for SARG runs and 80 for FOCES runs. The spectra in the time series observations were separated only by the CCD readout time, thus enabling us to obtain the high- est temporal resolution possible. Our initial temporal cadence was designed to detect prominence-like transient features in the Balmer lines. Spectral types and RV standards were acquired with the same setup and configuration as the target. These stan- dards were reduced and analysed in the same way as the tar- Fig. 1. Chromospheric activity indicators. The dashed line indi- cates quiescent state, while solid line indicates flare state. From top to bottom and left to right: He I D3 region, Ca ii K, Hα and Hβ get. The data were bias-subtracted, overscan-corrected and flat- fielded using standard routines in IRAF2 package. 2 IRAF is distributed by the National Optical Observatory, which is operated by the Association of Universities for Research in Astronomy, Inc., under contract with the National Science Foundation. 4 M. Hern´an-Obispo et al.: Massive candidate orbiting young K5 star The wavelength calibration was obtained by taking spectra of a Th-Ar lamp. Using Coud´e spectrographs allowed a stable environment for the wavelength calibration, since flexures are not possible. Details about the spectrographs used can be seen in Pfeiffer et al. (1998) for FOCES spectrograph and Gratton et al. (2001) for SARG spectrograph. In order to enhance the accuracy in calibration we used about 10–12 lines identified per order; across all orders for SARG spectra and about 80 orders for FOCES spectra. The orders were calibrated simultaneously and the total fit has an rms value typically lower than 0.003 Å. The spectra were normalized by a polynomial fit to the observed continuum. Heliocentric radial velocities were determined using a weighted cross-correlation method. The spectra of the star were correlated order by order against spectra of several RV standards with similar spectral type. Orders with chromospheric features and telluric lines were excluded. We calculated the uncertainties based on the cross-correlation peak height and the antisymmetric noise as described by Tonry & Davis (1979). Also, by measur- ing RVs of the standard stars, we estimated the systematic errors and the accuracy of the RV measurements with our instrumental setup. The accuracy between standards for the same run and be- tween runs is less than 0.05 km s−1. Additional echelle data were acquired in DDT mode at the FOCES spectrograph on December 2008. The telescope config- uration and the setup were identically to previous FOCES runs, except for two nights in which a different CCD was used. Data were taken over 10 consecutive nights but due to bad weather conditions only five nights were acquired. Because of the time limitation in DDT mode, only one RV standard was observed. 3.2. Photometricdata The purposes of these observations were to determine the photo- metric period and to look for photometric variability. In addition, the study of the light curve, as well as spectrocopy, allow us to characterize the active regions in the photosphere (Catalano et al. 2002, Frasca et al. 2005, Biazzo et al. 2007). CCD differential aperture photometry was obtained using the 2.0 m fully robotic Liverpool Telescope (Steele et al. 2004) at the Observatorio del Roque de los muchachos in La Palma, Spain. The observations were scheduled in monitoring mode. We ob- tained 22 photometric epochs during November and December 2007. Our observational strategy permitted us to obtain a photo- metric epoch every 3 nights on average. Each epoch consisted in alternating r′ and g′ exposures3, thereby obtaining quasi- simultaneous two band photometry. Custom made software4 was used to automatically extract the photometry. By analysing intra night scatter we can infer a photometric accuracy of 3 mmag and 4 mmag per exposure (r′ and g′ bands respectively, see Fig. 2). We fit the best sine-wave model to the photometry sampling many periods between 0.1 and 50 days, on both bands. Plotting the post-fit residuals as a function of the period, a very strong minimum on the post-fit residuals is found at 2.801 ± 0.001 days in both bands (see Fig. 3). We note that the period and the am- plitude are similar with those given by the SuperWASP survey (Norton et al. 2007). The different amplitude in each band is consistent with large spot or spot group covering at least 4% of the surface. As can be seen in Fig. 2, the amplitude is larger at shorter wavelength, i. 3 Sloan r' and g' filters were used 4 ATP, TATOOINE http://www.am.ub.es/∼anglada/atp/atp testing.htm Automatic Photometry. 0.06 0.04 0.02 0 -0.02 -0.04 -0.06 0.02 0 -0.02 ) g a m ( e d u t i n g a m e v i t a l e R ) g a m ( - C O g' band r' band 0 0.2 0.4 0.6 0.8 1 Phase Fig. 2. Photometry phased to the 2.801 days period. A linear trend and a zero point have been subtracted to both bands The residuals with respect to a simple sine-wave model are show in the lower panel. 0.05 0.04 0.03 0.02 0.01 ) g a m n i ( s m r t i f - t s o P RV period Photometric period g' band r' band 0 0.125 0.25 0.5 1 2 4 8 16 32 Period (days) Fig. 3. Postfit residuals to the photometry as a function of the period. The sharper minima correspond to the 2.801 days period in both bands. The RVs period is marked in grey to illustrate the absence of related photometric signals. e. at g′ band in this case. This color variation is correlated with variation in magnitude. The star appears redder when fainter, at minimum light and therefore bluer when brighter, at maximum light. The full analysis of the photometry and its relation to the star activity requires simultaneous discussion with the spectroscopic data and a more detailed study of the star will be presented else- where (Hern´an–Obispo et al. 2009b, in prep.) Table 2. Observing runs M. Hern´an-Obispo et al.: Massive candidate orbiting young K5 star 5 Date Telescope Instrument CCD chip # Spect. range (Å) Orders Dispersion (Å/pix) FWHMc (Å) N. Obs. 29/03-6/04 2004 21-22/11/2004 15/04/2006 2-5/10/2007 12-13/12/2008 19-21/12/2008 2.2ma TNG b TNG b 2.2ma 2.2ma 2.2ma FOCES SARG SARG FOCES FOCES FOCES 2048x2048 24µm Site#1d 2048x4096 13.5µm EEV 2048x4096 13.5µm EEV 2048x2048 24µm Site#1d 2048x2048 15µm LORAL#11i 2048x2048 24µm Site#1d 3720 - 10850 4620 - 7920 4620 - 7920 3720 - 10850 3830 - 10850 3620 - 7360 100 52 52 100 96 100 0.04 - 0.13 0.07 - 0.11 0.07 - 0.11 0.04 - 0.13 0.03 - 0.07 0.04 - 0.13 0.08 - 0.35 0.07 - 0.17 0.07 - 0.17 0.08 - 0.35 0.09 - 0.26 0.08 - 0.35 19 43 14 10 2 3 a 2.2 m telescope at the German Spanish Astronomical Observatory (CAHA) (Almer´ıa, Spain). b 3.58 m Telescopio Nazionale Galileo (TNG) at Observatorio del Roque de los Muchachos (La Palma, Spain). c The spectral resolution is determined as the FWHM at the arc comparison lines ranges. 4. On the nature of the RV variations Variations in the RV peak-to-peak amplitude of up to ∼2 km s−1 were observed during all the observing runs. These variations are significantly larger than the individual measurement errors (0.10 to 0.20 km s−1) or the systematic error (0.05 km s−1), even when we consider the scatter between runs with different spec- trographs and setups. 4.1. SearchingforperiodicalsignalsonRV A Least squares periodogram (see Appendix A) reveals one very significant peak at 7.783 days (see Fig. 4a). The data set contains 91 independent RV measurements. However, many of them are clustered together within groups of a few hours. The values we use to generate the periodogram and for orbital fitting (shown in Table 3), are averaged on a nightly basis. Fig. 4b shows the em- pirical False Alarm Probability (FAP) as a function of the Power. The 7.783 days peak has a FAP of 0.35%. It is worth noting that the RV period is larger than the pho- tometric period. Nevertheless to test if the RV period could arise from rotational modulation we searched for significant frequen- cies in the data points of the photometry. There is no significant power at the RV period, and no secondary peaks are found in the aliasing frequencies of the RVs or the photometric period after the main signals are removed. To illustrate the absence of related photometric signals, we marked the RV period in Fig. 3, that shows the post-fit residuals of photometric data. In addition to this, there is no signal at photometric period in the RV data, as can be seen in Fig. 4a, that shows the RV periodogram. 4.2. Stellaractivityjitter It is well known that spurious RV variations can be induced by stellar activity, especially due to changes in the profile of spectral lines caused by the presence of active regions, the so-called stel- lar jitter (Saar & Donahue 1997, Saar 2009). The high level of activity detected in BD+20 1790, induced us at first to relate RV variations with active regions. Since we ruled out the possibility of variations due to systematic errors or any seasonal effect, the main concern was to determine if stellar activity was responsi- ble. It is widely accepted that the relationship of bisectors of the cross-correlation function (CCF) and RV is a powerful method to determine whether the RV variation may be due to stellar ac- tivity or a planetary companion (Queloz et al. 2001, Mart´ınez- 7.78 days 0.35% FAP 1% FAP 5% FAP 10 Period(days) 100 1000 FAP distribution Cumulative FAP 50 40 30 20 10 r e w o P 0 1 1 P A F 0.1 0.01 7.78 days peak 0.35% FAP 0.001 0 10 20 Power 30 40 50 Fig. 4. a. Up: Least-Squares Periodogram of the nightly av- eraged radial velocity measurements. The 7.78 days peak has a FAP of 0.35%. The dotted horizontal line illustrates a FAP lower than 1% and the dashed horizontal line a FAP lower than 5%.b. Down: Empirical False Alarm Probability as a function of the Power (red line). The gray bars illustrate the distribution of False Alarms with an arbitrary normalization used to derive the Empirical FAP. Note that the Y axis is in logarithmic scale. Fiorenzano et al. 2005). The CCF was determined by using the same procedure as for the RV case, computing it for the regions which include the photospheric lines which are more sensitive to spot presence, while excluding chromospheric lines and telluric lines. The bisector inverse slope (BIS), defined as the difference of the average values of the top and the bottom zones, was com- puted to quantify the changes in the CCF bisector shape by us- 6 M. Hern´an-Obispo et al.: Massive candidate orbiting young K5 star Table 3. Radial Velocity JD days 2452388.3341a 2452389.3513a 2452390.3670a 2453099.3573a 2453100.3692 2453101.3748 2453102.3876 2454375.6480 2454378.6804 2453331.6400 2453332.6800 2453841.4250 2454812.7429 2454813.7240 2454820.5057 2454821.5126 2454822.5483 RV (km/s) σ (km/s) 9.23 8.94 8.52 7.82 6.96 7.34 7.84 7.96 7.72 8.71 8.16 7.73 7.76 7.67 7.73 7.53 7.96 0.19 0.14 0.38 0.06 0.10 0.07 0.05 0.08 0.04 0.03 0.03 0.03 0.21 0.18 0.16 0.16 0.14 a From L´opez-Santiago 2005 Fig. 6. Periodogram for bisectors of all runs. There is no clear period for bisector variations. subsequently the spectroscopic indices) which is most likely due to rotational modulation of plage-like structure emission. As is shown in Sect. 2, all chromospheric activity indicators are in emission above the continuum, indicating a very high level of activity. To avoid the photospheric contribution to the spectral profiles we applied the spectral subtraction technique described in detail by Montes et al. (1995). This technique makes use of the program STARMOD developed at Penn State University (Barden 1985) and lately modificated by Montes et al. (1995). Also, in order to control the error and minimize the uncertain- ties, some routines of the astronomical data reduction package REDucmE 5 developed at Universidad Complutense de Madrid (Cardiel 1999) were used. In these subtracted spectra, spectro- scopic indices have been defined and computed following Saar & Fisher (2000), Kuster et al.(2003), Bonfils et al. (2007). Both, Ca ii IRT and Ca ii H & K indices were only determined for FOCES runs, due to the wavelength range coverage of the spec- trograph. To avoid contamination from telluric lines we only consider the 8662Å Ca ii IRT line. We searched for periodic signals in the spectroscopic indices by computing their Least squares periodograms. Fig. 7 shows the variation with time (or- bital phase folded in this case) for Ca ii IRT, Ca ii H & K, Hα and Hβ indices. The corresponding periodogram computed shows more noise rather than a clear signal. This result is also seen in the indices figures as a non-modulation of the activity index. As an example, Fig. 8 shows the periodogram for the Hα index. As pointed out by Walter (1994), the rotational modulation of chromospheric lines due to plages is not always detectable in very active stars. Furthermore, in this case the flares could contaminate the data, masking the actual period of variation of the indices. In order to investigate this possibility we removed the data affected by flare events. Due to the different wavelength range coverage of spectrographs, we considered only Hα and Hβ indices. For Hα index, we have found a tentative rotational modulation with a period of 2.77 days, similar to photometric period (see Fig. 9). However, the postfit residuals show in Fig. 10 that this could be a misleading signal, even pure noise. For Hβ index, no clear modulation has been found. The lack of variability of BIS and spectroscopic indices with RV 5 http://www.ucm.es/info/Astrof/software/reduceme/reduceme.html Fig. 5. Bisector velocity span vs. Radial velocity for all the observing runs. Symbols represent the different runs: stars for FOCES 04A, diamonds for SARG 04B, circles for SARG 06A and triangles for FOCES 07B. The lack of correlation indicates that RV variations are not due to stellar activity. ing the method described by Queloz et al. (2001). In choosing the span zones we avoided wings and cores of the CCF profiles, where errors of bisectors measurements are large. In Fig. 5 it can be seen that there is a lack of correlation between the BIS and RV variation for all the observing runs. This indicates that the RV variations are not due to variations in the asymmetry of the photospheric lines profile, and subsequently not due to stel- lar activity variations. The least squares periodogram of bisec- tors shows two tentative peaks around 2.8 days, and is shown in Fig. 6. We have estimated the stellar jitter from Santos et al. (2000), that takes into account the Ca ii H & K index. Assuming an av- erage value for Ca ii H & K index of about -4.2 and by using eq. [4], we derived a value for the stellar jitter of up to 10 m s−1. This stellar jitter is added in quadrature to the RV error. As an additional test we investigated the variation of stellar ac- tivity indicators, especially those that are ascribed to the pres- ence of plage-like structures on the chromosphere, like Balmer lines, Ca ii H & K and Ca ii IRT. The emission flux for these lines in active stars usually shows a periodic modulation (and M. Hern´an-Obispo et al.: Massive candidate orbiting young K5 star 7 200 150 r e w o P 100 50 0 1 10 Period(days) 100 1000 Fig. 8. Periodogram for Hα index. There is no clear period for index variations. Fig. 9. Hα index for the data without flare events. It can be seen a modulation with a period of about 2.77 days, similar to photo- metric period. Fig. 7. Spectroscopic index for chromospheric activity indica- tors, phased folded orbital period. From top to bottom: Hα (squares), Hβ (circles), Ca ii IRT (triangles) and Ca ii H & K (stars). The dashed line is indicating the quiescent state. Error bars for indices are of order of 0.001 period, and the absence of a photometric period larger than 2.8 days, strongly support the planetary companion hypothesis. Fig. 10. Postfit residuals to Hα index for no-flare data as a func- tion of the period. There is no clear period for index variations. 4.3. RVwavelengthdependence Desort et al. (2007) (hereafter D07) pointed out that the color de- pendence (with wavelength) of the RV peak-to-peak amplitude with spots can be used as a diagnostic to discriminate between stellar activity or planetary companions. Due to the contrast be- 8 M. Hern´an-Obispo et al.: Massive candidate orbiting young K5 star tween spots and the surrounding photosphere being greater in the visible than at IR wavelengths, it is expected that an attenu- ation of RV amplitude towards red wavelengths would be seen. Observationally this effect has been shown by e.g. Mart´ın et al. (2006), Hu´elamo et al. (2008) and Prato et al. (2008). If the RV variations are due to a planet, the RV amplitude should be the same at every wavelength range. We investigated a possi- ble chromatic dependence by computing the RV in two differ- ent ranges of wavelength, one for red and near-IR wavelengths (7650 to 10000 Å) and the other for blue (4300 to 4800 Å). The resulting RV peak-to-peak amplitude is 2.19±0.20 km s−1 for the near-IR range and 2.20±0.20 km s−1 for the blue range. The values differ by only 0.5% and agree within the uncertainties. Additional RV infrared follow-up can allow us to confirm this. In a forthcoming paper (Hern´an–Obispo et al. 2009d, in prep.), we will present the first results of the study of the RVs of BD+20 1790 in the near-IR range. 4.4. RVvariationbyempiricalspotsandplages? To estimate an order of magnitude of the expected RV ampli- tude due to spots, we use empirical relations derived by Saar & Donahue 1997 (hereafter SD97) and D07. These relations con- nect the RV amplitude with the spot filling factor fs and vsini. We consider both relations by D07 and SD97, because D07 re- lations take into account the spectral type and the whole spectral range (except telluric and chromospheric lines) to compute the empirical RV, whereas SD97 uses a single line and G5V spectral type. Using SD97 eq. [1], we derived an amplitude of up to 575 m s−1 and by using D07 eq [5] we similarly estimate an ampli- tude of up to 600 m s−1. As mentioned, these results are taken as a quantitative estimation. There are more effects that are not taken into account here, like the spot location at stellar surface given by the colatitude θ, and the spot temperature. SD97 eq. [1] and D07 eq. [5] considered the simple case of an equatorial spot, but SD97 assumed a T spot = 0 K and D07 assumed an spot temperature 1000 K cooler than the photosphere. The difference between the RV amplitude derived from both equations could be due to this different spot temperature. On the other hand, we can estimate the spot filling factor that could produce the RV signal of our data. We considered an aver- age semi-amplitude of 1 km s−1. The fs estimated from SD97 is therefore 23% while D07 indicates 19%. The fs measured from photometric variation is about 4%. These results indicate that the spot filling factor needed to explain the RV variation due purely to spots is not in agreement with the photometry. Saar (2003) and Saar (2009) showed significant efforts to model plage-induced RV jitter. Although the models are mostly applicable to solar-like stars, we could estimate the plage fill- ing factor fp that could produce the RV signal by using the Saar (2009) equation that connects the RV amplitude with vsini > 6 km s−1. This fp estimated is about 70%, that strongly suggets that the RV variation is not due to chromospheric plages. Fig. 11. Radial velocity amplitude variation with wavelength, computed for LO Peg profiles spots affect the line profiles. However, BD+20 1790 has a low vsini to model the photosphere by generating Doppler imaging spot maps. To carry out a realistic approximation to the prob- lem, we construct realistic spot maps by using spectra of another star with similar characteristics, LO Peg, that is widely stud- ied in the literature and its photospheric activity is well-known (Jeffries & Jewell 1993, Jeffries et al. 1994, Eibe et al. 1998, Eibe et al. 1999, Barnes et al. 2005). LO Peg is a K5V–K7Ve star, identified by Jeffries & Jewell (1993) as a member of the Local Association, with an estimated age of 20–30 Myr. Jeffries et al. (1994) determined the inclination to be 50◦. The level of activity is similar to BD+20 1790, but LO Peg is a rapid rota- tor (vsini ∼69 km s−1). The LO Peg photometry suggests a spot filling factor of up to 1.5%. Using the Doppler imaging program, DoTS (Collier Cameron 1997), and an input starspot image derived for LO Peg (Barnes et al. 2005), we generated a set of line profiles for a star with vsini = 10 km/s (i.e. matching that of BD+20 1790) over a complete rotation phase. The profiles thus contain asym- metries due to starspots from the observed LO Peg image. We used appropriate temperatures for the BD+20 1790 photosphere and estimated the spots to possess temperatures which were up to 1000 K cooler. Profiles were generated for the three differ- ent wavelengths of 4000 Å, 6717 Å and 10000 Å. The radial velocity variations were then calculated in order to estimate the relative amplitudes due to spot induced variations at each of the three wavelengths. The RV attenuation with wavelengths relative to 4000 Å is 16% at 6717 Å and about 30% at 10000 Å , as illustrated in Fig. 11. Assuming at a first approach the same fs for LO Peg and BD+20 1790, the RV signal for BD+20 1790 should be about 1.5 km s−1 at 10000 Å. However, in Hern´an–Obispo et al. 2009d (in prep.), we find only about 0.5% attenuation in the near-IR region relative to the visible, 6717 Å region. This result is an additional argument in support the existence of the planetary companion. 4.5. WhatwouldtheRVsignalbewithoutaplanet? It is important to remark that empirical relations derived by SD07 and D07 do not take into account the chromatic effect of spots on the RV signal. We therefore investigated how much RV signal would be expected in the absence of a planet, and the degree of RV attenuation with wavelength (assuming the RVs are due to cool spots). In order to quantify the attenuation if the cause of variations were spots, we try to investigate how much 4.6. RVjitterfromflares We have estimated the rate of flare occurrence as the fractional amount of the total observing time (for all runs) where a flare was detected. Thus, we get a flare frequency of occurrence of ∼ 40%. This higher rate raises the question of how much RV jitter we should expect from large flares, if any. Saar (2009) presents the first approach to this issue, concluding that RV jitter due to M. Hern´an-Obispo et al.: Massive candidate orbiting young K5 star 9 Fig. 12. Up: Hα index vs. BIS. The dashed line is indicating the quiescent state. Down: Hβ index vs. BIS. For both, it is seen that the scatter for BIS is higher when occur flare events. Error bars for the indices are of order of 0.001. flare occurrence would be non-negligible, although probably be a stochastic jitter component. Chromospheric activity indicators exhibited an enhancement at flare state, the broad emission of Balmer lines and He I D3 in emission being the most notable features (see Fig. 1). As pointed out by Saar (2009), although these lines are excluded when we measure RV, it is possible that a significant core filling in photospheric lines occurs when there is a flare event. The cause could be upper photospheric heating. Results by Houdebine (1992) state that heating is propagated down to low photospheric levels. A second related problem is the effect of large flares on BIS. While it has not been studied until now, it is expected to be more pronounced, since bisectors are more sensitive to changes in line profiles. To our knowledge it is reported here for the first time. Fig. 12 a shows the relationship of Hα index vs. BIS, where the dashed line corresponds to the quiescent state, and higher values for Hα index indicate the occurrence of a flare event. It is seen that the scatter for BIS is higher when a flare occurs. Outliers at quiescent state correspond to a low S/N rate. Similar BIS behaviour is seen in Fig. 12 b, that shows Hβ index vs. BIS. 5. Orbital solution for BD+20 1790 b We computed the orbital solution for the RV data using a stan- dard Keplerian fit with the RV period estimated by the Least squares periodogram. The fit was obtained firstly only consider- ing the FOCES data, averaged by night, in order to avoid intra- night scatter. After this, we added the SARG data to improve the fit. The results for the fit considering only FOCES data or all data from the two spectrographs were compatible within un- certainties. With the addition of RVs measured in winter 2008 Fig. 13. Radial velocity variability of BD+20 1790.a. Up: Circular orbit. b. Down: Eccentric orbit. Values marked with circle symbol represent FOCES runs except stars that represent DDT FOCES 08B run. Diamond symbol are for SARG runs. (DDT FOCES 08b run), the Least squares periodogram is strik- ingly improved, and the 7.78 day peak clearly dominates the power spectrum. Attempts to perform a Keplerian fit using the second and the third highest periodogram peaks produced sig- nificantly worse folded curves. We have included to perform the fit the RV set computed by L´opez–Santiago (2005). A first fit (see Fig. 13a) derives a close-in massive planet (a = 0.066 AU, M2 sin i = 6.54M jup) in a circular orbit (e = 0.05) with a rota- tional period of 7.7834 days and a reduced χ2 of 1.07. Also we present a second fit (see Fig. 13b) with the same period for an eccentric orbit (a = 0.066 AU, M2 sin i = 6.15M jup, e = 0.14, χ2 = 0.997). Due to the sampling of the data, we cannot discard a possible eccentric orbit. Orbital elements for both solutions are compiled in Table 4 and discussed in the next section. As additional test, we computed the orbital solution remov- ing the data affected by flare events. The fit derives a solution (a = 0.066 AU, M2 sin i = 6.54M jup, e = 0.01, K = 0.91 km s−1) compatible with the solution when considering all the data. The fit is presented in Fig. 14. 10 M. Hern´an-Obispo et al.: Massive candidate orbiting young K5 star ant planets is on average 4 times greater than those with more distant planetary companions. For the 'close-in' sub-sample, the X-ray luminosity is LX = 1028.5 erg s−1 on average. The X-ray luminosity of BD+20 1790 is 5 times brighter than this average which is consistent with chromospheric and X-ray emission in- duced by the presence of a massive close-in companion (Lanza 2009). Even though the stellar activity could swallow the RV signal of a planetary companion, we can detect it for BD+20 1790 b since it is a massive planet. The RV variation is large enough even though the RV accuracy is tipically about 150 m s−1. Due to the observational strategy (the data are not part of a planet-search program), the eccentricity is poorly constrained. Indeed there is no "a priori" reason to discard an eccentric or- bit since the circularization time-scale computed is up to several Gyr, but more data is required to properly characterize the ec- centricity. RV optical and infrared follow-up over twice the RV period will enable us to constrain the orbital solution as well as confirm the presence of the planet. More massive exoplanets M2 sin i ∼5M jup with orbital periods longer than about 6 days have eccentricities significantly larger than lower mass planets (Udry & Santos 2007). Another possibility is that additional un- detected longer period planets are maintaining the eccentricity of BD+20 1790 b. Both situations have been discussed in detail by Wu & Murray (2003). It is worth noting however that the star is metal-rich, as presented in Sec. 2. The existence of a correlation between stellar metal- licity and planet mass has been reported by e.g. Santos et al. (2001), Fischer & Valenti (2005), Guillot et al. (2006). Massive planets tend to form around metal-rich stars, i.e., planets that or- bit around metal-rich stars also have higher mass cores. Compared to other planets of similar masses and orbits6, and taking into account statistical results described in recent reviews (Udry & Santos 2007), BD+20 1790 b does not exhibit unusual characteristics, except for its young age and its relatively high mass. We used a complimentary method to determine the stel- lar age from Mamajek & Hillenbrand (2008) (hereafter MH08), that uses the fractional X-ray luminosity, RX = LX /Lbol. MH08 demonstrate that RX has the same age-inferring capability as the chromospheric index R′ HK. By using their equation [A3] we es- timated an age for BD+20 1790 of up to 35 Myr. Considering a value for logR′ HK = -4.2 on average, we can also estimate the age with the new relation proposed by MH08, by equation [3]. We computed an age of up to 58 Myr. These values are in agreement with the range estimated by L´opez-Santiago et al. (2006). Lowrance et al. (2005) included this star BD+20 1790 in a coronographic survey for substellar companions using the coronograph on NICMOS/HST and the 200 inch Hale Telescope (Palomar Osbservatory). No companions were found beyond 10 AU. However, the orbital solutions we find suggest a semi-major axis below 0.1 AU, clearly beyond their resolving capabilities. Great care must therefore be taken when extrapolating prop- erties of early stellar evolution stages from the characteristics of the latter stages, since the current knowledge about planetary system evolution is still somewhat speculative. The exoplane- tary zoo is such that new planets with unusual properties require a replanting of planet formation and migration scenarios. Planets discovered around young stars could be the missing link that re- construct the scenarios between exoplanets and protoplanetary disks. Indeed, further study of BD+20 1790 b has the potential 6 Observational data for the more than 370 exoplanets are compiled on the Extrasolar Planets Encyclopaedia (http://exoplanet.eu), man- tained by J. Scheneider Fig. 14. Radial velocity variation of BD+20 1790 computed con- sidering only the data that are not affected by flares. Circle symbol represent FOCES runs except stars that represent DDT FOCES 08B run. Diamond symbol are for SARG runs. Table 4. Orbital Parameters of BD+20 1790 b Parameter Porb T a conj a e K γ ω M2sini rms χ2 Solution 1 7.7834 ± 0.0004 3085.8 ± 0.5 0.066 ± 0.001 0.05 ± 0.02 0.93 ± 0.03 8.22 ± 0.01 200.4 ± 21.8 6.54 ± 0.57 138.9 1.071 Solution 2 7.7834 ± 0.0004 3086.30 ± 0.18 0.066 ± 0.002 0.14 ± 0.04 0.84 ± 0.06 8.12 ± 0.04 120.7 ± 14.0 6.15 ± 0.59 132.3 0.997 days HJD AU km s−1 km s−1 degrees M jup m s−1 a Time of periastron passage 6. Discussion The lack of a relation between the BIS and spectroscopic indices with the RV period, as well as the different RV and photometri- cal period strongly suggest that the RV variations are due to a planetary companion. However, it is possible that the RV varia- tions are actually due to a combination of phenomena (activity and planet). Stellar magnetic activity may be influenced and enhanced by the presence of a close-in giant planet, as proposed by Cuntz et al. (2000), Cuntz & Shkolnik (2002) and Lanza (2008). Thus, the presence of a planetary companion could be an interpretation for the high level of stellar activity detected. In a recent paper, Lanza (2009) proposes a new model that predicts the formation of prominence-like structures in very highly active stars with close-in giant planets. Also, as presented in Sec. 2 and Sec. 4.6, the large flares, with energy releases in the superflare regime, and the high rate of flare ocurrence, could find a source in addi- tion to stellar activity in the reconnection of the stellar coronal field as the planet moves inside the Alfv´en radius of the star (Ip et al. 2004). In a forthcoming paper we explore in detail these possible star-planet interactions (Hern´an–Obispo et al. 2009c, in prep). In addition, as suggested by the statistical analysis by Kashyap et al.(2008), the X-ray flux from stars with close-in gi- M. Hern´an-Obispo et al.: Massive candidate orbiting young K5 star 11 to improve our understanding of planetary systems at early evo- lutionary stages. 7. Conclusions This paper describes the investigation of RV variations for the young and active K5V star BD+20 1790. Based upon the anal- ysis of the BIS of the CCF, as well as activity indicators and photometry, the presence of a planetary companion is shown to be the best interpretation. The orbital solution results in a com- panion with a mass in the planetary regime. No photometric pe- riod larger than 2.8 days strongly supports the planetary origin of the observed RV variations. Two solutions for the orbit are com- puted and discussed. The presence of a close-in massive planet could also be an explanation for the high level of stellar activity. Since the RV data are not part of a planet search program, we can consider our results as serendipitous evidence of a planetary companion. Indeed additional RV optical and infrared follow-up will enable us to constrain the orbital solution as well as confirm the presence of the planet. This is thus far the youngest main sequence star for which a planetary candidate has been reported. Acknowledgements. We thank Calar Alto Observatory for allocation of direc- tor's discretionary time to this programme. This work was supported by the Spanish Ministerio de Educaci´on y Ciencia (MEC) under grant AYA2005-02750, Ministerio de Ciencia e Innovacin (MICINN) under grant AYA2008-06423-C03- 03 and The Comunidad de Madrid under PRICIT project S-0505/ESP-0237 (ASTROCAM). MCGO acknowledges financial support from the European Commission in the form of a Marie Curie Intra European Fellowship (PIEF- GA-2008-220679). MHO and GAE thank Dr. Chriss Moss, support astronomer at the LT for his help and patience. Also MHO thanks Dr. Santos Pedraz, sup- port astronomer at the Calar Alto Observatory for his help with DDT run. MHO is grateful to Dr. Jos´e Antonio Caballero for valuable discussions, and also Dr. Laurence R. Doyle for his suggestions that was the initial inspiration for this work. This research has made use of the SIMBAD database, operated at CDS, Strasbourg, France. The authors gratefully acknowledge the valuable comments and suggestions of an anonymous referee, that are helped to improve the paper. References Barden, S. C., 1985, ApJ, 295, 162 Barnes, J. R., Collier Cameron, A., James, D. J. and Donati, J.-F., 2000, MNRAS, 314, 162 Barnes, J. R., Cameron, A. C., et al., 2005, MNRAS, 356, 1501 Biazzo, K., Frasca, A., Henry, G. W. et al., 2007, ApJ, 656, 474 Bonfils, X., Mayor, M., Delfosse, X. et al., 2007, A&A, 474, 293 Bouvier, J., Alencar, S. H. P., Boutelier, T., et al., 2007, A&A, 463, 1017 Byrne, P. B., Eibe, M. T. and Rolleston, W. R. J., 1996, A&A, 311, 651 Cardiel, N., 1999, PhD Thesis, Universidad Complutense de Madrid Carpenter, J. M., Bouwman, J., Silverstone, M. D. et al., 2008, ApJS, 179, 423 Carroll, B. W. and Ostlie, D. A., 2007, An Introduction to Modern Astrophysics, 2nd edition, Pearson International Edition, Addison Wesley Catalano, S., Biazzo, K., Frasca, A. et al., 2002, A&A, 394, 1009 Collier Cameron, A. and Robinson, R. D., 1989a, MNRAS, 236, 57 Collier Cameron, A. and Robinson, R. D., 1989b, MNRAS, 238, 657 Collier Cameron, A. and Woods, J. A., 1992, MNRAS, 258, 360 Collier Cameron, A., 1997, MNRAS, 287, 556 Collier Cameron, A., 2001, in IAU Symposium, "Recent Insights into the Physics of the Sun and Heliosphere: Highlights from SOHO and Other Space Missions", 203, p229 Collier Cameron, A., Jardine, M. M. and Donati, J.-F., 2002, in ASP Conf. Ser., "Stellar Coronae in the Chandra and XMM-NEWTON Era", 277, p397 Cumming, A., 2004, MNRAS, 354, 1165 Cuntz, M. and Saar, S. H. and Musielak, Z. E., 2000, ApJ, 533, L151 Cuntz, M. and Shkolnik, E, 2002, 323, 387 Desort, M., Lagrange,A.-M. et al., 2007, A&A, 473, 983 Donati, J.-F., Mengel, M., Carter, B. D. et al., 2000, MNRAS, 316, 699 Eibe, M. T., 1998, A&A, 337, 757 Eibe, M. T., Byrne, P. B., Jeffries, R. D. et al., 1999, A&A, 341, 527 Esposito, M., Guenther, E., Hatzes, A. P. et al., 2006, "Tenth Anniversary of 51 Peg-b: Status of and prospects for hot Jupiter studies", p127 Ferreira, J. M., 2000, MNRAS, 316, 647 Fischer, D. A. and Valenti, J., 2005, ApJ, 622, 1102 Fleming, T. A., Schmitt, J. H. M. M. and Giampapa, M. S., 1995, ApJ, 450, 401 Frasca, A., Biazzo, K., Catalano, S. et al., 2005, A&A, 432, 647 Gratton, R. G., Bonanno, G., Bruno, P. et al, 2001, Exp. Ast., 12, 107 Gonzalez, G., Laws, C., Tyagi, S. et al., 2001, AJ, 121, 432 Guillot, T., Santos, N. C., Pont, F. et al., 2006, A&A, 453, L21 Hern´an-Obispo, M., de Castro, E., Cornide, M., 2005, ESA Special Publication, 560, 647 Hern´an-Obispo, M., de Castro, E., G´alvez, M.C., "Highlights of Spanish Astrophysics IV", 2007, Kluwer Academic Publishers Houdebine, E. R., 1992, Irish Astronomical Journal, 20, 213 Huelamo, N., Figueira, P. et al., 2008, ArXiv e-prints, 0808.2386 Huerta, M., Johns-Krull, C. M. et al., 2008, ApJ, 678, 472 Ip, W.-H., Kopp, A. and Hu, J.-H., 2004, ApJ, 602, L53 Jardine, M., Collier Cameron, A., Donati, J.-F. et al., 2001, MNRAS, 324, 201 Jardine, M. and van Ballegooijen, A. A., 2005, MNRAS, 361, 1153 Jeffries, R. D. and Jewell, S. J., 1993, MNRAS, 264, 106 Jeffries, R. D., Byrne, P. B., Doyle, J. G. et al., 1994, MNRAS, 270, 153 Jeffries, R. D. 1995, MNRAS, 273, 559 Kashyap, V. L., Drake, J. J. and Saar, S. H., 2008, ApJ, 687, 1339 Kurucz, R. L., 1993, IAU Colloq. 138: Peculiar versus Normal Phenomena in A-type and Related Stars, APCS , 44, 87 Kurster, M., Endl, M., Rouesnel, F. et al., 2003, A&A, 403, 1077 Lanza, A. F., 2008, A&A, 487, 1163 Lanza, A. F., 2009, A&A, 505, 339 L´opez-Santiago, J. 2005, PhD Thesis, Universidad Complutense de Madrid L´opez-Santiago, J., Montes, D., Crespo-Chac´on, I., et al., 2006, ApJ, 643, 1160 Lowrance, P. J., Becklin, E. E., Schneider, G., Kirkpatrick, et al., 2005, AJ, 130, 1845 Mamajek, E. E. and Hillenbrand, L. A., 2008, ApJ, 687, 1264 Mason, K. O., Hassall, B. J. M., Bromage, G. E. et al., 1995, MNRAS, 274, 1194 Mart´ın, E. L., Guenther, E., Zapatero Osorio, M. et al., 2006, ApJ, 644, L75 Mart´ınez Fiorenzano, A. F., Gratton, R. G., Desidera, S. et al., 2005, A&A, 442, 775 Mayor, M. and Queloz, D., 1995, Nature, 378, 355 Montes, D., de Castro, E., Fernandez-Figueroa, M. J. et al., 1995, A&AS, 114, 287 Norton, A. J., Wheatley, P. J., West, R. G., et al., 2007, A&A, 467, 785 Paulson, D. B., Cochran, W. D. and Hatzes, A. P., 2004, AJ, 124, 3579 Paulson, D. B. and Yelda, S., 2007, PASP, 118, 706 Pfeiffer, M. J., Frank, C., Baumueller, D. et al, 1998, A&AS, 130, 381 Prato, L., Huerta, M., Johns-Krull, C. M. et al., 2008, ApJ, 687, L103 Queloz, D., Henry, G. W., Sivan, J.P., et al., 2001, A&A, 379, 279 Reid, I. N., Cruz, K. L., Allen, P. et al., 2004, AJ, 128, 463 Rubenstein, E. P. and Schaefer, B. E., 2000, ApJ, 529, 1031 Santos, N. C., Mayor, M., Naef, D. et al., 2000, A&A, 361, 265 Santos, N. C., Israelian, G. and Mayor, M., 2001, 373, 1019 Saar, S. H. and Donahue, R. A., 1997, ApJ, 485, 319 Saar, S. H. and Fischer, D., 2000, ApJ, 534, L105 Saar, S. H., 2003, in Astronomical Society of the Pacific Conference Series "Scientific Frontiers in Research on Extrasolar Planets", 294, p65 Saar, S. H., 2009, American Institute of Physics Conference Series, 1094, 152 Scargle, J. D., 1982, ApJ, 263, 835 Setiawan, J., Weise, P., Henning, T., et al., 2007, ApJ, 660, L145 Setiawan, J., Henning, T. and Launhardt, R. et al., 2008, Nature, 451, 38 Sneden, C., 1973, ApJ, 184, 839 Steele, I. A., Smith, R. J., Rees, P. C. et al., 2005, "Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series", 5489, 679 Tonry, J. and Davis, M., 1979, AJ, 84, 1511 Udry, S., Santos, N. C., 2007, ARA&A, 45, 397 Walter, F. M., 1994, in IAU Symposium "Stellar Surface Structure", 176, p355 Wu, Y., Murray, N., 2003, ApJ, 589, 605 Appendix A: Periodogram We use a Least squares periodogram approach to identify and visually illustrate the relevant periods in the data. It differs from the more classic Lomb-Scargle periodogram (Scargle 1982) in a few key aspects. For a given period P, a linear model of the form vr = γ+A cos 2π/Pt + B cos 2π/Pt is fitted using a weighted least squares to the data and the χ2 of the residuals is obtained. The χ2 minima reveal the candidate signals of interest. One can rep- resent the root mean square of the residuals (RMS) with respect to the period to show the relevant periods as minima (This ap- proach is used to illustrate the photometric periods in Fig. 3). To 12 M. Hern´an-Obispo et al.: Massive candidate orbiting young K5 star recover a more familiar view of a periodogram, one can compute the Power of each period P as Power(P) = χ2 none = none − χ2 (χ2 P)/2 χ2 P/(nobs − 3) obs σ !2 Xi vi − hvi (A.1) (A.2) that follows a Fisher-F Distribution with 2 and nobs − 3 de- grees of freedom and can be used to obtain a first hint of the False Alarm probability of a given solution. This definition of the power measures how much the χ2 of the fit improves when a sinusoid of period P is included (see Cumming 2004 for a more detailed description). Since analytical approaches tend to give optimistic confi- dence levels, it is desirable to obtain the False Alarm Probability of a solution empirically. To make this, we generate a large num- ber of synthetic datasets (105) with the same sampling cadence (same dates) but only containing random noise. Then, for each realization, we compute the Least Squares periodogram and find the period with higher power, which will be a false alarm. A histogram of False Alarms as a function of the Power is ob- tained and its complementary cumulative distribution gives the False Alarm Probability of a given peak in our signal. The Least Squares periodogram of the RVS data and its associated empiri- cal FAP probability distribution are shown in Fig. 4. Compared to the Lomb-Scargle periodogram, this approach allows a proper weighting of each observation and can be easily generalized to include other time-dependent effects in the signal at the period search level.
1912.05889
1
1912
2019-12-12T11:45:11
HST/STIS capability for Love number measurement of WASP-121b
[ "astro-ph.EP", "astro-ph.SR" ]
Data from transit light curves, radial velocity and transit timing observations can be used to probe the interiors of exoplanets beyond the mean density, by measuring the Love numbers $h_2$ and $k_2$. The first indirect estimate of $k_2$ for an exoplanet from radial velocity and transit timing variations observations has been performed by taking advantage of the years-spanning baseline. Not a single measurement of $h_2$ has been achieved from transit light curves, mostly because the photometric precision of current observing facilities is still too low. We show that the Imaging Spectrograph instrument on-board the Hubble Space Telescope could measure $h_2$ of the hot Jupiter WASP-121b if only few more observations were gathered. We show that a careful treatment of the noise and stellar limb darkening must be carried out to achieve a measurement of $h_2$. In particular, we find that the impact of the noise modelling on the estimation of $h_2$ is stronger than the impact of the limb darkening modelling. In addition, we emphasize that the wavelet method for correlated noise analysis can mask limb brightening. Finally, using presently available data, we briefly discuss the tentative measurement of $h_2 = 1.39^{+0.71}_{-0.81}$ in terms of interior structure. Additional observations would further constrain the interior of WASP-121b and possibly provide insights on the physics of inflation. The possibility of using the approach presented here with the Hubble Space Telescope provides a bridge before the high-quality data to be returned by the James Webb Space Telescope and PLATO telescope in the coming decade.
astro-ph.EP
astro-ph
Draft version December 13, 2019 Typeset using LATEX preprint2 style in AASTeX62 9 1 0 2 c e D 2 1 . ] P E h p - o r t s a [ 1 v 9 8 8 5 0 . 2 1 9 1 : v i X r a HST/STIS capability for Love number measurement of WASP-121b Hugo Hellard,1, 2 Szil´ard Csizmadia,1 Sebastiano Padovan,1 Frank Sohl,1 and Heike Rauer1, 2, 3 1Deutsches Zentrum fur Luft und Raumfahrt, Institut fur Planetenforschung Rutherfordstrasse 2, 12489 Berlin, Germany, DE 2Technische Universitat Berlin Strasse des 17. Juni 135, 10623 Berlin, Germany, DE 3Freie Universitat Berlin, Institut fur Geologische Wissenschaften Malteserstrasse 74-100, 12249 Berlin, Germany, DE (Received; Revised; Accepted) Submitted to ApJ ABSTRACT Data from transit light curves, radial velocity and transit timing observations can be used to probe the interiors of exoplanets beyond the mean density, by measuring the Love numbers h2 and k2. The first indirect estimate of k2 for an exoplanet from radial velocity and transit timing variations observations has been performed by taking advantage of the years-spanning baseline. Not a single measurement of h2 has been achieved from transit light curves, mostly because the photometric precision of current observing facilities is still too low. We show that the Imaging Spectrograph instrument on-board the Hubble Space Telescope could measure h2 of the hot Jupiter WASP-121b if only few more observations were gathered. We show that a careful treatment of the noise and stellar limb darkening must be carried out to achieve a measurement of h2. In particular, we find that the impact of the noise modelling on the estimation of h2 is stronger than the impact of the limb darkening modelling. In addition, we empha- size that the wavelet method for correlated noise analysis can mask limb brightening. Finally, using presently available data, we briefly discuss the tentative measurement of h2 = 1.39+0.71−0.81 in terms of interior structure. Additional observations would further constrain the interior of WASP-121b and possibly provide insights on the physics of inflation. The possibility of using the approach presented here with the Hubble Space Telescope provides a bridge before the high-quality data to be returned by the James Webb Space Telescope and PLATO telescope in the coming decade. Keywords: HST photometry -- Planets and Satellites: interiors -- Planets and Satel- lites: individual (WASP-121b) Corresponding author: Hugo Hellard [email protected] 2 Hellard et al. 1. INTRODUCTION Celestial bodies, and in particular planets or- biting close to their Roche limit will deform in response to tidal interactions with their host star and, to a lower extent, to their own ro- tation. The radial deformation of the planetary outer shape can be quantified using the Love number h2, while the redistribution of mass within the planet can be expressed through the Love number k2 (Love 1911). The perturbing potential, Vp, is usually expressed as a sum of harmonics of degree n, Vp,n: Vp = Vp,n. (1) ∞(cid:88) n=2 (cid:18)Rp (cid:19)n d For example, the tidal potential of degree n, Vt,n, is equal to: Vt,n = Gms d Pn(λ), (2) where G is the gravitational constant, ms is the stellar (i.e., perturber) mass, d is the center-to- center distance, Rp is the mean planetary radius (i.e., the radius the planet would have if it were isolated and non-rotating), Pn is the Legendre polynomial of degree n, and λ = sin θ cos ϕ, with θ and ϕ the colatitude and east longitude, respectively. Likewise, the resulting potential at the planet Vr, and the total radial deformation ∆r can be expressed as: ∞(cid:88) ∞(cid:88) n=2 Vr = ∆r = Vr,n ∆rn (3) (4) n=2 The Love numbers of degree n, hn and kn are defined as (Love 1911; Kopal 1959): ∆rn = hn Vp,n g Vr,n = knVp,n, (5) (6) where g is the mean surface acceleration of the unperturbed planet. The derivation of the Love numbers depend on the planetary interior. In case of hydrostatic equilibrium (i.e., the interior behaves as a fluid), hn and kn only depend on the radial density profiles, and there is a simple relation hn = 1 + kn with 0 < kn < 1.5 (Munk & MacDonald 1960). A value of 0 would de- scribe a point-mass surrounded by a mass-less envelope, while a value of 1.5 would represent a homogeneous body. For instance, k2 (cid:39) 0.985 for the fluid Earth (Lambeck 1980), and k2 (cid:39) 0.6 for Jupiter (Iess et al. 2018). While hydrostatic equilibrium is a fair assumption for hot Jupiters - since gas behaves as a fluid - it may not be true for ice giants or rocky planets. In that case, the Love numbers depend on additional interior parameters such as rigidity and viscosity, and on the timescale of the perturbation (Correia & Rodr´ıguez 2013). In either case, the knowledge of these numbers provides additional informa- tion on planetary interiors (e.g., Baumeister et al. 2019; Kellermann et al. 2018; Padovan et al. 2018). We usually focus on the second degree Love numbers h2 and k2 because they are the most sensitive to the interior and the easiest to measure (Gavrilov & Zharkov 1977). There- fore we have 0 < h2 < 2.5, where 0 represents an infinitely rigid body (i.e., a non-deformable sphere) while 2.5 describes a homogeneous fluid body. The interior of WASP-121b is expected to be in hydrostatic equilibrium since it is mainly composed of gas, and the latter behaves as a fluid. In addition, close-in giants are ex- pected to be locked in a 1:1 spin-orbit resonance (Ragozzine & Wolf 2009), a configuration for which the linear approximation for the compu- tation of the fluid Love numbers holds (Padovan et al. 2018). In that case the theoretical range of values for h2 is between 1 and 2.5. However, non-linear effects may change the value of the Love numbers by several percents (Wahl, Hub- bard, & Militzer 2016, 2017). Accordingly, we HST/STIS capability for Love number measurement 3 opt for a less-biased wider prior for h2 which can accommodate these a priori-unknown ef- fects. We do not extend the upper bound on h2 over 2.5 because it is highly unlikely to have a gas giant close to being homogeneous, since it is mostly made of gas, which in the outer part is very compressible. It has been shown that k2 can be indirectly mea- sured from radial velocity variations (RVVs) and/or transit timing variations (TTVs) obser- vations (Csizmadia, Hellard, & Smith 2019), while h2 can be measured from transit light curve (TLC) observations (Correia 2014; Akin- sanmi et al. 2019; Hellard et al. 2019). Depend- ing on the assumed system's dynamics (e.g., stellar and planetary rotation rates, orbits), es- timates of k2 recently became available for few exoplanets (e.g., Buhler et al. 2016; Hardy et al. 2017; Bouma et al. 2019; Csizmadia, Hel- lard, & Smith 2019), while h2 has never been measured from TLCs. This is because the pho- tometric uncertainty of current telescopes is still above the required levels. The upcoming JWST (launch planned for 2021) and PLATO (launch planned for 2026) missions will provide accurate enough TLC observations to precisely measure h2 (Hellard et al. 2019). In this paper we show that, in the meantime, the Imaging Spectrograph (STIS) instrument on-board the Hubble Space Telescope (HST) has the capabil- ity to carry out Love number measurements of WASP-121b from TLCs. We prove this point by using two transits observed by HST/STIS (Section 2). In Section 3 we present the transit model. To achieve a precise measurement of h2, we show in Section 4 that a careful treatment of the noise should be performed and that the wavelet method performs well for that purpose. Furthermore, a cautious modelling of the stel- lar limb darkening (LD) should be used, and a thorough analysis of the fitted stellar intensity profile must be carried out. We conclude with a discussion on the retrieved Love number and its potential to better understand the physics of inflation (Section 5). 2. DATA We used observations of two primary tran- sits of WASP-121b made by HST/STIS, in the frame of the Panchromatic Comparative Ex- oplanet Treasury (PanCET) survey (Progam 14767; P.I.s D. Sing and M. L´opez-Morales). The first visit was made on October 24, 2016 (hereinafter Lv1) and the second one on Novem- ber, 6, 2016 (hereinafter Lv2), with the G430L grating. The data were published in Evans et al. (2018), where the reader can find further de- tails on the data reduction process. In Evans et al. (2018), two other primary transits were pub- lished (one with the G750L grating and another with the G141 grism of the Wide Field Camera 3). However, the data of these two extra tran- sits do not cover the ingress and egress phases of the transit, which are important for Love num- ber measurements (Carter & Winn 2010; Cor- reia 2014). Accordingly, we did not use these two transits in our analysis1. The raw white light curves are presented in Figure 1. One can easily identify the systematics coming from the thermal expansion and contraction of the tele- scope along each orbit, the so-called thermal breathing of HST2. The treatment of such time- correlated noise is crucial when retrieving infor- mation from TLCs (see Section 4.1). In the next section we briefly present the transit model used to retrieve the Love number h2 from the TLC. 3. TRANSIT MODEL The object WASP-121b orbits its F6V host star at a distance of roughly 0.03 AU in about 1.3 days (Delrez et al. 2016). Considering the 1 We ran an additional analysis included these two transits and did not obtain any improvements on the h2 measurement. 2 see STScI Instrument Science Report ACS 2008-03 4 Hellard et al. (cid:18) r(θ, ϕ) = Rp 1 + q 4(cid:88) n=2 − 1 3 h2(1 + q)F 2 p d (cid:18)Rp (cid:19)n+1 d (7) (cid:19) P2(cos(Θ)) , hnPn(λ) (cid:18) Rp (cid:19)3 where q is the ratio between the stellar and planetary masses, Fp is the ratio between the orbital and rotational periods, Θ is the obliq- uity. We assume a circular orbit (Delrez et al. 2016) (d is equal to the semi-major axis), a syn- chronized rotation (Fp = 1), and a rotational axis perpendicular to the orbital plane (Θ = θ). Because the sensitivity of the planetary Love numbers to the interior decreases with increas- ing degree (see, e.g., Padovan et al. 2018), we fix h3 = h4 = 1 and assume a spherical star. The integration of the occulted stellar area dur- ing transit phases result in the modelled TLC (Hellard et al. 2019). The fitted transit param- eters are the inclination (i), the epoch (E0), the limb darkening coefficients (LDCs, see Section 4.2), the normalized semi-major axis (d/Rs), the normalized planetary mean radius (Rp/Rs), and h2. 4. LIGHT CURVE ANALYSIS 4.1. Noise analysis The noise present in the TLC may be discrim- inated between uncorrelated noise (white noise) and time-correlated noise or systematics (red noise). The latter may come from stellar flares or instrumental systematics, for instance. The amount of white noise should not be lower than the expected photon noise, σph, for the consid- ered instrument, which can be approximated as (Kjeldsen & Frandsen 1992): σph = 0.011D−1∆λ−1/2∆t−1/2100.2V , (8) where D is the telescope diameter in centimeter (240 cm for HST), ∆λ is the equivalent width of Figure 1. HST/STIS raw data for Lv1 and Lv2. The systematics due to thermal breathing are clearly visible. mass estimates available in the literature (Del- rez et al. 2016), this results in a planet orbiting at 2.13 Roche radii, where the Roche radius cor- responds to the orbital distance at which tidal interactions overcome the planet's self gravity. Therefore, WASP-121b is highly deformed due to tidal interactions with its host star, and the planetary shape should not be modelled as a sphere. To further support that, we calculate a simple estimate of the radial deformation the planet experiences due to tides. Using the val- ues from Delrez et al. (2016) and assuming a Jovian-like Love number h2 = 1.6, we derive ∆r2/Rp = 0.05 with Equations (2) and (4) . A radial deformation of 5%, which translates into roughly 3600 km, is clearly not negligible. Therefore, we adopt the shape model described by Hellard et al. (2019), where the radius is given by: 3210123time since mid-transit [h]0.9800.9850.9900.9951.000stellar fluxLv1Lv2 HST/STIS capability for Love number measurement 5 the filter in nanometer (280 nm for G430L), ∆t is the exposure time in seconds (253 s), and V √ is the stellar magnitude in the V band (10.4 for WASP-121). We obtain σph = 21 ppm/ ∆ t. There are three major ways of modelling the red noise. If the source and behavior of the systematics are known, the first way consists in fitting a function which best represents that behavior. Examples include a first order poly- nomial in time, a second order polynomial in the detector position, or a fourth order polyno- mial in the HST phase. However, the system- atics' behavior is often not well - or at least not completely - understood. Therefore the second option consists in modelling the correlated noise through a Gaussian Process (GP). This method requires the selection of some kernel functions, which, in principle, should be physically moti- vated (Foreman-Mackey et al. 2017). However, there is no clear connection between the ther- mal breathing of HST and the physical param- eters on which we have a handle on. Therefore, we did not investigate the GP method. The last option corresponds to the so-called wavelet method (Carter & Winn 2009). An in-depth description of the method is beyond the scope of this paper but we summarize here the main idea (see also Cubillos et al. 2017). The red noise is assumed to have a power spectral den- sity varying as 1/f γ, where f is the frequency. The residuals are projected onto an orthonornal wavelet basis with two dimensions: scale and location (in time). This transform greatly sim- 3 Carter & Winn (2009) emphasized that while σw is the standard deviation of the white noise, σr is generally not the standard deviation of the correlated noise. plifies the calculation of the likelihood function. The method requires the selection of a wavelet basis, and Wornell (1996) and Carter & Winn (2009) recommend the fourth-order Daubechies wavelet basis, which performs well for 1/f γ noise. We further followed the recommendation of Carter & Winn (2009) and fixed γ = 1, which does not decrease the performance of the noise analysis. In summary, the wavelet method is parametrized by only two parameters: σw and σr (see footnote 3). These parameters are re- lated to the coefficients' variance of the wavelet transform, see Cubillos et al. (2017) for addi- tional details. We summarize in Table 1 the models we con- sidered to analyze the TLCs. It is important to note that each coefficient of any polynomial baseline function has to be fitted separately for the Lv1 and Lv2 data sets. Hence, we report the total number of free parameters for the noise models in the Table. Additionally, we always used the wavelet method to calculate the likeli- hood function for two reasons: (1) it may help taking into account some systematics not well modelled by the baseline functions, and (2) it allows us to consistently compare the models. Therefore, the total number of free parame- ter for the noise model is equal to 2 (wavelet method) plus twice the number of coefficients of the baseline function (the factor 2 comes from the fact that we corrected the baseline for each visit separately). 6 Hellard et al. Table 1. Transit functions considered for the analysis of the light curves. In the Table, c0 is a constant and p(xN ) denotes a N-order polynomial (without the zero-order term) in time (x = t), Hubble Space Telescope phase (x = φ), and detector position (x = X and/or x = Y ). The symbol tr denotes a noise-free transit curve, while w denotes the red noise component estimated from the wavelet method. Nf ree is the number of free parameters for the noise models. Model # model 1 Baseline function B1 = c0 + p(t1) a B2 = c0 + p(t1 + φ4 + X 1 + Y 1) b model 2 model 3 B3 = c0 · (1 + p(φ4)) · (1 + p(X 1)) · (1 + p(Y 1)) c model 4 B4 = 1 Transit function Nf ree 2×2 + 2 2×8 + 2 2×7 + 2 tr · B1 + w tr · B2 + w tr · B3 + w tr · B4 + w 2 a(Evans et al. 2018) b(Alexoudi et al. 2018) c(Wakeford et al. 2018) We used a Differential Evolution Markov Chain (DE-MC) to explore the parameter space, and we minimize the wavelet-based likelihood function to obtain the posterior distributions of all fitted parameters. We used the Python library MC3 (Cubillos et al. 2017) with 12 par- allel chains, each consisting of roughly 170,000 steps (2 million steps in total). We applied a to- tal burn-in period of 140,000 steps, and checked that convergence was reached by ensuring a Gelman-Rubin test value smaller than 1.01 for every parameter (Gelman & Rubin 1992). As mentioned in Akinsanmi et al. (2019) and Hellard et al. (2019), there is a strong corre- lation between the planetary mean radius and h2, and we have no prior information on these parameters. One way to refine the Love number estimation is to perform a two-step analysis: the first step uses a DE-MC with uniform priors on the planetary mean radius and h2, until a sta- tionary distribution is reached. The second step uses a DE-MC where the prior on the LDCs and planetary mean radius are the posterior distri- butions from the first step. The final distribu- tions are used for parameter estimations. This two-step analysis allows the DE-MC to be more efficient, and provides refined parameter esti- mation. It was first suggested by Dunkley et al. (2005) and was used in various studies later on (e.g., Umetsu & Diemer 2018; Shin et al. 2019). A careful check of the Bayesian Information Criterion (BIC, see Equation 9), distribution of the residuals, and fitted white noise value are carried out to ensure that the refinement step best explains the data. We summarize the change in priors in Table 2, while the BIC is defined as (Kass & Raftery 1995): BIC = χ2 + k ln(n), (9) where χ2 is the usual chi-squared of the fit, k is the number of model parameters, and n the number of data points. The following results are always the outputs from the second step and we emphasize, if rele- vant, when the refinement step decreases the fit quality (i.e., when BIC2 > BIC1). We present in Figures 2 - 5 the systematics-free data, best fit, residuals, and distribution of the residuals for each models (see Table 1). HST/STIS capability for Love number measurement 7 Figure 2. Systematics-free data and best fit (top left panel), residuals with white noise errorbars (bottom left panel), and distribution of the residuals (right panel) for Model 1. stdres corresponds to the standard deviation of the residuals, while the orange solid line in the right panel corresponds to the Gaussian fit of the residuals' distribution. Figure 3. Same as Figure 2 but for Model 2. Model 1: It is clear from Figure 2 that some periodic systematics remain in the residuals. Furthermore, the distribution of the residuals is far from being Gaussian, and BIC2 > BIC1. Therefore, the first-order polynomial in time does not properly describe the noise behavior in this data set. Model 2: From Figure 3 the residuals do not exhibit any periodic behavior anymore, but somehow the in-transit phases are better fit- ted than the out-of-transit phases. Further- more, the distribution of the residuals is still not Gaussian. Overall, Model 2 describes bet- ter the noise behavior than Model 1, and this is reflected in a lower BIC despite the higher number of parameters. Model 3: From Figure 4 it is clear that the dis- tribution of the residuals is not Gaussian, and that the best fit does not describe well the data. Additionally we have BIC2 > BIC1. Model 3 does not properly describe the noise behavior in this data set. Model 4: This assumption-free model is the best, as shown in Figure 5. There is no periodic behavior in the residuals, and the distribution 0.9850.9900.9951.000stellar fluxw=137±90 ppmstdres=94.97 ppmBIC2 = 125.61 > BIC1Lv1Lv2best fit3210123time since mid-transit [h]5000500residuals [ppm]3002001000100200300residuals [ppm]0123456probability density×1030.9850.9900.9951.000stellar fluxw=75±72 ppmstdres=47.76 ppmBIC2 = 122.00Lv1Lv2best fit3210123time since mid-transit [h]2500250residuals [ppm]15010050050100150residuals [ppm]012345678probability density×103 8 Hellard et al. Figure 4. Same as Figure 2 but for Model 3. Figure 5. Same as Figure 2 but for Model 4. Table 2. Summary of the priors in the two-step analysis. In the Table, post. dist. denotes posterior distribution. The parameters not mentioned in this Table have unchanged, uniform priors. We refer the reader to Section 4.2 for an extensive discussion on LD. Step 1 Step 2 Parameter up = u1 + u2 N (0.6, 0.1) a post. dist. step 1 um = u1 − u2 N (0.4, 0.1) a post. dist. step 1 post. dist. fit 1 U(0.05, 0.2) Rp/Rs h2 U(0, 2.5) U(0, 2.5) a(Evans et al. 2018) of the residuals is well approximated by a Gaus- sian. Two outliers appear in the residual time series, but an analysis without these two points did not lead to a significantly better result. Most importantly, the BIC is roughly twice as low as Model 2, because the noise model con- tains only two parameters. We conclude from this analysis that the wavelet method alone best describes the noise behavior in the TLC, without overfitting the data. With only two parameters, it prevents the fitting of complicated baseline models with multiple pa- rameters. This is of particular interest when the noise sources are not well - or not fully - understood, as it is often the case. 0.9850.9900.9951.000stellar fluxw=105±55 ppmstdres=110.11 ppmBIC2 = 191.73 > BIC1Lv1Lv2best fit3210123time since mid-transit [h]2500250residuals [ppm]3002001000100200300residuals [ppm]0.00.51.01.52.02.53.03.5probability density×1030.9850.9900.9951.000stellar fluxw=148±85 ppmstdres=59.63 ppmBIC2 = 65.60Lv1Lv2best fit3210123time since mid-transit [h]2500250residuals [ppm]2001000100200residuals [ppm]012345678probability density×103 HST/STIS capability for Love number measurement 9 Table 3. Summary of the de- rived BIC and Love number val- ues, h2, for each noise model. Model # BIC derived h2 1.35+0.71−0.79 model 1 0.95+0.73−0.60 model 2 1.31+0.70−0.77 model 3 1.39+0.71−0.81 model 4 125.61 122.0 191.73 65.60 In Table 3 we summarize the derived BIC and Love number values for each noise model. The errorbars present similar amplitudes, regard- less of the model. Although all peak values are within their 1-sigma uncertainty, their differ- ences range from 0.04 (between models 1 and 4) to 0.44 (between models 2 and 4), which would drastically change the interpretation of the planetary interior. Surprisingly, the second best model (model 2) presents the less consis- tent peak value with the best model (model 4); and a peak value lower than 1 suggesting non-linear effects taking place in the interior. Therefore, the impact of the noise model in the estimation of h2 remains high, though we showed that the wavelet method provides the most reliable result. In the next subsection we perform a limb dark- ening study, and show that a careful analysis must be carried out when combining the wavelet method with LD. In particular, we show that the wavelet method can mask limb brightening. 4.2. Limb darkening study In this subsection the model used to fit the data is Model 4 from Table 1. We test several LD laws, in addition to the quadratic one de- fined as (Kopal 1950): I(µ) I0 = 1 − u1(1 − µ) − u2(1 − µ)2, (10) where µ = cos(γ), with γ the angle between the direction to the observer and the normal to the stellar surface. I0 is the normal emergent intensity at the stellar center. The logarithmic law is defined as (Klinglesmith & Sobieski 1970): I(µ) I0 = 1 − u1(1 − µ) − u2µln(µ). (11) The square-root law is defined as (D´ıaz-Cordov´ez & Gim´enez 1992): = 1 − u1(1 − µ) − u2(1 − √ µ). (12) I(µ) I0 The power-2 law is defined as (Hestroffer 1997; Maxted 2018): = 1 − c(1 − µα). I(µ) I0 (13) The three-parameter law is defined as (Sing et al. 2009): ui(1 − µi/2), (14) = 1 − i=4(cid:88) i=2 I(µ) I0 = 1 − i=4(cid:88) i=1 I(µ) I0 while the four parameter law is defined as (Claret 2000): ui(1 − µi/2). (15) As described in Section 4.1, we perform two fits by updating the priors on the LDCs and plan- etary mean radius. For the quadratic LD law we have estimated values for the LDCs in the G430L waveband from the literature (Evans et al. 2018). However, no estimated values for any other LD law are available in this specific wave- band and/or for these stellar parameters (i.e., effective temperature and metallicity). From a theoretical point of view, some tables are avail- able in the literature for several LD law, mostly for other wavebands relevant to the CoRoT, Ke- pler, and TESS instruments. The Python pack- age ldtk (Parviainen & Aigrain 2015) allows 10 Hellard et al. the user to compute theoretical model-specific LDCs using the PHOENIX library (Husser et al. 2013). We find that even for the quadratic law, the theoretical LDCs are far from the fit- ted values. Hence, we did not use any priors from this package. Additionally, it is not clear what the bounds for the LDCs are, especially for the three- and four-parameter laws. Thus one has to ensure the fitted stellar intensity profile is nowhere negative or does not present limb brightening. Although the latter effect can appear in the presence of bright spots or plages (Csizmadia et al. 2013), quiet stars are more appropriate for such fine TLCs analyses. Therefore, we wish to avoid limb brightening by target selection. We show in Figures 6 - 11 the results for each LD law, in the same fashion as in Figures 2 - 5, but we add the best-fitted stellar intensity profile and compare it with theoretical expec- tations for other observing facilities. The latter were chosen based on their wavelength cover- age (as consistent as possible with the G430L grating) and on the availability of LDCs tables (King 2010; Claret & Bloemen 2011; Claret 2017; Maxted 2018). For the 3-parameter and 4-parameter laws, only the output from step 1 (see Table 2) is shown, see below for details. Quadratic LD: Same as Model 4 in the pre- vious section. In addition, the stellar intensity profile is in agreement with theoretical calcula- tions for other observing facilities. Logarithmic LD: The standard deviation of the residuals increases and the Gaussianity of its distribution decreases. Furthermore we have BIC2 > BIC1. Although the best fitted stellar intensity profile agrees with theoretical calcu- lations, we conclude that the stellar LD is not properly modelled by a logarithmic law in this data set. Square-root LD: The standard deviation of the residuals and derived BIC are smaller than the ones obtained with a quadratic LD law. The stellar intensity profile fits to theoretical expectations in other wavebands. However, the Gaussianity of the residual distribution de- creases. Although the derived h2 agrees with the quadratic LD law, the best-fit value is at the edge of the 1σ confidence interval. Therefore, we do not favour the square-root law. power-2 LD: As for the square-root law, the standard deviation of the residuals and BIC are slightly smaller than the ones obtained with the quadratic LD law. Furthermore, the stellar intensity profile fits the expectations in other wavebands. However, the Gaussianity of the residual distribution slightly decreases. Al- though the fitted h2 agrees with the quadratic law, its precision somewhat decreases because the LDCs are not well constrained. Conse- quently we decide not to favour the power-2 law, despite the slight improvement in the fit. We acknowledge that this law may best explain the stellar LD in this data set, only with better constrained LDCs. 3-parameter LD: Despite a lower standard deviation of the residuals and BIC value than the ones derived for the quadratic LD law, it is striking that the transit shape is unusual. The first indication of a low-quality model comes from the distribution of the residuals, which happens to be far from Gaussian. The main scientific information indicating a wrong mod- elling lies in the stellar intensity profile, which exhibits limb brightening close to the stellar center and toward the edge of the stellar disk. 4-parameter LD: The exact same comments emphasized for the 3-parameter LD law hold for the 4-parameter law. The standard devia- tion of the residuals and the BIC value are even smaller. However, the stellar intensity profile shows a limb brightening close to the stellar center and toward the edge of the stellar disk. HST/STIS capability for Love number measurement 11 Figure 6. Systematics-free data and best fit (top left panel), residuals with white noise errorbars (bottom left panel), distribution of the residuals (middle panel), and stellar intensity profiles (right panel) for the quadratic LD law. stdres corresponds to the standard deviation of the residuals, while the orange solid line in the middle panel corresponds to the Gaussian fit of the residuals' distribution. Figure 7. Same as Figure 6 but for the logarithmic LD law. Figure 8. Same as Figure 6 but for the square-root LD law. In Table 4 we summarize the derived BIC and Love number values for each LD law. The peak values remain consistent with each other, with differences ranging from 0.02 (between the quadratic, logarithmic, and square-root laws) to 0.05 (between the logarithmic, square-root, and power-2 laws). The peak value of the second best law (power-2, 1.36) is only lower than the one of the best law (quadratic, 1.39) by 0.03. Therefore, it appears that the impact of the limb-darkening modelling in the determination of h2 is weaker compared to the impact of the systematics' modelling (see Section 4.1). 0.9850.9900.9951.000stellar fluxw=148±85 ppmstdres=59.63 ppmBIC2 = 65.60Lv1Lv2best fit3210123time since mid-transit [h]2500250residuals [ppm]2001000100200residuals [ppm]012345678probability density×1030.00.20.40.60.81.00.40.50.60.70.80.91.0stellar intensitybest fit (HST/STIS)CoRoTKeplerTESS0.9850.9900.9951.000stellar fluxw=149±85 ppmstdres=76.81 ppmBIC2 = 84.12 > BIC1Lv1Lv2best fit3210123time since mid-transit [h]5000500residuals [ppm]3002001000100200300residuals [ppm]012345678probability density×1030.00.20.40.60.81.00.20.40.60.81.0stellar intensitybest fit (HST/STIS)CoRoTKeplerTESS0.9850.9900.9951.000stellar fluxw=150±85 ppmstdres=57.79 ppmBIC2 = 63.86Lv1Lv2best fit3210123time since mid-transit [h]2500250residuals [ppm]2001000100200residuals [ppm]0.00.20.40.60.81.01.2probability density×1020.00.20.40.60.81.00.20.40.60.81.0stellar intensitybest fit (HST/STIS)CoRoTKeplerTESS 12 Hellard et al. Figure 9. Same as Figure 6 but for the power-2 LD law. Figure 10. Same as Figure 6 but for the 3-parameter LD law. Output from step 1. Figure 11. Same as Figure 6 but for the 4-parameter LD law. Output from step 1. We showed that in the presence of poorly con- strained LDCs, one has to be very cautious when using the wavelet method for correlated noise analysis. This method may mask limb brightening while still providing a reasonable light curve. Hence, one must verify the distri- bution of the residuals, and most importantly, the fitted stellar intensity profile. We found that a quadratic law best describes the stel- lar LD in this data set. This is in agreement with most transit light curve analyses, since quadratic LDCs are the most studied, hence the most constrained. The community should strive to improve our understanding of stellar LD as it plays a crucial role in transit curve analysis in general, and in the precision of Love number measurements in particular. 5. DISCUSSION 0.9850.9900.9951.000stellar fluxw=149±85 ppmstdres=55.85 ppmBIC2 = 62.11Lv1Lv2best fit3210123time since mid-transit [h]2500250residuals [ppm]2001000100200residuals [ppm]0.00.20.40.60.81.0probability density×1020.00.20.40.60.81.00.20.40.60.81.0stellar intensitybest fit (HST/STIS)CoRoTKeplerTESS0.9850.9900.9951.000stellar fluxw=147±85 ppmstdres=51.19 ppmBIC1 = 62.31Lv1Lv2best fit3210123time since mid-transit [h]2500250residuals [ppm]2001000100200residuals [ppm]0.00.20.40.60.81.01.2probability density×1020.00.20.40.60.81.00.40.60.81.01.21.41.61.8stellar intensitybest fit (HST/STIS)CoRoTKepler0.9850.9900.9951.000stellar fluxw=149±85 ppmstdres=37.48 ppmBIC1 = 56.86Lv1Lv2best fit3210123time since mid-transit [h]2500250residuals [ppm]15010050050100150residuals [ppm]0.000.250.500.751.001.251.501.75probability density×1020.00.20.40.60.81.00.51.01.52.0stellar intensitybest fit (HST/STIS)CoRoTKepler HST/STIS capability for Love number measurement 13 Table 4. Summary of the derived BIC and Love number values, h2, for each LD law. power-2 LD law BIC derived h2 1.39+0.71−0.81 1.41+0.71−0.82 1.41+0.71−0.82 1.36+0.73−0.82 quadratic logarithmic square-root 65.60 84.12 63.86 62.11 3-parameter 62.31 a 4-parameter 56.86 a aBIC output from step 1. These LD laws present nonphysical limb brightening, therefore we do not re- port the derived Love number. -- -- Section 4.1, the white noise level reached af- ter N complete observed transits, per minute exposure, σN , is given by: (cid:114) (cid:114) 1 σN = σw ∆t . (16) N √ 1 minute √ ∆t and Equation (16), Using σw = 148ppm/ we deduce that 12 complete observed transits are needed to reach 90 ppm/ 1 minute. While this is valid for the G430L grating, the required number of observed transits could decrease by using a wider wavelength coverage (e.g., the G141 grism of the Wield Field Camera 3). This proves that the Hubble Space Telescope has the capability to carry out Love number measure- ments. The analysis presented in the previous Sec- tion leads to a tentative 2-sigma detection of the Love number, h2 = 1.39+0.71−0.81, and the poste- rior distributions of the transit parameters are summarized in Figure 12. Even though the un- certainty is high, it is essential to note that the peak of the distribution is far from 0, a value corresponding to a spherical body. Addition- ally, the peak is not far from the latest Jovian measurement of 1.62 (Iess et al. 2018), and it matches the measurement for Saturn of 1.39 (Lainey et al. 2017). We performed two addi- tional analyses where we either set 1.0 as h2's lower bound (as the interior of the hot Jupiter WASP-121b is expected to be in hydrostatic equilibrium) or extended the bounds on h2 to 0 < h2 < 3.5 (to accommodate unexpected distortions) but found no improvement in the estimation of h2. √ Hellard et al. (2019) showed that a maximal 1 minute is re- white noise level of 90 ppm/ √ quired to reliably retrieve h2. We calculated a white noise level of σw = 148 ± 85 ppm/ ∆t for one complete observed transit. Assuming the systematics can be removed as showed in Theoretical models for homogeneous H-He plan- ets indicate that an object with the mass of WASP-121b would have a radius similar to Jupiter (e.g., Fortney et al. 2007). The mea- sured 1.9 RJup indicates beyond doubt that WASP-121b is an inflated hot-Jupiter. Sev- eral mechanisms have been proposed to provide a suitable energy source for the inflation of hot- Jupiters, with Ohmic dissipation (Batygin & Stevenson 2010) being possibly the most likely (Thorngren & Fortney 2018). The interior den- sity profile of a non-inflated planet with the mass of WASP 121b would be relatively well approximated by a polytropic relation of index n = 1 (Guillot 2005), which has a fluid Love number h2 of about 1.52 (using the Matrix- propagator approach, see Padovan et al. 2018). The lower central value of 1.39 inferred in this study provides a further confirmation of the inflated nature of WASP-121b. Given the de- pendence of h2 on the density profile in the interior (e.g., Padovan et al. 2018), and the different locations of the energy deposition for different inflation mechanisms (e.g., Komacek, & Youdin 2017), improving the error bars on fu- ture determinations of h2 of hot-Jupiters, using 14 Hellard et al. Figure 12. Posterior distributions and correlation plots of the transit parameters (noise Model 4 and quadratic LD law, see Section 4). Dotted lines represent the mean and 68% interval confidence, while the plain blue line corresponds to the best fit. the method proposed here, will provide a direct way to better constrain the physics of inflation. 6. CONCLUSIONS Using two primary transits of WASP-121b precisely observed by the Imaging Spectro- graph on-board the Hubble Space Telescope, we tested the instrument capability for Love number measurements. We first performed a thorough noise analysis and showed that the wavelet method alone best describes the time- correlated behaviour of the noise, without the need to fit complicated baseline functions with many parameters. Second, we performed a stel- lar limb darkening study and found that it is best described by a quadratic law. Most im- portantly we emphasized that, in some cases, the wavelet method can mask limb brightening when the limb darkening coefficients are poorly constrained. Therefore we strongly encourage the community to further study stellar limb darkening processes, as it plays a crucial role in 0.0030.0020.0010.0000.001E00.300.450.600.75up0.00.20.40.6umParameterValue & Uncertaintyi[deg]86.82+1.941.94E0[JD]2456635.70742+0.00040.0004up=u1+u20.56+0.070.07um=u1u20.35+0.100.10dRs3.72+0.070.12RpRs0.1316+0.00340.0033h21.39+0.710.813.23.43.63.8dRs0.1260.1320.1380.144RpRs80.082.585.087.590.0i0.51.01.52.02.5h20.0030.0020.0010.0000.001E00.300.450.600.75up0.00.20.40.6um3.23.43.63.8dRs0.1260.1320.1380.144RpRs0.51.01.52.02.5h2 HST/STIS capability for Love number measurement 15 fine transit modelling. Our analysis provided a tentative 2-sigma detection of the Love number: h2 = 1.39+0.71−0.81, in which the impact of the noise modelling in its determination was found to be stronger than the impact of the limb darken- ing modelling. We showed that a total of 12 complete observed transits are required to pre- cisely estimate h2 with the G430L grating. This number could decrease by increasing the wave- length coverage, e.g., using the G141 grism of the Wide Field Camera 3. A precise estimate of h2 would refine the constraints on the interior of WASP-121b and provide direct insights on the physics of inflation. We acknowledge support from the DFG via the Research Unit FOR 2440 Matter under planetary interior conditions. Sz.Cs. thanks the Hungarian National Research, Development and Innovation Office for the NKFI-KH-130372 grants. We warmly thank Thomas M. Evans for kindly sharing his Hubble Space Telescope observations of WASP-121b. REFERENCES Akinsanmi, B., Barros, S. C. C., Santos, N. C., et Evans, T. M., Sing, D. K., Goyal, J. M., et al. al. 2019, A&A, 621, A117 2018, AJ, 156, 283 Alexoudi, X., Mallonn, M., von Essen, C., et al. Foreman-Mackey, D., Agol, E., Ambikasaran, S., 2018, A&A, 620, A142 & Angus, R. 2017, AJ, 154, 220 Baumeister, P., Padovan, S., Tosi, N., et al. 2019, Fortney, J. J., Marley, M. S., & Barnes, J. W. ApJ, accepted. https://arxiv.org/abs/1911.12745 Batygin, K., & Stevenson, D. J. 2010, ApJL, 714, L238 Bouma, L. G., Winn, J. N., Baxter, C., et al. 2019, AJ, 157, 217 Buhler, B. P., Knutson, H. A., Batygin, K., et al. 2016, ApJ, 821, 26 Carter, J. A., & Winn, J. N. 2009, ApJ, 704, 51C Carter, J. A., & Winn, J. N. 2010, ApJ, 709, 1219 Claret, A. 2000, A&A, 363, 1081 Claret, A., & Bloemen, S. 2011, A&A, 529, A75 Claret, A. 2017, A&A, 600, A30 Correia, A. C. M. 2014, A&A, 570, L5 Correia, A. C. M. & Rodr´ıguez, A. 2013, ApJ, 767, 128 Csizmadia, Sz., Pasternacki, Th., Dreyer, C., et al. 2013, A&A, 549, A9 Csizmadia, Sz., Hellard, H., & Smith, A, M. S. 2019, A&A, 623, A45 Cubillos, P., Harrington, J., Loredo, T. J., et al. 2017, AJ, 153, 3 Delrez, L., Santerne, A., Almenara, J.-M., et al. 2016, MNRAS, 458, 4025 2007, ApJ, 659, 1661 Gavrilov, S. V., & Zharkov, V. N. 1977, Icarus, 32, 443 Gelman, A., & Rubin, D. B. 1992, StaSc, 7, 457 Guillot, T. 2005, Ann. Rev. Earth Planet. Sci., 33, 493 Hardy, R. A., Harrington, J., Hardin, M. R., et al. 2017, ApJ, 836, 143 Hellard, H., Csizmadia, Sz., Padovan, S., et al. 2019, ApJ, 878, 119 Hestroffer, D. 1997, A&A, 327, 199 Husser, T. O., Wende-von Berg, S., Dreizler, S., et al. 2013, A&A, 553, A6 Iess, L., Folkner, W. M., Durante, D., et al. 2018, Nature, 555, 220 Kass, R. E., & Raftery, A. E. 1995, J. Am. Stat. Assoc., 90, 430 Kellermann, C., Becker, A., & Redmer, R. 2018, A&A, 615, A39 King, D. K. 2010, A&A, 510, A21 Kjeldsen, H., & Frandsen, S. 1992, PASP, 104, 413 Klinglesmith, D. A., & Sobieski, S. 1970, AJ, 75, 175 D´ıaz-Cordov´ez, J., & Gim´enez, A. 1992, A&A, Komacek, T. D., & Youdin, A. N. 2017, ApJ, 844, 259, 227 94 Dunkley, J., Bucher, M., Ferreira, P. G., et al. Kopal, Z. 1950, Harvard College Observatory 2005, MNRAS, 356, 3 Circular, 454, 1 16 Hellard et al. Kopal, Z. 1959, Close Binary Systems (New York: Parviainen, H., & Aigrain, S. 2015, MNRAS, 453, Wiley) Lainey, V., Jacobson, R. A., Tajeddine, R., et al. 2017, Icarus, 281, 286 Lambeck, K. 1980, The Earth's Variable Rotation: Geophysical Causes and Consequences (Cambridge: Cambridge Univ. Press) Love, A. E. H. 1911, Some Problems of Geodynamics (Cambridge: Cambridge Univ. Press) 4 Ragozzine, D. & Wolf, A. S. 2009, ApJ, 698, 1778 Shin, I.-G., Ryu, Y.-H., Yee, J. C., et al. 2019, AJ, 157, 146 Sing, D. K., et al. 2009, A&A, 505, 891 Thorngren, D. P., & Fortney, J. J. 2018, AJ, 155, 214 Umetsu, K., & Diemer, B. 2018, ApJ, 836, 231 Wahl, S. M., Hubbard, W. B., & Militzer, B. 2016, ApJ, 831, 14 Wahl, S. M., Hubbard, W. B., & Militzer, B. Maxted, P. F. L. 2018, A&A, 616, A39 2016, Icarus, 282, 183 Munk, W. H. & MacDonald, G. J. F. 1960, The rotation of the Earth; a geophysical discussion (Cambridge: Cambridge Univ. Press) Padovan, S., Spohn, T., Baumeister, P., et al. 2018, A&A, 620, A178 Wakeford, H. R., Sing, D. K., Deming, D., et al. 2018, AJ, 155, 29 Wornell, G. W. 1996, Signal Processing with Fractals: A Wavelet-based Approach (Upper Saddle River, NK: Prentice Hall PTR)
1211.7121
2
1211
2012-12-04T17:52:42
A proposal for community driven and decentralized astronomical databases and the Open Exoplanet Catalogue
[ "astro-ph.EP", "astro-ph.IM", "physics.comp-ph", "physics.data-an" ]
I present a new kind of astronomical database based on small text files and a distributed version control system. This encourages the community to work collaboratively. It creates a decentralized, completely open and democratic way of managing small to medium sized heterogeneous astronomical databases and catalogues. The use of the XML file format allows an easy to parse and read, yet dynamic and extendable database structure. The Open Exoplanet Catalogue is based on these principles and presented as an example. It is a catalogue of all discovered extra-solar planets. It is the only catalogue that can correctly represent the orbital structure of planets in arbitrary binary, triple and quadruple star systems, as well as orphan planets.
astro-ph.EP
astro-ph
A PROPOSAL FOR COMMUNITY DRIVEN AND DECENTRALIZED ASTRONOMICAL DATABASES AND THE OPEN EXOPLANET CATALOGUE Hanno Rein Institute for Advanced Study, Princeton, NJ 08540 October 14, 2018 ABSTRACT I present a new kind of astronomical database based on small text files and a distributed version control system. This encourages the community to work collaboratively. It creates a decentralized, completely open and democratic way of managing small to medium sized heterogeneous astronomical databases and catalogues. The use of the XML file format allows an easy to parse and read, yet dynamic and extendable database structure. The Open Exoplanet Catalogue is based on these principles and presented as an example. It is a catalogue of all discovered extra-solar planets. It is the only catalogue that can correctly represent the orbital structure of planets in arbitrary binary, triple and quadruple star systems, as well as orphan planets. 1. INTRODUCTION Astronomical catalogues and databases are almost as old as astronomy itself. Lists of star names and constellations were compiled by many ancient civilizations including the Chinese, Persians, Babylonians, Arabs and Greeks. Our technology has changed dramatically since then. However, the basic prin- ciples of how these catalogues and databases are compiled and maintained has not changed much at all. This proposal presents a new way to manage astronomical databases using modern technology. In recent years, the internet allowed the open source com- munity to make extensive use of distributed version control systems. Almost all of the software that we use every day was developed using these systems. This includes astrophys- ical projects like the pencil code1. But perhaps the most prominent example is the Linux kernel. The Linux kernel has over 15 million lines of code and an average of 3500 lines are added each day2. The number of active contributors varies a lot from project to project. Most projects involve only a hand- ful of developers. However, over 1000 developers contribute to a new release of the Linux kernel3. Linus Torvalds is the official maintainer of the project and releases each final ver- sion. However, his contributions to the kernel itself do not even give him a place in the top 30. At first sight, it may seem surprising that such a large group of individuals can work together to achieve a common goal. It might be even more surprising to claim that a similar ap- proach could benefit astronomy. But there are many similari- ties between the astronomy and open source community. We are both dedicated to keeping our results as open and easily accessible as possible. We both care about security, reliabil- ity and credibility. We are both large communities with many people having diverse backgrounds and needs. I therefore pro- pose to adopt parts of this workflow to astronomical databases and catalogues. In the following I will list the main advantages of this ap- proach. The two main ideas are using human-readable XML files to store the data and using git as a version control sys- tem to manage the database. I discuss technical details regard- ing those two ideas in Sections 3 and 4. Although I mention 1 http://www.nordita.org/pencil-code/ 2 Statistic adopted from http://royal.pingdom.com/2012/04/16/ linux-kernel-development-numbers/. 3 Number taken from http://lwn.net/Articles/395961/. specific technologies, the ideas are very general and do not rely on either XML or git. Finally, I will present one exam- ple of such a database that I have already been implementing, the Open Exoplanet Catalogue. 2. MAIN ADVANTAGES There are many advantages to using an open, decentralized approach when managing astronomical databases and cata- logues. In my opinion, the main advantages are the following: Community driven: Anyone can contribute data to the cat- alogue. This includes the original observer who col- lected the data, other astronomers who spot a typo or even interested members of the general public who might want to help4. Democratic principles: The maintainer of the database is re- sponsible for including updates into the main database. This person checks the validity of the submissions and makes sure only those that he thinks are credible are added. However, sometimes a person might disagree with the maintainer's decisions. That person can then simply fork5 the project. This allows him to stay up-to-date with the latest updates to the main database. But at the same time he can add his own data which he thinks is credible but the original maintainer disagrees. If the maintainer later decides that it was indeed a credible addition, he can just merge the new fork back into the main database. This is an important point. Let me explain this with a analogy to the Linux kernel. In that case most people trust Linus Torvalds and will use his version (the official version). They do not know if they can trust the other maintainers of the more than 1000 forks of the Linux kernel that are available on github6. However, in certain cases you might still want to have a look at one of these forks. For example when you want to run Linux on 4 One cannot underestimate the enthusiasm of people to do something like this. Just look at wikipedia and the citizen science project zooniverse. 5 A fork is a copy of the database. It also copies the entire history of the database. 6 Github is a free hosting platform for projects based on the git version control system. It is used by many open source projects including the Linux kernel. 2 a new Laptop that needs a new driver which has not been included in Torvalds' version yet. A very similar workflow might occur in astronomy, for example when a new extra-solar planet has been announced but has not been confirmed yet. Decentralized structure: The concept of forks and the re- sulting decentralized structure have many more advan- tages. Suppose the original maintainer stops maintain- ing the database because he has left the field, is on holi- day or simply doesn't have time anymore. Other mem- bers of the community that rely on this database can easily continue updating it without anyone having to formally give up control over the original version. Everyone can copy the full database including the entire history of changes to a local machine. One can analyze, edit, delete and add data without an internet connection. There is no need to host and maintain a server (of course this is still possible). The decentralized structure is furthermore a natural way to back up data. Let us once again take the Linux kernel as an example. The project has probably been forked by millions of people. No matter how many servers fail or get hacked, there will always be one copy that is still intact. Furthermore, it is even possible to continue working together on the project while a major server is not accessible. Openness: Anyone can see the entire database structure and all available data. There is no web interface required that would hide important parts of the data. This also speeds up the development of the database because it does not involve programming a user interface. Of course a web interface can always be implemented at a later time to allow simplified access to the data. Correctness and privacy: Many catalogues contain data from several different groups. Using a version control system to manage the database allows anyone to con- tribute their own data and make corrections to existing data. This reduces the number of errors compared to catalogues which are compiled by a single third party. Observers can enter and verify new and confidential data on a local copy of the database. The update can be pushed to the main repository at the same time an announcement is made or an embargo lifted. Easily extendable: Separating a database into small text files If the makes reading and maintaining it very easy. database in written in the XML file format, adding a new field to the database is trivial. One only needs to change those files for which new data exists. All the other files remain unchanged. Any existing script that parses the database will be fully backwards compatible without any additional work. Keeping track of history: The entire history of each change to the database is recorded in the version control sys- tem. This allows one to go back and redo an analysis with only that part of the data that was available at an earlier time. One can also search for trends in the data. For example, one can easily verify that the reported ec- centricity for a given planetary systems is almost cer- tainly decreasing as a function of time because new LISTING 1. -- Example of a database entry in the Open Exoplanet Cata- logue: CoRoT-5. <mass>1.03</mass> <radius>1.186</radius> <temperature /> <planet> 1 <name>CoRoT-5</name> 2 <rightascension>06 45 07</rightascension> 3 <declination>+00 48 55</declination> 4 <distance>400</distance> 5 <star> 6 7 8 9 10 11 12 13 14 15 <name>CoRoT-5 b</name> <list>Confirmed planets</list> <mass>0.467</mass> <radius>1.388</radius> <period>4.0378962</period> <semimajoraxis>0.0494 </semimajoraxis> 16 17 18 19 20 21 <eccentricity>0.09</eccentricity> <inclination>85.83</inclination> <description>CoRoT-5 b orbits an F -type star in the constellation Monoceros. </description> <discoverymethod>transit </discoverymethod> <lastupdate>09/07/18</lastupdate> <discoveryyear>2008 </discoveryyear> 22 23 </star> </planet> data becomes available and partly counteracts a strong bias in fitting Keplerian orbits (Shen and Turner 2008). Studies like this are not possible if only the current ver- sion of a catalogue are available. Blame and Credit: Each line in each file can be traced back to the person who made changes to it and to the peer-reviewd publication where the data is taken from. Therefore the entry database is verifiable. If found that a person made dubious entries across many different files, this can be easily corrected by reverting all those commits to the version control system. 3. HUMAN-READABLE XML FILES It is important to stress that the discussion in this paper is very general and does not depend on the specific file format. There are many options and the choice is often just a matter of personal preference. There are only two requirements for a file format to work nicely with a version control system: the files have to be text files and each value should be placed in a separate line. I chose XML for the Open Exoplanet Catalogue because of the flexible tree structure that allows me to con- struct arbitrary star systems and because of the vast number of parsing libraries that are readily available for many differ- ent languages. 3.1. Basic features of the XML file format XML stands for Extensible Markup Language. It is a met- alanguage that allows anyone to define their own customized markup languages. To astronomers, XML might be most fa- miliar due its similarity to HTML, the language in which the World Wide Web is written. However, XML is not about the displaying or formatting but rather about the creating and sharing of information. A document is a tree structure that consists of tags such as LISTING 2. -- Example of a python script that outputs a list of planet names and masses from the Open Exoplanet Catalogue. 3 <tag attributename="attributevalue">text</tag>. Each tag can contain text and/or other tags. Tags can also con- tain optional attributes. In contrast to HTML, each tag must be closed explicitly. XML contains a lot more features such as the possibility to precisely define and verify a customized language. Most of these features are not important for the fol- lowing discussion and I will therefore not go into any more detail. Listing 1 shows an example of a database entry written in the XML file format. It describes the planetary system CoRoT-5 and is taken from the Open Exoplanet Catalogue. Without any further introduction, header file or manual, any astronomer can pick out the most important pieces of infor- mation. For example, you can easily convince yourself that this is a Hot Jupiter orbiting a 1.03 solar mass star at 0.04947 astro- nomical units and a period of 4.03 days. It is located 400 par- secs away from the Sun at the coordinates α = 6h45m7s, δ = 0◦48′55". In this particular example there are no units given the in the XML file itself because of well established conven- tions (e.g. quoting the star's mass in units of Solar mass). However, adding units to tags in the form of an attribute is straightforward: <mass unit="solarmass">1.03</mass>. Errorbars and any additional meta data can be added in a sim- ilar way. Tags for information that does not exist for a given object can simply be removed. This dramatically enhanced readability and also saves disk space. One can also empha- size the missing value with an empty tag, for example: <temperature />. I hope this example has convinced you that this format makes the entire database human readable. This is of course no sub- stitute for a full description of each fields and all the implicit conventions used. But it takes away a lot of unnecessary over- head. The XML file in Listing 1 is also a good example to show how hierarchical information can be stored. The planet clearly belongs to the star. Now, imagine someone discoveries a wide binary companion to CoRoT-5. Adding another star to the system is trivial, it is just another <star> tag at the end of the file. Even something exotic as adding a moon which is orbit- ing the planet can be added very naturally. None of this breaks the structure of the database. Scripts parsing this database re- main fully functional, as described below7. Using only small text files to store the database has another big advantage over more traditional approaches that store ev- erything in one big file or database such as MySQL: small text files work nicely together with version control systems. These systems were originally designed to track changes in source code. It also makes sense form a logical point of view as each line contains exactly one piece of information. 3.2. Scriptablility XML files are great for scripting. Listing 2 shows an exam- ple of a short python script that queries the Open Exoplanet Catalogue and outputs the name and mass of all planets. 7 This statement is of course without loss of generality; one can always construct a script which will not work anymore. 1 2 3 4 5 6 7 import xml.etree.ElementTree as ET, glob for filename in glob.glob("*.xml"): tree = ET.parse(open(filename, 'r')) planets = tree.findall(".//planet") for planet in planets: print planet.findtext("./name"), print planet.findtext("./mass") The use of a scripting language like python and the XML file format allows one to easily create complex queries. For example, imagine you are only interested in planets that are in multi-planetary system. Adding the single line if len(planets)>1: after line 4 in Listing 2 is all that is needed to accomplish that. The syntax is so simple that it does not even require any expla- nation. Scripts can also be used to perform automatic updates to the database or even create the entire database from data in a completely different format. Because the entire database is stored in a version control system, there is no risk of acciden- tally deleting information. 4. VERSION CONTROL SYSTEM GIT The version control system git has seen a tremendous suc- cess in the open source world. It has become the de facto standard for distribution version control systems. Several al- ternatives with almost equal functionality exist (for example Bazaar and Mercurial). Although the following discussion is very general, it will focus on git. This is not meant to be an introduction to git. I want to point out how git can be used for maintaining an astronomical database. There are many good tutorials on git and you are strongly encouraged to have a look at those. An interactive and well made intro- duction can be found at http://try.github.com/. 4.1. Workflow Every copy (fork) of a git repository contains the entire history of every change ever committed to the database. There is no central server, no master and no slaves. The copy on your laptop is completely equal to the copy hosted on your web-server. However, more people might have access to the copy on your web-server. When starting to work on a new database or catalogue, you create a new repository on a local machine and start adding data. You can then push the repository to another publicly accessible or private computer (ssh access is enough to al- low git to communicate). A possibility worth looking into is the online platform http//github.com that was already mentioned above. It is a free hosting service for git reposi- tories. Collaborators can then copy (pull) the repository and make their own changes and additions. When they think their changes are good enough to require your attention, they will send you a pull request. You review the changes and check for consistency of the data files. For example, you might need to manually resolve a conflict if the same line in the same file was edited by more than one person. Once the changes are approved, anyone else will be able to download the newest version. There are several option when the maintainer does not agree or is not satisfied with the submitted changes. He can simply ignore the changes or try to improve them. He can also create 4 a new branch within the same repository and call it experi- mental or controversial. This allows others to see and discuss the submission but does not pollute the main database with questionable data. It is worth noting that a human readable format uses more disk space than a binary format. However, git automatically compresses data so that this should not be a big issue for most small to medium sized catalogues. In addition to that, there is no need to add tags to the XML file which do not con- tain any data. For example, a tag containing the value for the Rossiter-McLaughlin effect (Rossiter 1924; McLaughlin 1924) only needs to be added to those planets that had such an observation. submodule functionality is 4.2. Use of git-based catalogues in other git repositories git's ideally suited for databases and catalogues. It allows an external git repository to be added to another git repository as a submodule. This cre- ates a logical separation between a database and scripts that make use of it. But at the same time, this configuration can keep the database in sync with the public version. 4.3. References to scientific publications Including references is absolutely essential for any scien- tific database. However, it increases the size, readability and complexity of any data format dramatically. Once again, git comes to our rescue. The by far simplest way to add reference to a database that is storred in git is the commit message. This enables anyone to trace back each value to the person who made the entry in the database, but also to the scientific publication where the information is coming from. Note that this is only possible if the file format is constructed such that there is one value (or at most one set of values that comes from the same publication) per line. 5. THE OPEN EXOPLANET CATALOGUE The Open Exoplanet Catalogue is an example of a database that incorporates all the ideas presented above. It is a daily updated catalogue of all discovered extra-solar planets. Fur- thermore, it is to my knowledge the only catalogue that cor- rectly stores and represents planets in an arbitrary star system (s- and p-type binaries, triple and quadruple star systems, etc). The catalogue is hosted on github at https://github.com/ hannorein/open_exoplanet_catalogue/. The project was started in late 2010. The motivation came from the simple fact that I needed a catalogue of all discovered extra-solar planets but I was unhappy with ex- isting catalogues. Some catalogues, for example http:// exoplanets.org (Wright et al. 2011) are updated only ir- regularly and often lag several weeks behind important discoveries. Other websites, most importantly http:// exoplanet.eu (Schneider 2011), are usually updated very quickly after an announcement is made. But there were sev- eral typos and other inconsistencies in the catalogue that I needed to keep track of and change manually after every up- date because there was no way for me to directly contribute to the database itself. In addition to that, certain pieces of infor- mation (such as the discovery method) were available on the website, but not in the downloadable files. I ended up hav- ing to parse the (always changing) HTML files to find that information. The original dataset was taken from the two websites men- tioned above but has since been modified heavily. As of De- LISTING 3. -- Example of the database entry of the circumbinary planet PH-1b from the Open Exoplanet Catalogue. 1 <name>PH-1</name> 2 <binary> 3 4 5 <semimajoraxis>1000</semimajoraxis> <binary> 6 7 8 9 10 11 12 13 14 15 16 17 <semimajoraxis>0.1744</ semimajoraxis> <star> <name>PH-1 Aa</name> <mass>1.384</mass> </star> <star> <name>PH-1 Ab</name> <mass>0.386</mass> </star> <planet> <name>PH-1 (AaAb) b</name> <name>Kepler-64 b</name> <semimajoraxis>0.634 </semimajoraxis> </planet> <semimajoraxis>60</semimajoraxis> <star> <name>PH-1 Ba</name> <mass>0.99</mass> </star> <star> <name>PH-1 Bb</name> <mass>0.51</mass> </star> </binary> <binary> 18 19 20 21 22 23 24 25 26 27 28 29 30 31 </binary> </binary> cember 4th 2012 the database contains 861 planets. The cata- logue also contains a separate dataset of the latest Kepler Ob- jects of Interest (Borucki and the Kepler Team 2011) which uses the same data format. The compressed size of the main catalogue without images is less than 80kb. It is also used as an input catalogue for the iOS application Exoplanet8. 5.1. Planets in multiple star system The Open Exoplanet Catalogue can represent planetary systems in arbitrary hierarchical systems. The basic struc- ture contains only three objects: <planet>, <star> and <binary>. A binary hosts two stars, a star and a binary or two other binaries. Both stars and binaries can host planets. These simple rules allows one to construct any system. List- ing 3 shows the circumbinary system PH-1 (AaAb) b (also known as Kepler-64, see Schwamb et al. 2012) which is part of a quadruple system. In a standard, non-hierarchical database such as a simple table, it is almost impossible to represent systems like these. On the other hand the flexibility of a tree based file format like XML, allows one to add these on the fly to an exist- ing database and in a very natural way. Note that the script in Listing 2 is still fully functional even for this complicated quadruple system. 8 http://exoplanetapp.com 5.2. Images and other binary data The Open Exoplanet Catalogue contains image files of di- rectly imaged exoplanets and artistic illustrations often re- leased by various space agencies. This is an example of how binary data and images can be attached to a catalogue based on the XML file format. Binary data can be included in XML files in the same way e-mails can include attachments9. In most cases a much simpler approach, storing images and bi- nary data in a separate file, is sufficient. To link the XML file to the binary file, one only has to add a tag with the filename. In the Open Exoplanet Catalogue, the <image> tag is used to store images related to a certain planet. An additional tag <imagedescription> tag is used to store a description, a reference to the scientific paper and copyright information for the image. 5.3. Other user interfaces The entire database can be directly accessed via github. Be- cause each entry is a human-readable XML file, one can sim- ply read that file online in a web browser. In fact, some of these files might show up on the first page when you google a specific planetary system. However, an interactive and graphical user interface is of- ten very useful. I created such a website, the Visual Exoplanet Catalogue10. It is based on the Open Exoplanet Catalogue and is using it as a back-end. The catalogue is also easily acces- sible with the free iOS application Exoplanet8. I'm looking forward to other new and exciting projects that will make use of this catalogue. 5.4. List of tags Table 1 contains a list and the definitions of all tags cur- rently used in the Open Exoplanet Catalogue. This list will certainly grow with time. 6. CONCLUSIONS My experience of talking to astronomers and astrophysi- cists is that many are not very familiar with the resources and tools that the computer science and open source commu- nity uses. However, the benefits listed in Section 2 which these tools can provide to any new catalogue or database are 5 tremendous. A long ASCII table, a FITS file, a MySQL database or a complicated web-form is just not good enough anymore. With this proposal and the Open Exoplanet Cata- logue, I hope to encourage discussion, experimentation and innovation. The ideas and concepts presented here are not meant to be final and leave a lot of room for improvements. Many of the details will strongly depend on the specific kind and usage scenario of the database. The examples here are bi- ased towards a catalogue of extra-solar planets. Other areas where heterogeneous datasets are compiled into catalogues and where ideas very similar to those presented here can be readily applied are for example a catalogue of gravitational lenses and a database of solar system objects. The Open Exoplanet Catalogue has already proven to be very useful for several of my own projects (e.g. Rein 2012). But it is by no means perfect. For example, it is missing important information such as the Rossiter-McLaughlin mea- surements and most certainly contains errors. But the impor- tant point is that YOU can help improve it. Visit the cata- logue's website on github, create an account, clone the repos- itory and if you find a typo or want to add new data, simply make the changes (there is no need to ask for permission). If you are an observer, please consider adding your own data when you announce a new discovery. The idea of a demo- cratic, decentralized astronomical database can only work if many people contribute to it. I am excited to find out if we, as a community, can find enough people to maintain this cat- alogue. If you would like to contribute to the catalogue but have problems getting started with git or if you have any com- ments on this proposal, please do not hesitate to contact me. ACKNOWLEDGMENTS Hanno Rein would like to thank Dave Spiegel for inspiring conversations regarding the structure of the Open Exoplanet Catalogue. Scott Tremaine, Dave Spiegel, Chi-kwan Chan, Leonidas Moustakas, Peter Williams and Wladimir Lyra gave valuable feedback on this paper. Hanno Rein was supported by the Institute for Advanced Study and the NFS grant AST- 0807444. Support for the Exoplanet App came from the Royal Astronomical Society via a public outreach grant. REFERENCES Y. Shen and E. L. Turner, ApJ 685, 553 (2008), arXiv:0806.0032. R. A. Rossiter, ApJ 60, 15 (1924). D. B. McLaughlin, ApJ 60, 22 (1924). J. T. Wright, O. Fakhouri, G. W. Marcy, E. Han, Y. Feng, J. A. Johnson, A. W. Howard, D. A. Fischer, J. A. Valenti, J. Anderson, and N. Piskunov, PASP 123, 412 (2011), arXiv:1012.5676 [astro-ph.SR]. J. Schneider, in EPSC-DPS Joint Meeting 2011 (2011) p. 3. W. J. Borucki and the Kepler Team, ApJ 736, 19 (2011), arXiv:1102.0541 [astro-ph.EP]. M. E. Schwamb, J. A. Orosz, J. A. Carter, W. F. Welsh, D. A. Fischer, G. Torres, A. W. Howard, J. R. Crepp, W. C. Keel, C. J. Lintott, N. A. Kaib, D. Terrell, R. Gagliano, K. J. Jek, M. Parrish, A. M. Smith, S. Lynn, R. J. Simpson, M. J. Giguere, and K. Schawinski, ArXiv e-prints (2012), arXiv:1210.3612 [astro-ph.EP]. H. Rein, MNRAS , L526 (2012), arXiv:1208.3583 [astro-ph.EP] 9 For example, this can be done using a so called Multipart/related MIME type, RFC 2112, see http://tools.ietf.org/html/rfc2112. 10 http://exoplanet.hanno-rein.de 6 Tag <system> <planet> <star> <binary> <declination> <rightascension> <distance> <name> <semimajoraxis> <eccentricity> <periastron> <longitude> <ascendingnode> <inclination> <period> <mass> Can be child of - <system>, <binary>, <star> <system>, <binary> <system>, <binary> <system> <system> <system> <system>, <star>, <planet> <binary>, <planet> <binary>, <planet> <binary>, <planet> <binary>, <planet> <binary>, <planet> <binary>, <planet> <binary>, <planet> <planet>, <star> Description This is the root container for an entire planetary system This is the container for a single planet. The planet is a free floating (orphan) planet if this tag is a child of <system>. This is the container for a single star. A star can be host to one or more planets (circum-stellar planets). A binary consists of either two stars, one star and one binary or two binaries. In addition a binary can be host to one or more planets (circum-binary planets). Declination Right ascension Distance from the Sun Name of this object. This tag can be used multiple times if the object has multiple Names. Semi-major axis of a planet (heliocentric coordinates) if child of <planet>. Semi-major axis of the binary if child of <binary>. Eccentricity Longitude of periastron Mean longitude at a given Epoch (same for all planets in one system) Longitude of the ascending node Inclination of the orbit Orbital period Mass (or m sin(i) for radial velocity planets) <radius> <planet>, <star> Physical radius <temperature> <age> <metallicity> <spectraltype> <magV> <discoverymethod> <planet>, <star> <planet>, <star> <star> <star> <star> <planet> <description> <image> <planet> <planet> <imagedescription> <new> <planet> <planet> <discoveryyear> <lastupdate> <planet> <planet> Temperature (surface or equilibrium) Age Stellar metallicity Spectral type Visual magnitude Discovery method of the planet. For example: timing, RV, transit, imaging. Short description of the planet Filename without extension of a picture of the planet. File is stored in the images directory. Short description and copyright information of the image. The value for this tag is 1 if the system has been added recently (since the last update or within 48 hours) Year of the planet's discovery Date of the last (non-trivial) update All tags currently used in the Open Exoplanet Catalogue. TABLE 1 Unit ±dd mm ss hh mm ss parsec AU degree degree degree degree day Jupiter masses (<planet>), Solar masses (<star>) Jupiter Solar radii (<star>) Kelvin Gyr log, relative to solar (<planet>), radii yyyy yy/mm/dd
1801.09474
2
1801
2018-02-22T18:06:47
C/O vs Mg/Si ratios in solar type stars: The HARPS sample
[ "astro-ph.EP", "astro-ph.SR" ]
Aims. We present a detailed study of the Mg/Si and C/O ratios and their importance in determining the mineralogy of planetary companions. Methods. Using 499 solar-like stars from the HARPS sample, we determine C/O and Mg/Si elemental abundance ratios to study the nature of the possible planets formed. We separated the planetary population in low-mass planets ( < 30 $\rm M_{\odot}$) and high-mass planets ( > 30 $\rm M_{\odot}$) to test for possible relation with the mass. Results. We find a diversity of mineralogical ratios that reveal the different kinds of planetary systems that can be formed, most of them dissimilar to our solar system. The different values of the Mg/Si and C/O ratios can determine different composition of planets formed. We found that 100\% of our planetary sample present C/O < 0.8. 86\% of stars with high-mass companions present 0.8 > C/O > 0.4, while 14\% present C/O values lower than 0.4. Regarding Mg/Si, all stars with low-mass planetary companion showed values between 1 and 2, while 85% of the high-mass companion sample does. The other 15\% showed Mg/Si values below 1. No stars with planets were found with Mg/Si > 2. Planet hosts with low-mass companions present C/O and Mg/Si ratios similar to those found in the Sun, whereas stars with high-mass companions have lower C/O.
astro-ph.EP
astro-ph
Astronomy & Astrophysics manuscript no. 30743_corr February 23, 2018 c(cid:13)ESO 2018 C/O vs. Mg/Si ratios in solar type stars: The HARPS sample L. Suárez-Andrés1, 2, 3,(cid:63), G. Israelian1, 2, J.I. González Hernández1, 2, V. Zh. Adibekyan4, E. Delgado Mena4 , N. C. Santos4, 5 and , S. G. Sousa4, 5 1 Instituto de Astrofísica de Canarias, E-38205 La Laguna, Tenerife, Spain 2 Depto. Astrofísica, Universidad de La Laguna (ULL), E-38206 La Laguna, Tenerife, Spain 3 Isaac Newton Group of Telescopes, Apartado de Correos 321, E-38700 Santa Cruz de la Palma, Spain 4 Instituto de Astrofísica e Ciências do Espaço, Universidade do Porto, CAUP, Rua das Estrelas, 4150-762 Porto, Portugal 5 Departamento de Física e Astronomia, Faculdade de Cîencias, Universidade do Porto, 4169-007 Porto, Portugal Received March 7, 2017; accepted January 14, 2018 ABSTRACT Aims. We aim to present a detailed study of the magnesium-to-silicon and carbon-to-oxygen ratios (Mg/Si and C/O) and their importance in determining the mineralogy of planetary companions. Methods. Using 499 solar-like stars from the HARPS sample, we determined C/O and Mg/Si elemental abundance ratios to study the nature of the possible planets formed. We separated the planetary population in low-mass planets ( < 30 M(cid:12)) and high-mass planets ( > 30 M(cid:12)) to test for a possible relation with the mass. Results. We find a diversity of mineralogical ratios that reveal the different kinds of planetary systems that can be formed, most of them dissimilar to our solar system. The different values of the Mg/Si and C/O can determine different composition of planets formed. We found that 100% of our planetary sample present C/O < 0.8. 86% of stars with high-mass companions present 0.8 > C/O > 0.4, while 14% present C/O values lower than 0.4. Regarding Mg/Si, all stars with low-mass planetary companion showed values between one and two, while 85% of the high-mass companion sample does. The other 15% showed Mg/Si values below one. No stars with planets were found with Mg/Si > 2. Planet hosts with low-mass companions present C/O and Mg/Si similar to those found in the Sun, whereas stars with high-mass companions have lower C/O. Key words. stars: abundances - stars: atmospheres – stars: planetary systems 1. Introduction The determination of the chemical composition of extra- solar planets has been the subject of numerous studies in recent years. One of the keystones has been the fact that stars hosting giant planets are considered to be metal-rich when compared with single stars (Gonzalez 1997; Santos et al. 2001, 2004; Fischer & Valenti 2005; Gonzalez 2006). Stars hosting low mass planets (those with masses below 30 M(cid:12)) do not seem to be preferentially metal-rich (e.g. Sousa et al. 2008, 2011a) Recent studies of chemical abundances in stars with and without planets have shown no important differences in [X/Fe] vs. [Fe/H] trends between the two groups of stars for refractory (González Hernández et al. 2010; Delgado Mena et al. 2010) and volatile elements (Suárez-Andrés et al. 2016, 2017). On the other hand, studies by Haywood (2008, 2009) and Adibekyan et al. (2012a) found evidence of α-elements enhancement in stars with planets. As both planetesimals and planets are formed within the same environment, their composition is expected to (cid:63) Send offprint requests to: L. Suárez-Andrés, e-mail: [email protected] be the same as that of their host star. This assumption might be true for refractory species, but not for the volatile ones (Lodders 2003; Thiabaud et al. 2014). Although this fact does not affect the magnesium-to-silicon ratio (Mg/Si) for rocky planets, it can affect the carbon-to-oxygen ratio (C/OP (Dorn et al. 2015; Thiabaud et al. 2015b). Elemental abundance ratios (C/O and Mg/Si) are important as they govern the distribution and formation of chemical species in the protoplanetary disc. Variations in C/O can be found among planetary atmospheres and host stars (Madhusudhan et al. 2012; Konopacky et al. 2013; Moses et al. 2013; Brewer et al. 2016b) and among planets in the same planetary system. This is due to different parameters (temperature, pressure, etc.) and processes at work during the planet formation stage, including possible migrations of planets from their birthplace (Öberg et al. 2011; Ali-Dib et al. 2014; Mad- husudhan et al. 2012; Thiabaud et al. 2015a,b; Brewer et al. 2016b) In the last few years, several studies have tried to un- derstand the formation and evolution of planets using the- oretical models. Bond et al. (2010a,b), Carter-Bond et al. (2012) studied planet formation scenarios with different ini- Article number, page 1 of 9 8 1 0 2 b e F 2 2 . ] P E h p - o r t s a [ 2 v 4 7 4 9 0 . 1 0 8 1 : v i X r a A&A proofs: manuscript no. 30743_corr tial composition, but they did not carry out a detailed study of the output volatile species. Elser et al. (2012) tried to study this planet formation model by adding the formation of solids, but they were unable to reproduce some features present in the solar system, such as high Fe on Mercury. Recently, Thiabaud et al. (2014); Marboeuf et al. (2014) made a complete study, presenting models not only for re- fractory species, but for volatiles as well. Thiabaud et al. (2015a) presented a complete study, taking into account the accretion of several compounds omitted in previous works such as He, H2, H2O, CO, CO2, CH3OH, CH4, NH3, N2, and H2S. Also, they computed the importance of volatile elements in protoplanetary discs and their implications in planetary formation (Thiabaud et al. 2015b). Their models show that the condensation of volatile species as a func- tion of radial distance allows for C/O enrichment in spe- cific parts of the protoplanetary disc of up to four times the solar value. This could lead to the formation of planets that can be enriched in C/O in their envelope up to three times the solar value. Their models are consistent with re- cent observations of hot-Jupiter atmospheres (Brewer et al. 2016b). The amount of carbides and silicates formed in planets is controlled by the carbon-oxygen ratio (e.g. Bond et al. 2010a) – if C/O < 0.8: Si will form solid SiO4− 4 and SiO2, serving as seeds for Mg silicates for which the exact composition will be determined by Mg/Si. – if C/O > 0.8: Si will be solid as SiC. Also, graphite and TiC will be formed. Silicates are an important ingredient in the formation of rocky planets, as they are the most abundant compounds in the mantle and crust of these planets (Morgan & An- ders 1980). Silicate distribution is ruled by Mg/Si of the planet host star. Concerning Mg/Si, the principal compo- nents could be, as proposed by Bond et al. (2010b): – if Mg/Si < 1, Mg forms orthopyroxene (MgSiO3) and the remaining Si forms other minerals, such as feldspar (CaAl2Si2O8, NaAlSiO8) or olivine(Mg2SiO4) – if 1 < Mg/Si < 2, Mg is distributed equally between pyroxene and olivine. – if Mg/Si > 2, Si forms olivine and the remaining Mg forms other oxides like MgO. Testing and improving planetary formation models is key to future studies of habitability, as these ratios are essential elements in defining the structure of the planet. In this article we present a study of C/O and Mg/Si in solar-type stars and their implications for possible terres- trial planetary formation. 2. Sample description The high-resolution spectra analysed in this article were obtained with the HARPS spectrograph at La Silla Ob- servatory (ESO, Chile) during the HARPS GTO, HARPS- 2 and HARPS-4 programmes (see Mayor et al. 2003; Lo Curto et al. 2010; Santos et al. 2011, for more information about the programme). The spectra have been used previ- ously in the analysis of stellar parameters, as well as the derivation of precise chemical abundances (see e.g. Sousa Article number, page 2 of 9 et al. 2008, 2011a,b; Adibekyan et al. 2012b; Tsantaki et al. 2013; Bertran de Lis et al. 2015; Suárez-Andrés et al. 2017). The studied sample is part of the 1111-star sample pre- sented by Adibekyan et al. (2012b) for which the abun- dance of volatiles were measured (Bertran de Lis et al. 2015; Suárez-Andrés et al. 2017). We limited our metallicity sam- ple for stars with [Fe/H] > -0.6, which is the lower [Fe/H] for the planet-host sample. We studied 499 FGK solar-type stars, with effective temperatures between 5250 K and 6666 K, metallicities from −0.59 to 0.55 dex, and surface gravi- ties from 3.81 to 4.82 dex. Of 499 stars, 99 are planet hosts1, whereas the other 400 are single stars (stars with no known planetary companion, also known as comparison stars. We urge the reader to take into account that stars considered as single stars might har- bour undetected planets (see Mayor et al. 2011; Faria et al. 2016). 3. Chemical abundances To obtain the carbon-to-oxygen (C/O) and magnesium- to-silicon (Mg/Si) elemental abundance ratios we adopted chemical abundances from Adibekyan et al. (2012b); Bertran de Lis et al. (2015) and Suárez-Andrés et al. (2017) for Mg and Si, O, and C, respectively. All these chemical abundances were obtained using the same high-resolution high-quality HARPS spectra and the same stellar parame- ters, allowing us to obtain precise elemental abundance ra- tios. For carbon we studied the CH molecular band located at 4300 Å (see Suárez-Andrés et al. 2017), whereas for the other elements atomic features were used. Recent studies of molecular CH features (Suárez-Andrés et al. 2017) have proved that they are as reliable as atomic ones, and provide consistent results. Although the adopted abundances come from different sources, O, Mg, and Si abundances were obtained following the same procedure (EW method). For the case of C, syn- thetic fitting was applied, but compared to the EW method results to confirm their validity (see Suárez-Andrés et al. 2017). Errors are significantly lower for Mg and Si (aver- ages 0.06 and 0.04, respectively) than for C and O (average 0.15 for both cases), which affects our final elemental abun- dance ratio errors. Adopted solar abundances elements are log (C)(cid:12) = 8.50 dex (Caffau et al. 2010), log (O)(cid:12) = 8.71 (Caffau et al. 2008), log (Mg)(cid:12) = 7.58 dex and log (Si)(cid:12) = 7.55 (Anders & Grevesse 1989). for all 4. Elemental abundance ratios Theoretical studies suggest that C/O and Mg/Si are very important in determining the mineralogy of terrestrial plan- ets. Since the mineralogy of planets is commonly studied in terms of absolute ratios, we use these instead of solar ra- tios2. 1 Data from www.exoplanet.eu and the SWEET-CAT cata- logue (www.astro.up.pt/resources/sweet-cat/) 2 We remind the reader that C/O (cid:54)= [C/O]. Please see Section 6 for more information. Suárez-Andrés, L. et al.: C/O vs. Mg/Si ratios in solar type stars: The HARPS sample. Elemental abundance ratios were calculated using the following equation A/B = NA/NB = 10log(A)/10log(B), (1) where log(A) and log(B) are absolute abundances. Er- rors were estimated by evaluating an increase or decrease in the log(A) − log(B) abundance ratio, due to the relative errors (For more details, see Delgado Mena et al. 2010). Median errors are represented in all figures.(0.14 for C/O and 0.06 for Mg/Si)) A sample table is shown in Table 1. Table 1: Sample of C/O and Mg/Si abundances for a set of stars (see online table). Star HD39091 HD82943 HD100508 HD100777 HD102117 HD212708 Teff 5991 5992 5384 5530 5620 5644 log g 4.40 4.42 4.39 4.31 4.28 4.31 [Fe/H] 0.09 0.28 0.35 0.25 0.26 0.26 C/O 0.61 ± 0.11 0.63 ± 0.13 0.54 ± 0.09 0.53 ± 0.04 0.48 ± 0.09 0.51 ± 0.09 Mg/Si 1.11 ± 0.10 0.99 ± 0.04 0.91 ± 0.01 0.96 ± 0.12 1.07 ± 0.01 1.06 ± 0.07 Fig. 1: Comparative C/O versus [Fe/H]. Green dots show data from our work; red stars show results from Nissen et al. (2014); blue squares from Teske et al. (2014); orange triangles from González Hernández et al. (2013) 4.1. C/O The C/O in planet-host stars can provide key informa- tion about the protoplanetary disc in which the planet was formed. The dependence on the distance for volatile ele- ments will affect the C/O expected in exoplanetary atmo- spheres, as volatiles are heavily affected by different dis- tances (or different ice line positions) during the early life- time of the nebula while planets accrete (Öberg et al. 2011; Ali-Dib et al. 2014; Brewer et al. 2016b). Figure 2 shows the C/Os derived in this paper as a function of [Fe/H] for these samples; stars with and without planets. We obtain a linear fit for both samples, with very little differences between them. We can see a dependence of C/O on metallicity, in agreement with other works, such as Nissen et al. (2014), Teske et al. (2014) and Brewer & Fischer (2016). Several authors have studied the C/O ratios in stars with planets (e.g. Petigura & Marcy 2011; Nissen 2013; Nissen et al. 2014; Teske et al. 2014). In Fig. 1 we can see our results compared with the work of Nissen et al. (2014), Teske et al. (2014), and González Hernández et al. (2013). To make this comparison possible, we scaled the carbon and oxygen abundances presented by these authors to our ref- erence values (logN(cid:12)(O) = 8.71 and logN(cid:12)(C) = 8.50). As seen in Fig. 1, there is a good fit at all metallicities between our work and these previous studies. To test for possible relationships between C/O and the masses of planetary companions, we separated the planet population into two groups: low-mass planets (LMP; with masses less than or equal to 30 M⊕) and high-mass planets (HMP; with masses greater than 30 M⊕). In those stars which host several planets, the most massive planet in the system was considered in our study. Our sample consists of 19 low-mass and 80 high-mass planet hosts. If we take a closer look at the C/O distribution for the studied samples in Fig. 3, we can see that the planetary samples exhibit almost the same behaviour, with an small offset, while the single star sample spreads more, with a significant higher Fig. 2: C/O versus [Fe/H]. Red open circles refer to single stars while green dots refer to low-mass planet host stars and blue diamonds to high-mass planet-hosts. FWHM (see Table 2). In our sample, 100% of stars with planets have C/O values lower than 0.8. Average value for the low-mass and high-mass samples are 0.46±0.11 and 0.50±0.10, respectively. Average value for the single stars sample is 0.45±0.13. Our stars are slightly carbon poor when compared to the solar reference (C/O(cid:12)=0.61). These results agree, within errors, with those presented in Nissen et al. (2014); Teske et al. (2014); González Hernández et al. (2013). We can see in Fig. 3 the C/O distributions for single stars and both planetary samples. A cumulative histogram (Fig. 4) shows us that each sample behaves in a different way. We performed a Kolmogorov -Smirnov test (K–S test) to confirm these behaviours. The null hypothesis is rejected Article number, page 3 of 9 A&A proofs: manuscript no. 30743_corr Table 3: K–S test for all the samples. C/O Mg/Si Sample 0.16 LMP - NP 0.31 HMP - NP 0.25 0.12 at level α=0.1 if our results are higher than Dnn(cid:48)=0.29 and Dnn(cid:48)=0.15, for the low-mass and high-mass samples, respectively, following the expression (cid:114) Dnn(cid:48) > c(α) n + n(cid:48) nn(cid:48) ; c(α = 0.10) = 1.22. (2) Fig. 3: C/O distributions for stars harbouring low-mass (green) and high-mass (blue) planets. Stars without planets are shown in red. We can see in Table 3 that the C/O is above the threshold limit. The probabilities of similarity are 4% and 0% for the low and high-mass samples, respectively, so we can assume that the samples do not come from the same distribution. 4.2. Mg/Si The magnesium and silicon elemental abundance ratio con- trols the exact composition of silicates that can be found in the planetary companion, as this ratio, along with Fe/Si, does not depend so strongly on the distance to the star as the C/O ratio does (Thiabaud et al. 2015a,b; Dorn et al. 2015). In Fig. 5 top panel we show Mg/Si ratios as a function of metallicity. As expected, the Mg/Si ratio decreases with [Fe/H], with a similar slope for all the studied sample (see Table 4) We also find a relation between values of Mg/Si and Teff, as shown in the bottom panel of Fig. 5. We ob- tained slopes and standard deviations for the three studied samples (see Table 4) and, due to the high scatter of the single star sample at Teff higher than 6100K, we cannot conclude any physical difference. However, as we can see that that slopes and scatter increase, we decided not to use these stars with Teff higher than 6100K to avoid introduc- ing errors to our analysis. Given the uncertainty of the implication of NLTE cor- rections for these hot stars, we will study, from now on, only the sample with Teff < 6100K. The average value for the low-mass and high-mass samples are 1.15±0.08 and 1.12±0.15, respectively. The average value for the single stars sample is 1.12±0.12. These results agree, within er- rors, with those presented in Brewer et al. (2016b). The distribution of Mg/Si for both planet-host stars and the comparison sample are shown in Figure 6. We can see that all samples exhibit the same behaviour. In our sam- ple 100% of the low-mass sample have an Mg/Si value of between 1.0 and 2.0 while 85% of the high-mass sample ex- hibit this behaviour. We also find that none of the low-mass planet hosts have Mg/Si values below 1.0, but 15% of the high-mass sample does. No stars with Mg/Si values greater than 2.0 were found. Figure 7 shows Mg/Si distributions for single stars and both planetary samples. All samples are concentrated around Mg/Si∼1.12, only 0.05 higher than our solar refer- ence (1.07) and within errors. Statistics for these histograms are shown in Table 2. A cumulative histogram (Fig. 7) shows that each sample behaves in the same way. We find no Fig. 4: Cumulative C/O distributions for single stars (red) and stars with planets (low-mass (green) and high-mass planets (blue)). Table 2: Statistics for the fitted histograms presented in Figures 11, 3, and 6. The low-mass planets (LMP) and the high-mass planets sample (HMP) C/O Mg/Si [C/O]corr Sample Single stars Planets: LMP Planets: HMP Single stars Planets: LMP Planets: HMP Single stars Planets: LMP Planets: HMP Centre-fit FWHM 0.45 0.52 0.52 1.11 1.06 1.08 -0.11 -0.01 -0.03 0.29 0.10 0.21 0.22 0.25 0.16 0.25 0.11 0.17 Article number, page 4 of 9 Suárez-Andrés, L. et al.: C/O vs. Mg/Si ratios in solar type stars: The HARPS sample. Table 4: Statistics for the three groups studied (Comparison sample, low-mass planets and high-mass planets) for Mg/Si vs. Teff (see Fig. 5, bottom panel). Mg/Si vs. Teff (for Teff < 6100 ) Slope -1.39E-04 -1.21E-05 -1.63E-04 Slopeerr 0.32E-04 8.98E-05 0.90E-04 Mg/Si vs. Teff (for Teff > 6100 ) Slope Slopeerr 2.54E-04 - -2.37E-04 - -7.17E-04 2.97E-04 NP LMP HMP NP LMP HMP Fig. 5: Mg/Si as a function of [Fe/H] (top panel) and Teff (bottom panel). Red open circles refer to single stars while green dots refer to low-mass planet host stars and blue di- amonds to high-mass planet-hosts. differences between the planetary samples, as all three sam- ples peak around 1.10. The high-mass sample has a broader distribution than the low-mass one, but the limited number of the low-mass sample does not allow us to infer any result other than the similarity between the samples. A similar re- sult can be found in Brewer et al. (2016b). We performed a K–S test to confirm these behaviours between the samples. As we can see in Table 3, right col- umn, we cannot reject the null hypothesis for the any plan- etary sample, so we have to assume that the samples come from the same distribution. 5. C/O vs. Mg/Si: Implications on planet formation We studied C/O as a function of Mg/Si as a way to study the possible scenario for the formation of planets. In Fig. 8 we see how stars with and without planets are distributed in a C/O against Mg/Si plot. As can be seen, stars with Fig. 6: Mg/Si distributions for stars harbouring low-mass (green) and high-mass (blue) planets. Stars without planets are shown in red. planets, both low-mass and high-mass, are concentrated at C/O values of 0.5-0.6 and Mg/Si values of ∼ 1.0. The solar values of C/O(cid:12) = 0.61 and Mg/Si(cid:12) = 1.07, derived from the adopted solar values for oxygen, carbon, magnesium, and silicon, are also represented. As presented in Section 4, 100% of the sample with low- mass companions have an Mg/Si value of between 1.0 and 2.0, while 85% of the high-mass companion sample does, which means that Mg is equally distributed between py- roxene and olivine. We also find 15% of high-mass planet hosts with Mg/Si values below 1.0 so Mg and Si will form mainly orthopyroxenes, whereas the remaining Si will take other forms, such as feldspars or olivine. No low-mass com- panions with Mg/Si values lower than 1.0 were found and no stars with Mg/Si values greater than 2.0 were found either. We highlight the fact that all stars with low-mass companions present Mg/Si > 1.00, whereas the high-mass planetary sample can be found for a wide range of Mg/Si values, including Mg/Si < 1.00. Regarding C/O, 100% of stars with planets have C/O values lower than 0.8, meaning that Si will take solid form Article number, page 5 of 9 A&A proofs: manuscript no. 30743_corr Fig. 7: Cumulative Mg/Si distributions for single stars (red) and stars with planets (low-mass (green) and high-mass planets (blue)). and SiO2. Only ∼15% of our sample has C/O < as SiO4− 0.4 (15 stars). The exact composition will be ruled by the magnesium-to-silicon ratio. 4 Recent models by Carter-Bond et al. (2012); Marboeuf et al. (2014); Thiabaud et al. (2014, 2015a,b) suggest that elemental abundance ratios may suffer great variations when studied in planet host stars and in planetary atmo- spheres. They proposed that elemental abundance ratios such as Mg/Si and Fe/Si would show similar values in both planet hosts and planetary atmospheres. As for C/O, they proposed large differences between the star and the planets. They suggest that migration plays a key role when form- ing elemental abundance ratios in planetary atmospheres and the location of planet formation, as the total C/O ra- tios are governed by ices. Öberg et al. (2011) studied the influence of different snowlines of oxygen and carbon-rich species. They suggest a common region between the H2O and CO snowlines with giant planetary formation. These snowlines and the migration from the original birthplace can affect the abundance of volatiles, but not the abun- dance of refractory elements, such as Mg, Si, and Fe. This assumption was confirmed for hot-Jupiter hosts by Brewer et al. (2016b), where they show that the average planet C/O is super-stellar, but their large uncertainties do not exclude the possibility of the 1:1 relation for the stellar and planetary C/O. Stars with only low-mass planets are likely to be found in the 1 < Mg/Si < 1.5 regime, although mixed with stars with high-mass planets. As seen in Fig. 8, there is a gap for stars with 1.1 < Mg/Si < 1.3 and C/O < ∼0.5. As noted in Section 2, stars without planets are likely to be stars with undetected planets. Also, as presented in Section 4, there is a dependence in temperature for hotter stars. Since smaller planets will be harder to detect around these hotter stars, our results can be suffering from this effect. More knowledge about these stars and their plausible planetary companion is needed to constrain planetary formation and its relation with elemental abundance ratios more clearly. Article number, page 6 of 9 Fig. 8: Top panel: C/O vs. Mg/Si. Red circles refer to single stars (400 stars), and green dots refer to stars harbouring planets (99 stars). Middle panel: same as the top panel but only for stars with Teff < 6100K (312 single stars, 19 low-mass planet hosts and 65 high- mass planet hosts) . Bottom panel: same as top panel but for solar analogues, with Teff = Teff,(cid:12) ± 300 K, logg = logg(cid:12)±0.2 dex and [Fe/H] = [Fe/H](cid:12)±0.2 K (58 single stars, 2 low-mass planet hosts and 9 high-mass planet hosts). In all panels, an orange star represents solar values. Suárez-Andrés, L. et al.: C/O vs. Mg/Si ratios in solar type stars: The HARPS sample. 6. C/O, Mg/Si, and masses We studied C/O and Mg/Si as a function of mass of plane- tary companion (the most massive planet in case of multiple planets). In Fig. 9 we can see a slight dependence on plan- etary mass, as the slopes are not null (-0.04 for Mg/Si and 0.04 for C/O), but the elemental abundance ratio errors do not allow us to make any further assumptions. As proposed by Thiabaud et al. (2015b), Mg/Si in stars will give a direct information about the composition of the planet, as no differences are expected between them. Our stellar Mg/Si values can be translated as planetary Mg/Si (see Santos et al. 2015; Dorn et al. 2015). Regarding C/O, and given the indirect relation between the star and the planet, our results are an estimation of the C/O that could be found in the planet. For example, WASP-12b shows a high C/O, above 1 (Kreidberg et al. 2015), while its host star shows C/O = 0.48 (Teske et al. 2014). Brewer et al. (2016b) have proposed that planetary C/O will be super-stellar (while O/H could be sub-stellar) in hot-Jupiters. (2015) for a sample of 589 stars, the dependence of [Mg/Si] on metallicity, that is the Galactic evolution of [Mg/Si] ra- tio, may play an important role in the internal structure and composition of terrestrial planets. Overall, they con- cluded that stars formed at different times and places in the Galaxy have a different probability of forming low-mass planets, and the composition of the formed planets will also depend on the chemical composition of the environment in which they formed. The dependence of [C/O] on metallicity, that is, the Galactic evolution of [C/O], may play an important role in the internal structure and composition of high-mass plan- ets (as Mg/Si does for low-mass planets (see Adibekyan et al. 2015)). Following this statement, we looked for possible relations between [C/O] and the masses of planetary com- panions. We separated the planetary population into the same two groups as in the previous analysis (low-mass and high-mass planets). In Fig. 10 we show the dependence of [C/O] on the metallicity for stars with and without detected planets. This figure shows that C/O has a clear dependance on [Fe/H]. To remove the trend of [C/O] with GCE we fitted all our data points (all stars with and without detected planets) with a quadratic dependence on metallicity and then subtracted the fit. Mean squared deviation for the fit is 0.06.3[C/O]corr represents the [C/O] after substracting the fitted values. Fig. 9: Mg/Si and C/O elemental abundance ratios as a function of the planetary mass. A correlation for the final sample (stars with Teff < 6100K) is provided for both sets, C/O and Mg/Si. [C/O] as a function of Fig. 10: for stars with and without detected companions. Trend line provides a quadratic fit to all the data points. [Fe/H] 7. Galactic trends in [C/O] Elemental abundance ratios can help us to understand how the planets are formed. In astronomy, solar ratios (e.g. [C/O]) are more frequently used, as opposed to absolute ra- tios (e.g. C/O). However, in order to investigate if the C/O ratios are affected by Galactic chemical evolution (GCE), (as Mg/Si is, see Adibekyan et al. 2015), we need to rely on solar ratios ([C/O]), therefore we present in this Section a study of [C/O] for our sample. Several works have already discussed the possible differ- ences in individual elemental abundances and [X/Fe] abun- dances ratios between stars without detected planets and stars hosting low-mass and high-mass planets (Adibekyan et al. 2012a,c, 2015). As presented in Adibekyan et al. Distributions of [C/O] for all the presented sub-samples are shown in Fig. 11. A slight difference can be found be- tween the centres of each Gaussian fitting, but they are within errors (see Table 2). To obtain a clearer picture, we tested the sample using a two Kolmogorov-Smirnov (K–S) test (see Table 5 and Fig. 12). The null hypothe- sis is rejected at level α=0.1 if our results are higher than Dnn(cid:48)=0.29 and Dnn(cid:48)=0.15, for the low-mass and high-mass samples, respectively. In the first case, [C/O], above the imposed threshold limit, the probabilities of similarity are 4% and 4.09E-03% for the low and high-mass sample, re- spectively, so we can assume that the samples do not come 3 [C/O]=-0.45*[Fe/H]2+0.38*[Fe/H]-0.08 Article number, page 7 of 9 A&A proofs: manuscript no. 30743_corr Fig. 11: [C/O] distributions before (upper panel) and after ([C/O]corr, lower panel) correcting for the Galactic chemical evolution (GCE) effects. Fig. 12: Cumulative distributions for [C/O], before (upper panel) and after (lower panel) the GCE correction. from the same distribution. In the second case, [C/O]corr, we cannot reject the null hypothesis, but, probabilities are 17% and 38%, respectively. Although we cannot assume any conclusion due to the result being below the threshold limit, we can see that probabilities of similarity rises signif- icantly for the high-mass sample. The dependence of [C/O] on metallicity may play an important role in the internal structure and composition of high-mass planets (as Mg/Si does for low-mass planets (see Adibekyan et al. 2015)) but more studies with more data are required to confirm or dismiss this assumption. Table 5: K–S test for GCE and non-GCE corrected samples. Sample LMP - NP HMP - NP [C/O] 0.31 0.28 [C/O]corr 0.23 0.14 8. Summary and conclusions We present a detailed study of the C/O and Mg/Si elemen- tal abundance ratios for 499 solar-type stars observed with the HARPS high-resolution spectrograph. In our sample, 99 of 499 are planet-hosts and 400 stars have no known planetary companion. All the stars within our sample have Article number, page 8 of 9 effective temperatures between 5250 K and 6666 K, metal- licities from −0.59 to 0.55 dex and surface gravities from 3.81 to 4.82 dex. We separated the planet population into two groups to test for possible relations between elemental abundance ratios and the masses of planetary companions: low-mass planets (LMP; with masses less than or equal to 30 M⊕) and high-mass planets (HMP; with masses greater than 30 M⊕). All samples show the same distribution in a histogram, with al centre-fit similar within errors. Regarding the planetary sample, we cannot discern the high-mass from the low-mass sample. We studied the probability of the samples being drawn from the same distribution by applying a K–S test to our samples. For the C/O case, our result suggest no similar- ity between the samples. For Mg/Si, we cannot reject the null hypothesis, so we have to assume that our samples come from the same distribution. Overall, 99% of our sam- ple present C/O < 0.8, while all our confirmed planet host stars present C/O values below 0.8 (with peaks around 0.47), suggesting that the composition of the bulk of the planets should be SiO4− 4 and SiO2, serving as seeds for Mg silicates. ∼15% of our sample has C/O < 0.4. Regarding Mg/Si, and after separating our planetary sample in low- mass and high-mass companions, we found that none of our low-mass planet host sample present Mg/Si values lower than one, while 15% of the high-mass sample does, sug- gesting a composition of pyroxene and feldspars for this sample. Suárez-Andrés, L. et al.: C/O vs. Mg/Si ratios in solar type stars: The HARPS sample. 100% of our low-mass sample present Mg/Si values be- tween one and two, so an equal proportion of olivine and pyroxene is expected. 85% of the high-mass sample present these values. These results agree (within the errors and tak- ing into account different solar reference values) with recent studies of C/O and Mg/Si in the solar neighbourhood, such as Brewer & Fischer (2016), as they obtained peaks for dis- tributions at ∼0.5 for C/O and ∼1.1 for Mg/Si. Given these results, several different planet compositions could be found in planet hosts that meet these C/O and Mg/Si restrains. In the last few years, several models have suggested that C/O in planet host stars and planetary atmospheres is not the same, owing to migrations and snowlines, although more observations of atmospheres in transiting exoplanets are required to confirm these models (Carter-Bond et al. 2012; Öberg et al. 2011; Marboeuf et al. 2014; Thiabaud et al. 2014, 2015a,b). These ratios give a suggestion of possible planetary abundances. As there is a direct relation between host star and planet abundances for Mg/Si, we can confirm the com- position of the planets. Most of our planetary sample (100% of the low-mass and 85% of the high-mass sample) will have pyroxene and olivine as main components, while the other 15% of high-mass companions will have a composition based on orthopyroxene and minerals such as feldspar or olivine. For C/O, as the elements composing this ratio are sen- sitive to icelines, the observed planetary values can be up to three-four times the solar value, as recent observations suggest (Brewer et al. 2016b). Planetary C/O values can be up to four times the proposed stellar values. thank the for his/her We studied [C/O] and its relation with metallicity (as Mg/Si is affected, see Adibekyan et al. 2015). We do not find a clear relation between high-mass planets and [C/O], but our results suggest that more data is needed to con- firm this assumption. Our results for [C/O]corr are below the threshold limit but suggesting a significant increase for the probability of similarity between the high-mass and the comparison star samples. Acknowledgements. We thor- ough review and highly appreciate the comments and sugges- tions, which significantly contributed to improving the qual- ity of the publication. J.I.G.H. acknowledges financial support from the Spanish Ministry of Economy and Competitiveness (MINECO) under the 2013 Ramón y Cajal programme MINECO RYC-2013-14875, and the Spanish ministry project MINECO AYA2014-56359-P. V.Zh.A., E.D.M., N.C.S. and S.G.S. acknowl- edge the support from Fundação para a Ciência e a Tecnologia (FCT) through national funds and from FEDER through COM- PETE2020 by the following grants UID/FIS/04434/2013 & POCI-01- 0145-FEDER-007672, PTDC/FIS-AST/7073/2014 & POCI-01-0145- FEDER-016880 and PTDC/FIS-AST/1526/2014 & POCI-01-0145- FEDER-016886. V.Zh.A., E.D.M., N.C.S. and S.G.S. also acknowl- edge the support from FCT through Investigador FCT contracts IF/00650/2015, IF/00169/2012/CP0150/CT0002 and IF/00028/2014/CP1215/CT0002. This work has made use of the VALD database, operated at Uppsala University, the Institute of As- tronomy RAS in Moscow, and the University of Vienna. IF/00849/2015/, reviewer Adibekyan, V., Santos, N. C., Figueira, P., et al. 2015, A&A, 581, L2 Adibekyan, V., Gonçalves da Silva, H. M., et al. 2017, submitted Ali-Dib, M., Mousis, O., Petit, J.-M., & Lunine, J. I. 2014, ApJ, 785, 125 arXiv:1612.07355 Anders, E., & Grevesse, N. 1989, Geochim. Cosmochim. Acta, 53, 197 Bergemann, M., Collet, R., Amarsi, A. M., 2016, al. et Bertran de Lis, S., Delgado Mena, E., Adibekyan, V. Z., Santos, N. C., & Sousa, S. G. 2015, A&A, 576, A89 Bond, J. C., O'Brien, D. P., & Lauretta, D. S. 2010a, ApJ, 715, 1050 Bond, J. C., Lauretta, D. S., & O'Brien, D. P. 2010b, Icarus, 205, 321 Brewer, J. M., & Fischer, D. A. 2016a, ApJ, 831, 20 Brewer, J. M., Fischer, D. A., & Madhusudhan, N. 2016b, arXiv:1612.04372 Caffau, E., Ludwig, H.-G., Steffen, M., et al. 2008, A&A, 488, 1031 Caffau, E., Ludwig, H.-G., Bonifacio, P., et al. 2010, A&A, 514, A92 Carter-Bond, J. C., O'Brien, D. P., & Raymond, S. N. 2012, ApJ, 760, 44 ApJ, 725, 2349 Delgado Mena, E., Israelian, G., González Hernández, J. I., et al. 2010, Dorn, C., Khan, A., Heng, K., et al. 2015, A&A, 577, A83 Elser, S., Meyer, M. R., & Moore, B. 2012, Icarus, 221, 859 Faria, J. P., Santos, N. C., Figueira, P., et al. 2016, A&A, 589, A25 Figueira, P., Santos, N. C., Pepe, F., Lovis, C., & Nardetto, N. 2013, A&A, 557, A93 Fischer, D. A., & Valenti, J. 2005, ApJ, 622, 1102 Gonzalez, G. 1997, MNRAS, 285, 403 Gonzalez, G. 2006, PASP, 118, 1494 González Hernández, J. I., Israelian, G., Santos, N. C., et al. 2010, González Hernández, J. I., Delgado-Mena, E., Sousa, S. G., et al. 2013, ApJ, 720, 1592 A&A, 552, A6 Haywood, M. 2009, A&A, 482, 673 Haywood, M. 2009, ApJ, 698, L1 Konopacky, Q. M., Barman, T. S., Macintosh, B. A., & Marois, C. 2013, Science, 339, 1398 Kreidberg, L., Line, M. R., Bean, J. L., et al. 2015, ApJ, 814, 66 Lo Curto, G., Mayor, M., Benz, W., et al. 2010, A&A, 512, A48 Lodders, K. 2003, ApJ, 591, 1220 </pre> Madhusudhan, N., Lee, K. K. M., & Mousis, O. 2012, ApJ, 759, L40 Marboeuf, U., Thiabaud, A., Alibert, Y., Cabral, N., & Benz, W. 2014, A&A, 570, A36 Mayor, M., Pepe, F., Queloz, D., et al. 2003, The Messenger, 114, 20 Mayor, M., Marmier, M., Lovis, C., et al. 2011, arXiv:1109.2497 Morgan, J. W., & Anders, E. 1980, Proceedings of the National Academy of Science, 77, 6973 Moses, J. I., Madhusudhan, N., Visscher, C., & Freedman, R. S. 2013, ApJ, 763, 25 Nissen, P. E. 2013, A&A, 552, A73 Nissen, P. E., Chen, Y. Q., Carigi, L., Schuster, W. J., & Zhao, G. 2014, A&A, 568, A25 Öberg, K. I., Murray-Clay, R., & Bergin, E. A. 2011, ApJ, 743, L16 Petigura, E. A., & Marcy, G. W. 2011, ApJ, 735, 41 Portinari, L., Chiosi, C., & Bressan, A. 1998, A&A, 334, 505 Santos, N. C., Israelian, G., & Mayor, M. 2001, A&A, 373, 999 Santos, N. C., Israelian, G., & Mayor, M. 2004, A&A, 415, 1153 Santos, N. C., Mayor, M., Bonfils, X., et al. 2011, A&A, 526, A112 Santos, N. C., Adibekyan, V., Mordasini, C., et al. 2015, A&A, 580, Sousa, S. G., Santos, N. C., Mayor, M., et al. 2008, A&A, 487, 373 Sousa, S. G., Santos, N. C., Israelian, G., et al. 2011a, A&A, 526, L13 AA99 Sousa, S. G., Santos, N. C., Israelian, G., Mayor, M., & Udry, S. 2011b, Suárez-Andrés, L., Israelian, G., González Hernández, J. I., et al., A&A, 533, AA141 2016, A&A, 591, A69 Suárez-Andrés, L., I., Adibekyan, V. Zh., Delgado Mena, E., Sousa, S., Santos, N. C., 2017, A&A, 599, A96 Israelian, G., González Hernández, J. Teske, J. K., Cunha, K., Smith, V. V., Schuler, S. C., & Griffith, C. A. Thiabaud, A., Marboeuf, U., Alibert, Y., et al. 2014, A&A, 562, A27 Thiabaud, A., Marboeuf, U., Alibert, Y., Leya, I., & Mezger, K. 2015a, Thiabaud, A., Marboeuf, U., Alibert, Y., Leya, I., & Mezger, K. 2015b, 2014, ApJ, 788, 39 A&A, 574, A138 A&A, 580, A30 Tsantaki, M., Sousa, S. G., Adibekyan, V. Z., et al. 2013, A&A, 555, References Adibekyan, V. Z., Santos, N. C., Sousa, S. G., et al. 2012a, A&A, 543, Adibekyan, V. Z., Sousa, S. G., Santos, N. C., et al. 2012b, A&A, 545, Adibekyan, V. Z., Delgado Mena, E., Sousa, S. G., et al. 2012c, A&A, A89 AA32 547, A36 Adibekyan, V. Z., González Hernández, J. I., Delgado Mena, E., et A150 al. 2014, A&A, 564, L15 Article number, page 9 of 9
1008.4141
4
1008
2010-11-30T05:06:58
LHS6343C: A Transiting Field Brown Dwarf Discovered by the Kepler Mission
[ "astro-ph.EP", "astro-ph.SR" ]
We report the discovery of a brown dwarf that transits one member of the M+M binary system LHS6343AB every 12.71 days. The transits were discovered using photometric data from the Kelper public data release. The LHS6343 stellar system was previously identified as a single high-proper-motion M dwarf. We use high-contrast imaging to resolve the system into two low-mass stars with masses 0.45 Msun and 0.36 Msun, respectively, and a projected separation of 55 arcsec. High-resolution spectroscopy shows that the more massive component undergoes Doppler variations consistent with Keplerian motion, with a period equal to the transit period and an amplitude consistent with a companion mass of M_C = 62.8 +/- 2.3 Mjup. Based on an analysis of the Kepler light curve we estimate the radius of the companion to be R_C = 0.832 +/- 0.021 Rjup, which is consistent with theoretical predictions of the radius of a > 1 Gyr brown dwarf.
astro-ph.EP
astro-ph
Draft version August 9, 2018 Preprint typeset using LATEX style emulateapj v. 11/10/09 LHS 6343 C: A TRANSITING FIELD BROWN DWARF DISCOVERED BY THE KEPLER MISSION1 John Asher Johnson2,3, Kevin Apps4, J. Zachary Gazak5, Justin Crepp2, Ian J. Crossfield6, Andrew W. Howard7, Geoff W. Marcy7, Timothy D. Morton2, Carly Chubak7, Howard Isaacson7 Draft version August 9, 2018 ABSTRACT We report the discovery of a brown dwarf that transits one member of the M+M binary system LHS 6343 AB every 12.71 days. The transits were discovered using photometric data from the Kelper public data release. The LHS 6343 stellar system was previously identified as a single high-proper- motion M dwarf. We use adaptive optics imaging to resolve the system into two low-mass stars with masses 0.370±0.009 M⊙ and 0.30±0.01 M⊙, respectively, and a projected separation of 0.′′55. High-resolution spectroscopy shows that the more massive component undergoes Doppler variations consistent with Keplerian motion, with a period equal to the transit period and an amplitude consistent with a companion mass of MC = 62.9 ± 2.3 MJup. Based on our analysis of the transit light curve we estimate the radius of the companion to be RC = 0.833 ± 0.021 RJup, which is consistent with theoretical predictions of the radius of a > 1 Gyr brown dwarf. 1. INTRODUCTION Situated on the mass continuum between planets and hydrogen-burning stars are objects commonly known as brown dwarfs, with masses spanning approximately 13MJup up to 80 MJup(assuming Solar metallicity). Since the first discoveries of these substellar objects fifteen years ago (Oppenheimer et al. 1995; Basri et al. 1996), various surveys have found additional examples in num- bers exceeding the population of known exoplanets8. However, despite the large sample of brown dwarfs, very little is known about the physical properties or forma- tion mechanism(s) of these substellar objects (e.g. Basri 2006; Burgasser et al. 2007; Liu et al. 2008). Most known brown dwarfs have been discovered as soli- tary objects by wide-field, near-infrared (NIR) imaging surveys(e.g. Martin et al. 1997; Burgasser et al. 1999; Lawrence et al. 2007; Delorme et al. 2008). Their identi- fication is often based on spectral typing, with physical parameters derived from comparing photometric mea- surements to substellar evolutionary models. Knowledge beyond spectral typing is limited by the difficulty in mod- eling the complex molecular features that dominate the spectra of cool dwarfs (e.g. Allard et al. 2001; Cush- ing & Vacca 2006), a problem that persists above the hydrogen-burning mass limit for M-type dwarfs (Maness et al. 2007; Johnson & Apps 2009). The preferred method of measuring physical properties [email protected] 1 Based on observations obtained at the W.M. Keck Obser- vatory, which is operated jointly by the University of California and the California Institute of Technology. Keck time has been granted by Caltech, the University of California and NASA. 2 Department of Astrophysics, California Institute of Technol- ogy, MC 249-17, Pasadena, CA 91125 3 NASA Exoplanet Science Institute (NExScI) 4 Cheyne Walk Observatory, 75B Cheyne Walk, Horley, Sur- rey, RH6 7LR, United Kingdom 5 Institute for Astronomy, University of Hawai'i, 2680 Wood- lawn Drive, Honolulu, HI 96822 6 Department of Physics and Astronomy, University of Cali- fornia Los Angeles, Los Angeles, CA 90095 7 Department of Astronomy, University of California, Mail Code 3411, Berkeley, CA 94720 8 http://spider.ipac.caltech.edu/staff/davy/ARCHIVE/ of substellar objects such as masses, compositions and ages, is to study examples that are physically associated with brighter main-sequence stars. By assuming both the brown dwarf and its host star formed at the same time from the same molecular cloud, ages and chemical composition of the companion can be tied to the prop- erties of the brighter, more easily characterized compo- nent (Liu & Leggett 2005; Bowler et al. 2009). However, brown dwarfs in these favorable "benchmark" configu- rations are found in numbers far below the sample of exoplanets, despite their relative ease of detection com- pared to planet-mass companions. This observed feature of the substellar mass distribution of bound companions is known as the "brown dwarf desert," and the barren region extends over a wide swath around stars, extend- ing from ≈ 0.05 AU out to hundreds of AU (Marcy & Butler 2000; McCarthy & Zuckerman 2004; Grether & Lineweaver 2006; Johnson 2009). The existence of a deep minimum in the mass con- tinuum between stars and planets suggests that distinct formation mechanisms operate at either mass extreme, one for stellar objects and one for planets. However, the scarcity of objects in the brown dwarf desert makes it difficult to determine where this line should be drawn. For example, it is unclear whether a 20 MJup object in orbit around a main-sequence star formed like a mas- sive planet, or instead should be considered part of the extreme low-mass tail of the stellar initial mass func- tion (Kratter et al. 2010). Furthermore, an issue as fundamental as the mass-radius relationship below the hydrogen-burning limit is largely unconstrained by obser- vations. Understanding the nature and origins of brown dwarfs requires a much larger sample of detections. One promising avenue for increasing the sample of well- characterized substellar companions is through wide-field photometric transit surveys. Since the radii of objects are roughly constant from 1 -- 100 MJup (Baraffe et al. 1998), transit surveys are uniformly sensitive to com- panions throughout the entire brown dwarf desert. The transit light curve, together with precise radial veloci- ties (RV), provide both the absolute mass (as opposed to minimum mass, M sin i) and radius of the companion, 2 thereby directly testing the predictions of interior struc- ture models (Burrows et al. 1997; Torres et al. 2008). Once transits have been discovered, the door is opened up a wealth of follow-up opportunities that can measure properties such as the brown dwarf's albedo, temper- ature distribution, emission spectrum and atmospheric composition (see, e.g. the review by Charbonneau et al. 2005). Further, studying the distribution of physical characteristics of companions, and their relationships to the characteristics of their host stars, can inform theo- ries of the origins of brown dwarfs in the same way that the statistical properties of exoplanets inform theories of planet formation (Fischer & Valenti 2005; Torres et al. 2008; Johnson et al. 2010). In this contribution we present the discovery and char- acterization of a transiting brown dwarf orbiting a nearby low-mass star in the Kepler field. 2. PHOTOMETRIC OBSERVATIONS 2.1. Kepler Photometry The K epler space telescope is conducting a continuous photometric monitoring campaign of a target field near the constellations Cygnus and Lyra. A 0.95-m aperture Schmidt telescope feeds a mosaic CCD photometer with a 10◦ × 10◦ field of view (Koch et al. 2010; Borucki et al. 2010). Data reduction and analysis is described in Jenk- ins et al. (2010b) and Jenkins et al. (2010c), and the photometric and astrometric data were made publicly available as part of the first-quarter (Q0-Q1) data re- lease. The Q0 data have a time baseline UT 2009 May 5-11, and the Q1 data span UT 2009 May 13 through 2009 June 16. Among the 156,000 Long Cadence stellar targets in the Kepler field is the nearby, high-proper-motion M dwarf LHS 6343 (α = 19h10m14.31s, δ = +46◦57m25.0s; Reid et al. 2004). The photometric properties as listed in the Kepler Input Catalog (KIC; Batalha et al. 2010) are given in Table 1. The Q0-Q1 photometric time series of LHS 6343 contains a total of 2115 brightness measure- ments with a 29.4 minute cadence and a median internal measurement precision of 7× 10−5 (Jenkins et al. 2010c). As part of a study of the photometric variability of the closest stars in the Kepler field, one of us (K.A.) noted that the light curve of LHS 6343 exhibits four deep, pe- riodic dimming events spaced by 12.71 days. The light curve depths are constant to 0.3 mmag and exhibit no ob- vious additional dimming at intermediate periods, con- sistent with the signal of a transiting planet-sized ob- ject. The astrometry shows no shift in the center of light greater than 1 millipixel (4 mas, Jenkins et al. (2010a)), and there are no secondary eclipses evident at intermedi- ate phases. It is therefore unlikely that the source of the dimming events is a background eclipsing binary (EB). Examination of the Palomar Observatory Sky Survey (POSS) images shows no other stars at the current posi- tion of LHS 6343, further ruling false-positives involving an EB. The closest star in the archival images is 7.′′0 to the West of LHS 6343, near the edge of the Kepler photo- metric aperture. However, the star is only 4% as bright as LHS 6343, meaning that if it is an eclipsing binary it would have to nearly disappear to replicate the observed transit signal. This situation is ruled out by the lack of a large photocenter shift seen in the astrometric measure- TABLE 1 Observed Properties of LHS 6343 Parameter Value Source α δ µα (mas yr−1) µδ (mas yr−1) g′ r′ i′ Btot Vtot KP,tot Jtot Htot KS,tot ∆J ∆H ∆KS JA JB HA HB KS,A KS,B 19 10 14.33 +46 57 25.5 -145 -401 14.03 ± 0.02 13.06 ± 0.02 12.07 ± 0.02 KIC KIC KIC KIC KIC KIC KIC 15.009 ± 0.025 13.435 ± 0.018 13.104 ± 0.04 KIC 9.570 ± 0.021 8.972 ± 0.027 8.695 ± 0.011 2MASS 2MASS 2MASS PHARO PHARO PHARO 0.49 ± 0.05 0.48 ± 0.05 0.45 ± 0.06 10.10 ± 0.04 10.59 ± 0.06 9.51 ± 0.04 9.99 ± 0.07 9.25 ± 0.05 9.70 ± 0.08 ments. The Kepler photometric measurements phased at the 12.71 day period are shown in Fig. 1. 2.2. Nickel Z-band Photometry We observed the transit event predicted to occur on UT 2010 June 29 using the 1-m Nickel telescope at Lick Observatory on Mt. Hamilton, California. We used the Nickel Direct Imaging Camera, which uses a thinned Loral 20482-pixel CCD with a 6.3′ square field of view (Johnson et al. 2008). We observed through a Gunn Z filter, used 2 × 2 binning for an effective pixel scale of 0.′′37 pixel−1, and a constant exposure time of 75 sec- onds. We used the slow readout mode, with 34 s between exposures to read the full frame and reset the detector. The conditions were clear with ∼ 1.′′0 seeing. We began observing as soon as possible after sunset at an airmass of 1.4 and observed continuously for 5.4 hours bracketing the predicted transit midpoint, ending at an airmass of 1.05. We measured the instrumental magnitude of LHS 6343 with respect to the four brightest comparison stars in the field using an aperture width of 23 pixels and a sky annulus with an inner and outer radius of 28 and 33 pix- els, respectively. We converted the Nickel timestamps to BJDUT C using the techniques of Eastman et al. (2010) to be consistent with the Kepler data. The Nickel pho- tometric measurements phased at the 12.71 day period are shown in Fig. 1. 3. RADIAL VELOCITIES AND ORBIT SOLUTION We obtained spectroscopic observations of LHS 6343 at Keck Observatory using the HIRES spectrometer with a resolution of R ≈ 55, 000, with the standard iodine-cell setup used by the California Planet Survey (Howard et al. 2010). The transit depth, together with a rough stel- lar radius estimate of 0.4 R⊙ appeared consistent with a planet with a radius of ∼ 0.7 RJup. In anticipation of a low-amplitude Doppler signal we initially used 45 minute exposures through the iodine cell and C2 decker Kepler 1.00 Nickel Z−band LHS6343 y t i s n e n t I d e z i l a m r o N 0.98 0.96 0.94 0.92 Kepler Residuals (x5) Nickel Residuals 0.90 −3 −2 −1 0 Time Since Mid−Transit [hrs] 1 2 3 Fig. 1. -- The Kepler (upper) and Nickel (blue) light curves, phased at the photometric period. The best-fitting light curve models are shown for each data set (see § 5.3), and the residuals are shown beneath each light curve. The Kepler residuals have been multiplied ×5 for clarity. for sky-subtraction. The resulting signal-to-noise ratio (S/N) was ≈ 90 at 5500 A, near the center of the iodine absorption region. A cross-correlation analysis of the first two observa- tions revealed two peaks separated by ∼ 10 km s−1. The lack of a coincident background star in the POSS images suggests that the second set of lines must be from a physically-associated binary companion and that LHS 6343 is a double-lined spectroscopic binary. Adap- tive optics observations described in § 4 confirmed the existence of a wide binary companion at a projected sep- aration of 0.′′55. Hereafter, we refer to the more massive component as Star A, and the less massive component as Star B. In order to discern which component of the binary sys- tem is transited by a companion we made subsequent HIRES observations with a position angle oriented along the binary axis to ensure the light from both stars fell within the slit. The cross-correlation analysis of the third observed spectrum revealed that the deeper set of ab- sorption lines shifted by ∼ 10 km s−1 with respect to the first observation, indicating that Star A is orbited by a massive companion. Our remaining HIRES spectra were obtained without the iodine cell and with 3-minute exposure times. We measured the radial velocity of Star A using the cross- correlation analysis described by Johnson et al. (2004). However, we modified the procedure by constructing a double-lined cross-correlation template. We began with an iodine-free spectrum of the HD 265866 (M3V) for star 3 K = 9.6 km s−1 e = 0.056 ] 1 − s m k [ y t i c o e V l l i a d a R 15 10 5 0 −5 −10 C − O 0 10 20 30 1 0 −1 0 10 20 JD − 2465373 30 Fig. 2. -- Keck/HIRES radial velocity measurements of LHS 6343 A. The best-fitting Keplerian orbit solution is shown as a dashed line. The systemic velocity, γ = −46.0 ± 0.2 km s−1, has been subtracted for clarity. The lower panel shows the RV residuals about the best-fitting orbit. A, and added to this template a scaled, shifted version of itself to represent the spectrum of star B. For each observation we adjusted the scaling and Doppler-shift of spectrum B with respect to spectrum A until the cross- correlation peak was maximized. We then measured the centroid of the optimized cross-correlation function by fitting a parabola to the region near the resulting single peak. We corrected for shifts in the HIRES detector by using the telluric lines in the 630 nm α-band as a wavelength reference. We measured the absolute radial velocities of LHS 6343 A with respect to the absolute radial velocity of HD 265866 (M3V; Vr = +22.97 ± 0.50 km s−1; Chuback et al., in prep), and the resulting RV is corrected to the Solar System barycenter using the velocity corrections computed by the CPS data reduction pipeline. The full time series of radial velocity measurements is displayed in Fig. 2. The radial velocities are also listed in Table 2, along with the heliocentric Julian Dates of observation and internal measurement errors. We searched for the best-fitting Keplerian orbit solu- tion using the partially linearized, least-squares fitting procedure described in Wright & Howard (2009) and im- plemented in the IDL software package RVLIN9. We fixed the period and mid-transit time based on the light-curve analysis in § 5.3, which leaves only four free parameters: the velocity semiamplitude (KA), argument of perias- tron (ω), systemic velocity (γ) and eccentricity. We find that the RVs are described well by a nearly circular or- bit (e = 0.056 ± 0.032) with a velocity semiamplitude K = 9.6 ± 0.3 km s−1. The full spectroscopic orbit is given in Table 5 and shown in Fig. 2. The parameter uncertainties were estimated using a bootstrap Monte Carlo algorithm. For each of 5000 real- izations of the data, the measured RVs are perturbed by adding residuals randomly drawn from about the best- fitting orbital solution, with replacement. We chose this technique over an MCMC analysis out of concern that our RV measurement uncertainty is dominated by sys- 9 http://exoplanets.org/code/ 4 TABLE 2 Radial Velocities for LHS 6343 JD RV -2440000 (km s−1) Uncertainty (km s−1) 15373.095 15373.998 15377.078 15377.099 15378.030 15379.052 15380.827 15380.831 15395.983 15396.970 15404.974 15405.821 15406.865 15407.854 -38.40 -37.51 -49.42 -49.56 -52.56 -56.39 -55.83 -54.84 -47.92 -42.63 -55.90 -56.32 -53.69 -51.31 0.21 0.17 0.50 0.45 0.41 0.44 0.48 0.47 0.57 0.51 0.59 0.46 0.51 0.57 Fig. 3. -- PHARO KS-band adaptive optics image of LHS 6343, showing the two M-type components of the system. Star A is to the upper right of the image and Star B is to the lower left. tematic errors related to imperfect treatment of the sec- ond set of absorption lines, rather than photon noise. Thus, instead of assuming the RVs are normally dis- tributed about the model, we use the residuals them- selves as an estimate of the noise model. 4. PALOMAR ADAPTIVE OPTICS IMAGING We acquired near-infrared images of LHS 6343 on UT 2010 June 29 using the Palomar 200-inch telescope adap- tive optics system and PHARO camera (Hayward et al. 2001; Troy et al. 2000). These diffraction-limited ob- servations spatially resolve the target into a binary, as shown in Fig. 3. We used these observations to calculate the relative brightness of each component of the visual bi- nary and to subsequently constrain the range of possible masses (see § 5). Our results for differential magnitudes in each of the J,H,KS filters are listed in Table 1. We used aperture photometry To measure the differen- tial magnitudes of the two stars. However, given their rel- atively small angular separation, special care was taken to account for cross-contamination between the compo- nents. To remove the majority of flux contributed by the neighboring star, we used the spatial symmetry in the images. The amount of contaminating starlight was esti- mated by summing the counts over a region the same size as the photometric aperture located on the side opposite the star of interest. Once the contaminating light from the neighbor star is removed, we find that the photometric precision is lim- ited to several percent by uncertainties resulting from the subtraction residuals, as well as CCD non-linearity and PSF centroiding. These errors sources contribute simi- larly to the overall uncertainty and were added in quadra- ture, neglecting any correlations between PSF centroid- ing errors and contamination removal, which we found to be comparatively small. To this, we also added in quadrature the standard deviation in the mean flux ra- tio of the companions over the 20 images acquired in each bandpass. Observations in the KS band have a slightly larger uncertainty than in J and H, since the binary sep- aration subtends a smaller angle on the sky in units of resolution elements. 5. STELLAR PROPERTIES The physical properties of the two stellar components of the wide binary, hereafter Star A and Star B, are of central importance to measuring the properties of the substellar companion, LHS 6343 C. The mass of the com- panion is related to the mass of Star A (MA) and the companion's radius depends on the stellar radius, RA. The luminosity of Star B is also important for the mea- surement of the companion's radius, as its contribution to the total flux of the system dilutes the transit depth. In fact, for the specific case of LHS 6343 the precision with which we can measure the companion radius will depend critically on our estimate of the "third light" con- tribution of Star B, rather than the photometric precision of the transit light curve (Irwin et al. 2010). Because low-mass stars spend nearly their entire lives close to the zero-age main sequence, their observed prop- erties are a function of two physical characteristics: mass and, to a lesser extent, metallicity. There are two widely- used methods estimating the masses of M dwarfs. The first involves a comparison of observed properties such as absolute magnitude and color index, or luminosity and effective temperature, to tabulated stellar evolu- tion models, such as those calculated by Baraffe et al. (1998). However, studies of low-mass eclipsing bina- ries (EB) have demonstrated that these models system- atically under-predict stellar radii (Ribas 2006; L´opez- Morales 2007), even in cases when stellar activity should play a minimal role in shaping stellar structure (Torres 2007). The other method makes use of empirical relationship between the near-infrared luminosity of a star and its mass, as parametrized by Delfosse et al. (2000). We use the empirical relationships almost exclusively in order to avoid any systematic errors in the stellar evolution mod- els. However, since the mass-luminosity relationships re- quire absolute magnitudes, the distance to the star must be known to accurately estimate the stellar mass. Unfor- tunately, there is no published trigonometric parallax for LHS 6343, and the spectroscopic parallax of Reid et al. TABLE 3 Polynomial Coefficients for Magnitude-Mass and Mass-Radius Relationships j 0 1 2 3 4 5 bJ bH 14.888 -74.375 376.72 -1089.6 1601.6 -935.37 13.211 -57.464 271.11 -762.12 1103.9 -640.80 bKS 13.454 -67.439 338.43 -976.27 1433.9 -838.03 bV bR -9.229 5.017 -0.6609 0.03314 ... ... 0.000 1.268 -1.013 0.9391 ... ... (2004) is unreliable because it is based on the total mag- nitude and colors of the binary system, rather than the individual stars. While the binarity of LHS 6343 in some regards poses a challenge, having two stars with the same age and chemical composition, together with the available pho- tometric measurements, provide a unique opportunity to determine the physical characteristics of the two compo- nents. As we will demonstrate, the luminosity difference between the two stars constrains the mass ratio, while the total luminosity constrains the total mass, distance and metallicity of the system. An additional constraint on the mass is provided by the shape of the transit light curve. The slope of the ingress/egress yields the scaled semimajor axis a/RA ≡ aR, which is related to the density of Star A through Kepler's third law (e.g. Seager & Mall´en-Ornelas 2003; Sozzetti et al. 2007; Winn 2008). However, the true value of aR depends on the true transit depth (RC /RA)2, which in turn is related to the amount of dilution in the light curve due to Star B. In what follows we first relate the masses of Star A and Star B to the available observables. We then estimate the flux contribution of Star B in the Kepler bandpass, which provides a refined estimate of the stellar masses. The iteration of this procedure yields the stellar param- eters of Star A, which allows us to estimate the physical characteristics of the transiting object LHS 6343 C. 5.1. Stellar Masses The most useful data available to us comprises seven photometric measurements: the total near-infrared mag- nitudes in the 2MASS catalog denoted by Ti where i = J, H, KS corresponds to the three bands, respectively; the magnitude differences ∆J, ∆H, ∆KS from our AO imaging, denoted by ∆i. We also have the total John- son V - and B-band magnitudes Vtot = 13.435 ± 0.018 and Btot = 15.009 ± 0.025. Our Vtot agrees well with the value listed in the TASS catalog, Vtot = 13.38 ± 0.24 (Droege et al. 2006). These magnitudes yield the system color (B − V )tot = 1.574 ± 0.031. The individual apparent magnitudes in the ith NIR band (JHK) of Star A (mi,A) and Star B (mi,B) are related to the total magnitudes Ti and magnitude differ- ences ∆i through mi,A = 2.5 log10(1 + 100.4∆i) + Ti mi,B = ∆i + mi,A (1) Eqns 1 give three NIR apparent magnitudes for each star. These six NIR photometric measurements can be related 5 to the stellar masses (MA, MB) through the following equations mi(M, d) = Mi(M ) + 5(log10 d − 1) (2) The functions Mi(M ) give the absolute magnitudes in the NIR bands as a function of stellar mass, M , as de- termined by inversion of the empirical relationships of Delfosse et al. (2000), which we approximate with the polynomial Mi(M ) = 5 Xj=0 bi,jM j (3) The coefficients {bi} are listed in Table 3 for the i = {J, H, KS} bands. Under the assumption that the binary components share the same chemical composition, they should reside at the same distance from the average main sequence in the {V − KS, MKS } plane (Johnson & Apps 2009, hereafter JA09). JA09 provide a relationship between a star's metallicity, [Fe/H] ≡ F , and its "height" above the Solar-neighborhood mean main sequence, ∆MK. Since the stars share the same composition the must lie on the same isometallicity contour: their V − KS colors must be consistent with the same value of ∆MK, while the individual V -band luminosities must be consistent with the measured Vtot. This constraint can be expressed as Vtot(MA, MB, d, F ) = MV(MA, F ) − 2.5 log10(1 + 100.4[MV(MA,F )−MV(MB ,F )]) + 5 log10 d − 5, (4) where the absolute V-band magnitude, MV(M, F ), is related to stellar mass M and metallicity F by inverting the photometric metallicity calibration of JA0910. We approximate this inversion using the polynomial MV(M, F ) = 3 Xj=0 bV,j(cid:20)MKS (M ) +(cid:18) F − 0.05 0.55 (cid:19)(cid:21)j (5) The coefficients {bV } are listed in Table 3. In addition to the apparent magnitudes, the tran- sit light curve provides an additional constraint on MA through the scaled semimajor axis a/RA ≡ aR. This quantity is related to the mass and radius of Star A, and the period and mass of LHS 6343 C, through aR(MA, MC, P ) =(cid:18) G 4π2(cid:19)1/3 M 1/3 RA(MA) A P 2/3(cid:18)1 + MC MA(cid:19)1/3 , (6) 10 We use the JA09 relationship rather than the Schlaufman & Laughlin (2010, SL10) calibration because the former provides a better match to the mean metallicity of the Solar neighborhood. In principle, the SL10 relationship would serve our purposes just as well since both V -band metallicity relationships provide a means of relating Vtot to the stellar masses under the constraint that both stars lie on the same isometallicity contour in the {V − KS, MKS } plane. The only difference is the exact value of [Fe/H], which will be ≈ 0.1 dex lower using the SL10 calibration. 6 where RA(M ) is a function relating the stellar radius and mass. We use a polynomial fit to the masses and radii of well-characterized ({σM , σR} < 3%) low-mass eclipsing binaries tabulated by Ribas (2006) RA(M ) = 3 Xj=0 bR,jM j (7) The coefficients {bR} are listed in Table 3. The optimal stellar parameters can be obtained by minimizing the fitting statistic χ2 tot = + 3 3 σmi,A (cid:18) mi,A − mi(MA, d) (cid:18) mi,B − mi(MB, d) (cid:19)2 Xi=1 (cid:19)2 Xi=1 +(cid:18) Vtot − Vtot(MA, MB, d, F ) (cid:19)2 +(cid:18) aR − aR(MA, MC, P ) σmi,B σVtot σaR (cid:19)2 (8) However, in order to obtain the best-fitting parameters we must determine the corrected values of the transit depth and aR by accounting for the flux contribution of Star B, as described in the following section. 5.2. Flux Contribution of Star B Since both stars fall within the Kepler and Nickel pho- tometric aperture, we must estimate the flux contribu- tion of Star B in both the KP and Z bandpasses. The transformation between the Johnson B and V magni- tudes and the Kepler magnitude KP is given in the Ke- pler Guest Observer web page11. However, while we were able to measure individual V magnitudes for both stars based on their total V magnitude and derived physical properties using Eqn. 5. However, there is no suitable relationship between the available observables and the individual B magnitudes. We are therefore forced rely on stellar model grids to estimate the B − V colors of stars A and B. For this task, we selected the "Basic Set" of stellar model grids from the Padova group for Solar composition and an age of 5 Gyr (Girardi et al. 2002), which give predictions for the Johnson B and V magnitudes as a function of stellar mass. We selected the Padova models because the Baraffe et al. (1998) model grids do not give fluxes in the B or Z bandpasses. We tested the reliability of the Padova model grids using a sample of 6 metal-rich stars from Johnson & Apps (2009). We found that the models under-predict the B − V colors of our calibration stars by 0.20 ± 0.07, independent of stellar mass over a range of approximately 0.1 M⊙to 0.6 M⊙. We applied this correction to the Padova models to obtain a refined relationship between stellar mass and B − V color, and the uncertainty of this offset was folded into the final uncertainties in the colors. 11 http://keplergo.arc.nasa.gov/CalibrationZeropoint.shtml Using the estimates of the B − V colors of Star A and Star B, together with the individual V -band magnitudes from Eqn. 5 we can estimate the relative magnitudes of the stars in the Kepler bandpass, ∆KP , as a function of the stellar masses. We then use this value of ∆KP to correct the Kepler light curve parameters. For the Nickel light curve, we use the SDSS z-band magnitude from the Padova model as a proxy for the Gunn Z filter. Our reliance on low-mass stellar model grids for the magnitude differences is less than ideal, particularly in optical bandpasses, because of the difficulty in properly treating molecular opacities at optical wavelengths (e.g. Baraffe et al. 1998). In the analysis that follows, we at- tempt to ameliorate this imperfect knowledge by using large uncertainties for ∆KP and ∆Z, which are propa- gated throughout our analysis and reflected in the con- fidence intervals for our derived system properties. We discuss the impact of our model-based magnitude differ- ences in § 6.2. 5.3. Light Curve Analysis We fitted the Kepler and Nickel light curves using the analytic eclipse model of Mandel & Agol (2002) to com- pute the integrated flux from the uneclipsed stellar sur- face as a function of the relative positions of the star and planet. The parameters were the period P , inclina- tion i, ratio of the companion and stellar radii RC /RA, the scaled semimajor axis aR, time of mid-transit Tmid, and the parameters describing the limb-darkening of the star. For the limb-darkening we used the quadratic ap- proximations of Sing (2010) and Claret (2004) for the Kepler and Z bands, respectively. In our model we fixed the quadratic terms u2 at the tabulated values for each filter, and allowed the linear term u1 to vary under a penalty of the form exp[−(x − µ2 x], which is added to the fitting statistic. This treatment of the limb dark- ening was made based on the additional structure seen in the residuals of the Kepler light curve fit. Allowing the limb-darkening parameters to float freely results in unphysical values. x)/σ2 To properly model the light curves we made two mod- ifications to the typical light curve analysis. First, we corrected for the third-light component by adjusting the normalized flux level from the Mandel & Agol light curve model, fmod, such that fcorr = fmod + 10−0.4∆KP 1 + 10−0.4∆KP (9) In our fitting procedure, we treat ∆KP as a free pa- rameter under the normally-distributed, prior constraint N (∆KP , σKP ) based on the estimate of the magnitude differences described in § 5.2. We use a similar prescrip- tion for the flux contribution of Star B to the Nickel Z bandpass. Our other modification involved rebinning the analytic light curve to match the 29.4-minute cadence of the data using a method similar to that of Kipping (2010). In addition to modeling the companion transit, we also fitted a slowly-varying function to the out-of-transit por- tion of the Nickel light curve to account for differential extinction from the Earth's atmosphere. To account for time-correlated noise in the data we used a Daubechies fourth-order wavelet decomposition likelihood function TABLE 4 LHS 6343 Ephemeris Tmid (BJD - 2450000.0) Tmid - Ephemeris Telescope 4957.216636 ± 0.000073 4969.930410 ± 0.000099 4982.64437 ± 0.00013 4995.35807 ± 0.00015 5376.77274 ± 0.00025 0.0000020 ± 0.00012 -0.000045 ± 0.00014 0.000093 ± 0.00016 -0.000028 ± 0.00018 -0.0000050 ± 0.00027 K K K K N Note. -- K -- Kepler, N -- Nickel Z-band following the technique described by Carter & Winn (2009). Wavelet decomposition provides increased confi- dence in the derived parameter uncertainties over the tra- ditional χ2 likelihood by allowing parameters that mea- sure photometric scatter (uncorrelated white noise σw, and 1/f red noise σr) to evolve as free parameters. The method recovers the traditional χ2 fitting statistic in the case of σr = 0, and when σw is fixed at a value equal to the characteristic measurement error. We determined the best-fitting model parameters and their uncertainties using a Markov Chain Monte Carlo (MCMC) analysis with a Gibbs Sampler (Geman & Ge- man 1984; Ford 2005; Winn et al. 2009). The MCMC fit- ting algorithm was implemented using the Transit Anal- ysis Package (TAP; Gazak & Johnson 2011, in prep.), a graphical user interface -- driven analysis tool written in the Interactive Data Language (IDL). We constructed 10 chains containing 106 links using initial conditions based on a simple least-squares fit to the phased pho- tometry. We chose step sizes such that 30 -- 40% of the steps are accepted. We discarded the initial 105 links in each parameter chain to allow for "burn-in" before com- bining the chains. The resulting chains of parameters form the posterior probability distribution, from which we select the 15.9 and 84.1 percentile levels in the cu- mulative distributions as the "1σ" confidence limits. In most cases the posterior probability distributions were approximately Gaussian, and we therefore report only symmetric error bars for simplicity. The final, iterative fitting procedure and derived parameters are presented in § 6 6. PHYSICAL PROPERTIES OF THE LHS 6343 STELLAR SYSTEM We solve for the physical parameters of the LHS 6343 system using the following iterative procedure: 1. Fit the light curves to obtain the scaled semimajor axis, aR. We initially use ∆Kp = 0.7, σKP = 0.1, ∆Z = 0.5, and σZ = 0.1 in Eqn. 9 and refine these values with subsequent iterations. 2. Estimate MC based on our Keplerian fit to the RV time series by solving M 3 C sin3 i = K 3P 2πG (MA + MC)2p1 − e2 (10) We initially assume MC ≪ MB, and relax this as- sumption as MA is revised. 3. Use aR and MC to estimate MA and MB by mini- mizing Eqn. 8. 7 Measured Value Model Value ) e u a v l g n i t t i f − t s e B ( / y t i t n a u Q 1.01 1.00 0.99 aR Vtot JA JB HA HB KA KB 0 1 2 3 Parameter 4 5 6 7 Fig. 4. -- Comparison of the observed stellar and transit proper- ties (filled circles with error bars), and the corresponding quanities predicted by our model (red swaths; Eqns 2, 4 and 6). In order to show all parameters on the same scale, the quantities have been divided by the best-fitting model values. 42 37 ] c p [ d 32 0.3 0.2 0.1 0.0 −0.1 −0.2 0.45 0.40 0.35 0.30 ] / H e F [ ] n u S M [ A M 0.25 0.30 0.35 0.25 0.30 0.35 0.3 0.2 0.1 0.0 −0.1 −0.2 0.45 0.40 0.35 0.30 32 37 42 0.45 0.40 0.35 0.30 0.25 0.30 MB [MSun] 0.35 32 37 d [pc] 42 −0.2 −0.1 0.0 [Fe/H] 0.1 0.2 0.3 Fig. 5. -- Each panel shows the joint, posterior pdfs for two parameters at a time, marginalized over the remaining parame- ters. The contours show the iso-probability levels corresponding to {68.2, 95, 99.7}% confidence. 4. Test for convergence in MA, MB, MC and ∆KP by comparing the current values to those of the previous iteration. If the change is larger than 10% of the parameter uncertainties, then go to step 1. We find that convergence is rapid, requiring only three iterations. Our best-fitting stellar parameters are MA = 0.370 ± 0.009 M⊙, MA = 0.30 ± 0.01 M⊙, d = 36.6 ± 1.1 pc and [Fe/H] = 0.04 ± 0.08. The fit results in χ2 tot = 2.1 with 8 data points and 4 free parameters, indicating that our fit is acceptable but that our photometric measurement uncertainties are likely overestimated. The best-fitting model values are compared to the observed quanities in 8 Fig. 4. The marginalized, posterior probability density functions (pdf) are shown in Fig. 5. Our analysis of the light curve results in RC /RA = 0.226±0.003 and aR = 45.3±0.6, with ∆KP = 0.74±0.10 and ∆Z = 0.5 ± 0.1. The best-fitting transit model is shown in Fig. 1 for both the Kepler (blue) and Nickel data (red), along with the residuals. For both data sets we recover σr = 0, consistent with no red noise contam- ination. However, we do see additional structure in the in-transit residuals of the Kepler fit. We do not know the source of this increased scatter. Because there are so few points, and because the data do not appear to be time-correlated during any single transit event, we found that the extra scatter is accounted for by allowing the fitting procedure to increase the white-noise component, σw. We find σw = 1.1×10−3 (1.8-minute cadence) for the Nickel data and σw = 1.1 × 10−4 (29.4-minute cadence) for the Kepler data. 6.1. The Radii of LHS 6343 A and LHS 6343 C Given the mass of Star A, we can estimate the stellar radius by evaluating Eqn. 7 which results in a stellar radius RA = 0.378± 0.008 R⊙. The radius ratio from the light curve analysis, together with RA yields the radius of LHS 6343 C, RC = 0.833±0.021 RJup. The other physical properties of the brown dwarf are listed in Table 5. 6.2. The Dependence of RC on the Flux Contribution of LHS 6343 B In § 5.2 we describe our method of estimating the flux contribution of Star B, as parametrized by the magni- tude differences ∆KP and ∆Z. These values are of cen- tral importance to the determination of the corrected values of a/RA and RC /RA as derived from the anal- ysis of the light curves, and hence the physical proper- ties of the brown dwarf. Unfortunately, we do not have spatially -- resolved photometry of the binary stars, so we were forced to rely on theoretical stellar models to es- timate the magnitude differences. Owing to incomplete knowledge of molecular opacities, particularly in opti- cal bandpasses, the model grids provide only rough es- timates of the true flux contribution of Star B. To en- capsulate this imperfect knowledge in our analysis, we used relatively large uncertainties (0.1 mag) for ∆KP and ∆Z. These errors are propagated throughout our MCMC analysis and are reflected in the final physical properties of the LHS 6343 system. As a test, we performed independent fits to the Ke- pler and Z-band light curves. We found that the transit properties from the two analyses agreed extremely well, differing by only a fraction of a σ in each value. Figure 6 illustrates the dependence of RC on ∆KP and ∆Z, and compares the value of RC measured from the indepen- dent light curve fits. Based on this figure we feel that our value of ∆Z is reasonable since its value shouldn't be less than ∆J = 0.49. Similarly, and ∆Z shouldn't ex- ceed our estimate of ∆KP , and ∆KP shouldn't be much larger than our measured ∆V = 0.74. These results, along with the close agreement between the two mea- surements of RC , provide additional confidence that our model-based magnitude differences are valid. 6.3. Searching for Transit Timing Variations ] p u J R [ C R 0.90 0.88 0.86 0.84 0.82 0.80 0.78 Kepler (KP−band) Nickel (Z−band) 0.2 0.4 0.6 ∆ Z, ∆ KP 0.8 1.0 Fig. 6. -- The dependence of RC on ∆KP (upper, blue) and ∆Z (lower, red), based on the marginalized posterior pdfs. The con- tours represent the 68.2% and 95% confidence regions. The two filled circles with error bars show the final values of RC , ∆KP and ∆Z from our analysis, and the error bars represent the correspond- ing 68.2% confidence bounds in each dimension. To measure the individual transit mid-times we fixed all of the global parameters (RC /RA, a/RA, P , i and the limb-darkening coefficients) and fitted each transit event separately using the MCMC algorithm described in § 5.3. Table 4 lists the time at the mid-point of each transit Tmid; the difference between the measured values and the those predicted by a linear ephemeris; and the formal measurement uncertainties which range from 10 -- 23 seconds. We see no statistically significant timing variations. 6.4. Limits on the Secondary Eclipse Depth To place an upper limit on the secondary eclipse depth we fitted a simplified transit model to the Kepler photom- etry near one-half phase away from the primary eclipse. We fixed the transit parameters from the fit to the pri- mary eclipse, along with e and ω the from the RV anal- ysis; assumed no limb darkening; and used final value of ∆KP = 0.74 ± 0.10. By allowing only the transit depth, σw and the terms describing the out-of-transit normal- ization to vary, we place a 95% upper limit on the eclipse depth of d < 6.5 × 10−5. 6.5. Limits on the System Age Measuring the ages of M-type stars is notoriously dif- ficult, except for the rare cases when stars are in clusters or associations, or show indications of extreme youth. The space motion of LHS 6343 rules out membership in any known moving group or cluster. We therefore must rely on spectroscopic and kinematic indicators that can at least tell us if a star is either younger or older than a few Gyr. Our HIRES spectra show remarkably low chromo- spheric activity in the Ca II K line (Fig. 7). None of the 140 low-mass stars on the CPS M dwarf survey have a chromospheric S values, as measured on the Mt. Wilson scale (Wright et al. 2004), as low as what we measure for LHS 6343 A. After attempting to correct for the ≈ 30% contamination from star B to the emission core, we es- Ca II K Line Gl876 x u F l Barnard's Star Gl445 LHS6343 3920 3925 3930 Wavelength [Å] 3935 3940 Fig. 7. -- HIRES spectra near the CaII K line for LHS 6343 and several M dwarfs with ages greater than 1 Gyr. timate S ≈ 0.4 for LHS 6343 A. The most chromospher- ically quite M dwarf in the CPS sample is Gl 445, with an average of S = 0.5. LHS 6343 was not detected by ROSAT12, but at 36.6 pc this only rules out extremely young ages less than ∼ 100 Myr. Our HIRES spectra exhibit narrow lines in- dicating Vrot sin i < 2 km s−1. No rotational modulation of star spots is seen in the Kepler light curve, indicating either a long rotation period (Prot & 40 days) or ex- tremely low spot coverage, or both. Thus, the age of the LHS 6343 system is most likely greater than 1-2 Gyr. 7. DISCUSSION LHS 6343,C is remarkably similar to the brown dwarf NLTT41135 B, recently discovered by the MEarth transit survey (Irwin et al. 2010). Both brown dwarfs orbit one component of a nearby M+M visual binary, both stellar systems are hierarchical triples, and both were discovered through transit photometry. However, owing to its more favorable orbital inclination, transit photometry provides a direct measurement of the radius LHS 6343 C. The large mass ratio between parent star and compan- ion argues against the formation of LHS 6343 C in the protoplanetary disk of LHS 6343 A. This is because there is likely not enough material in the disks of M dwarfs to build a 62.9 MJup companion, whether through core accretion (Laughlin et al. 2004) or disk instability (Boss 12 The Rontgensatellit (ROSAT) was a joint German, US, and British X-ray observatory operational from 1990 to 1999. 9 OGLE−TR−123B CoRoT−3b CoRoT−15b 0.5 Gyr 1 Gyr 5 Gyr WASP−30b OGLE−TR−122B LHS6343C ] p u J R i [ s u d a R 1.4 1.2 1.0 0.8 0.6 20 40 60 Mass [MJup] 80 100 Fig. 8. -- The masses and radii of known low-mass, transiting objects. Also plotted are the predicted mass-radius curves from the Baraffe et al. (1998) interior structure models for 0.5, 1, 5 Gyr, top to bottom. Not shown are the young, eclipsing brown dwarfs in the 2MASS 2053-05 system (Stassun et al. 2006), which have radii well above the plot range. 2006). It is therefore more likely that LHS 6343 C formed in a similar manner as LHS 6343 B, through the fragmen- tation of a portion of their natal molecular cloud. Irwin et al. (2010) make similar arguments regarding the origin of NLTT 41135 B. Prior to this year, the only other brown dwarfs with accurate radius measurements were the substellar com- ponents of the 2MASS J05352184-0546085 eclipsing bi- nary system Stassun et al. (2006), and Corot-3 b (Deleuil et al. 2008). The 2MASS J05352184-0546085 system was discovered in a young star-forming region, and the two brown dwarfs have very large radii for their masses (0.669 R⊙ and 0.611 R⊙) because they are still undergo- ing gravitational collapse. In the latter portion of this year, two other transiting, field brown dwarfs have been discovered by transit sur- veys. The ground-based Wide Angle Survey for Plan- ets discovered WASP-30 b (60.96 ± 0.89 MJup, 0.89 ± 0.21 RJup; Anderson et al. 2010), and the space-borne CoRoT mission discovered CoRoT-15 b (63.3± 1.1 MJup, 1.12+0.3 −0.15 RJup; Bouchy et al. 2010). Both of these com- panions have masses comparable to LHS 6343 C, yet or- bit single, F-type stars that are much more massive than LHS 6343 A. The current sample of known transit- ing brown dwarfs are shown in Fig. 8, along with two very low-mass eclipsing M dwarfs. Also shown are the mass and radius predictions of the Baraffe et al. (1998) interior models for three different ages, illustrating that the radius of LHS 6343 C is consistent with the model predictions for a brown dwarf with an age > 1 Gyr yet < 5 Gyr. Our knowledge of the physical characteristics of brown dwarfs is starting to expand owing to the growing pro- ductivity and efficiency of wide-field transit surveys, both from the ground and in space. LHS 6343 is one of the closest stars in the Kepler field and points the way toward additional brown dwarf discoveries in the near future. 10 We thank Keivan Stassun and Leslie Hebb for provid- ing us with their B and V photometry of the LHS 6343 system. We thank Jonathan Irwin for his independent analysis of the Kepler light curve and for encouraging us to incorporate the third-light correction into our forward- modeling procedure; Josh Carter for independently con- firming our best-fitting light curve parameters; and John Gizis for pointing out an error in our calculation of the brown dwarf mass in an earlier draft of this manuscript. We gratefully acknowledge the tireless dedication and hard work of the Kepler team, without whom this project would not be possible. In particular, we thank Jon Jenk- ins and Lucianne Walkowicz for confirming the planet- like nature of the transits following the initial identifica- tion of transit events by K.A. We gratefully acknowl- edge the efforts and dedication of the Keck Observa- tory staff, especially Grant Hill, Scott Dahm and Hien Tran for their support of HIRES and Greg Wirth for support of remote observing. We made use of the SIM- BAD database operated at CDS, Strasbourge, France, and NASA's Astrophysics Data System Bibliographic Services. A. W. H. gratefully acknowledges support from a Townes Post-doctoral Fellowship at the U. C. Berke- ley Space Sciences Laboratory. G. W. M. acknowledges NASA grant NNX06AH52G. Finally, we extend special thanks to those of Hawaiian ancestry on whose sacred mountain of Mauna Kea we are privileged to be guests. Without their generous hospitality, the Keck observa- tions presented herein would not have been possible. REFERENCES Allard, F., et al. 2001, ApJ, 556, 357 Anderson, D. R., et al. 2010, ArXiv e-prints Baraffe, I., et al. 1998, A&A, 337, 403 Basri, G. 2006, Astronomische Nachrichten, 327, 3 Basri, G., Marcy, G. W., & Graham, J. R. 1996, ApJ, 458, 600 Batalha, N. M., et al. 2010, ApJ, 713, L109 Borucki, W. J., et al. 2010, Science, 327, 977 Boss, A. P. 2006, ApJ, 643, 501 Bouchy, F., et al. 2010, arxiv:1010.0179 Bowler, B. P., Liu, M. C., & Cushing, M. C. 2009, ApJ, 706, 1114 Burgasser, A. J., et al. 1999, ApJ, 522, L65 Burgasser, A. J., et al. 2007, Protostars and Planets V, 427 Burrows, A., et al. 1997, ApJ, 491, 856 Carter, J. A. & Winn, J. N. 2009, ApJ, 704, 51 Charbonneau, D., et al. 2005, ApJ, 626, 523 Claret, A. 2004, A&A, 428, 1001 Cushing, M. C. & Vacca, W. D. 2006, AJ, 131, 1797 Deleuil, M., et al. 2008, A&A, 491, 889 Delfosse, X., et al. 2000, A&A, 364, 217 Delorme, P., et al. 2008, A&A, 484, 469 Droege, T. F., et al. 2006, PASP, 118, 1666 Eastman, J., Siverd, R., & Gaudi, B. S. 2010, PASP, 122, 935 Fischer, D. A. & Valenti, J. 2005, ApJ, 622, 1102 Ford, E. B. 2005, AJ, 129, 1706 Geman, S. & Geman, D. 1984, IEEE Transactions on Pattern Analysis and Machine Intelligence, PAMI-6, 721 Girardi, L., et al. 2002, A&A, 391, 195 Grether, D. & Lineweaver, C. H. 2006, ApJ, 640, 1051 Hayward, T. L., et al. 2001, PASP, 113, 105 Howard, A. W., et al. 2010, ApJ, 721, 1467 Irwin, J., et al. 2010, ApJ, 718, 1353 Jenkins, J. M., et al. 2010a, ApJ, 724, 1108 Jenkins, J. M., et al. 2010b, ApJ, 713, L87 Jenkins, J. M., et al. 2010c, ApJ, 713, L120 Johnson, J. A. 2009, PASP, 121, 309 Johnson, J. A., et al. 2010, PASP, 122, 905 Johnson, J. A. & Apps, K. 2009, ApJ, 699, 933 Johnson, J. A., et al. 2004, AJ, 128, 1265 Johnson, J. A., et al. 2008, ApJ, 686, 649 Kipping, D. M. 2010, MNRAS, 408, 1758 Koch, D. G., et al. 2010, ApJ, 713, L79 Kratter, K. M., Murray-Clay, R. A., & Youdin, A. N. 2010, ApJ, Laughlin, G., Bodenheimer, P., & Adams, F. C. 2004, ApJ, 612, 710, 1375 L73 Lawrence, A., et al. 2007, MNRAS, 379, 1599 Liu, M. C., Dupuy, T. J., & Ireland, M. J. 2008, ApJ, 689, 436 Liu, M. C. & Leggett, S. K. 2005, ApJ, 634, 616 L´opez-Morales, M. 2007, ApJ, 660, 732 Mandel, K. & Agol, E. 2002, ApJ, 580, L171 Maness, H. L., et al. 2007, PASP, 119, 90 Marcy, G. W. & Butler, R. P. 2000, PASP, 112, 137 Martin, E. L., et al. 1997, A&A, 327, L29 McCarthy, C. & Zuckerman, B. 2004, AJ, 127, 2871 Oppenheimer, B. R., et al. 1995, Science, 270, 1478 Reid, I. N., et al. 2004, AJ, 128, 463 Ribas, I. 2006, Ap&SS, 304, 89 Schlaufman, K. C. & Laughlin, G. 2010, A&A, 519, A105+ Seager, S. & Mall´en-Ornelas, G. 2003, ApJ, 585, 1038 Sing, D. K. 2010, A&A, 510, A21+ Sozzetti, A., et al. 2007, ApJ, 664, 1190 Stassun, K. G., Mathieu, R. D., & Valenti, J. A. 2006, Nature, 440, 311 Torres, G. 2007, ApJ, 654, 1095 Torres, G., Winn, J. N., & Holman, M. J. 2008, ApJ, 677, 1324 Troy, M., et al. 2000, 4007, 31 Winn, J. N. 2008, ASPC, 398, 101 Winn, J. N., et al. 2009, ApJ, 693, 794 Wright, J. T. & Howard, A. W. 2009, ApJS, 182, 205 Wright, J. T., et al. 2004, ApJS, 152, 261 11 TABLE 5 System Parameters of LHS 6343 Parameter Value 68.3% Confidence Comment Transit Parameters Orbital Period, P [days] Radius Ratio, (RC /R⋆,A)corr Transit Depth, (RC /R⋆,A)2 Scaled semimajor axis, aR ≡ a/R⋆ Orbit inclination, i [deg] Transit impact parameter, b corr Other Orbital Parameters Semimajor axis between Star A and C [AU] Eccentricity Argument of Periastron ω [degrees] Velocity semiamplitude KA [km s−1] Systemic Radial Velocity γ [km s−1] Stellar Parameters MA [M⊙] MB [M⊙] R⋆,A [R⊙] ρA [ρ⊙] log gA [cgs] [Fe/H] Distance [pc] VA [mag] VB [mag] (B − V )A (B − V )B ∆KP [mag] ∆Z [mag] Brown Dwarf Parameters MC [MJ up] RC [RJ up] Mean density, ρC [g cm−3] log gC [cgs] Interval 12.71382 ±0.00004 0.226 0.051 45.3 89.50 0.40 0.0804 0.056 -23 9.6 -46.0 0.370 0.30 0.378 6.6 4.851 0.04 36.6 13.88 14.63 1.57 1.60 0.74 0.5 62.9 0.833 109 5.35 ±0.003 ±0.001 ±0.6 ±0.05 ±0.04 ±0.0006 ±0.032 ±56 ±0.3 ±0.2 ±0.009 ±0.01 ±0.008 ±0.4 ±0.008 ±0.08 ±1.1 ±0.03 ±0.06 ±0.07 ±0.07 ±0.10 ±0.1 ±2.3 ±0.021 ±8 ±0.02 A A A A A A B C C C C D D D A B D D E E F,G F,G F F,G B,C B B,C A Note. -- Note. -- (A) Determined from the parametric fit to the Kepler light curve. (B) Based on group A parameters supplemented by the photometric stellar mass determination described in § 5. (C) Based on our analysis of the Keck/HIRES RV measurements. (D) Based on our photometric mass and radius determinations described in § 5. (E) Eqn. 5. (F) Interpolation of Padova model grids. (G) (B −V ) colors from the Padova models have been corrected by adding empirically measured offset of 0.21 ± 0.07 mag. Gunn-Z magnitudes were approximated using SDSS z′-band, with an additive correction of 0.07 mag based on the difference between the model ∆J compared to the measured ∆J in Table 1.
1908.08585
2
1908
2019-12-12T00:43:02
An 11 Earth-Mass, Long-Period Sub-Neptune Orbiting a Sun-like Star
[ "astro-ph.EP" ]
Although several thousands of exoplanets have now been detected and characterized, observational biases have led to a paucity of long-period, low-mass exoplanets with measured masses and a corresponding lag in our understanding of such planets. In this paper we report the mass estimation and characterization of the long-period exoplanet Kepler-538b. This planet orbits a Sun-like star (V = 11.27) with M_* = 0.892 +/- (0.051, 0.035) M_sun and R_* = 0.8717 +/- (0.0064, 0.0061) R_sun. Kepler-538b is a 2.215 +/- (0.040, 0.034) R_earth sub-Neptune with a period of P = 81.73778 +/- 0.00013 d. It is the only known planet in the system. We collected radial velocity (RV) observations with HIRES on Keck I and HARPS-N on the TNG. We characterized stellar activity by a Gaussian process with a quasi-periodic kernel applied to our RV and cross correlation function full width at half maximum (FWHM) observations. By simultaneously modeling Kepler photometry, RV, and FWHM observations, we found a semi-amplitude of K = 1.68 +/- (0.39, 0.38) m s^-1 and a planet mass of M_p = 10.6 +/- (2.5, 2.4) M_earth. Kepler-538b is the smallest planet beyond P = 50 d with an RV mass measurement. The planet likely consists of a significant fraction of ices (dominated by water ice), in addition to rocks/metals, and a small amount of gas. Sophisticated modeling techniques such as those used in this paper, combined with future spectrographs with ultra high-precision and stability will be vital for yielding more mass measurements in this poorly understood exoplanet regime. This in turn will improve our understanding of the relationship between planet composition and insolation flux and how the rocky to gaseous transition depends on planetary equilibrium temperature.
astro-ph.EP
astro-ph
Preprint typeset using LATEX style emulateapj v. 12/16/11 AN 11 EARTH-MASS, LONG-PERIOD SUB-NEPTUNE ORBITING A SUN-LIKE STAR Andrew W. Mayo1,2,3,†,‡,(cid:63), Vinesh M. Rajpaul4, Lars A. Buchhave2,3, Courtney D. Dressing1, Annelies Mortier4,5, Li Zeng6,7, Charles D. Fortenbach8, Suzanne Aigrain9, Aldo S. Bonomo10, Andrew Collier Cameron5, David Charbonneau7, Adrien Coffinet11, Rosario Cosentino12,13, Mario Damasso10, Xavier Dumusque11, A. F. Martinez Fiorenzano12, Raphaelle D. Haywood7,2, David W. Latham7, Mercedes L´opez-Morales7, Luca Malavolta13, Giusi Micela14, Emilio Molinari15, Logan Pearce16, Francesco Pepe11, David Phillips7, Giampaolo Piotto17,18, Ennio Poretti12,19, Ken Rice20,21, Alessandro Sozzetti10, Stephane Udry11 9 1 0 2 c e D 2 1 . ] P E h p - o r t s a [ 2 v 5 8 5 8 0 . 8 0 9 1 : v i X r a ABSTRACT −0.035M(cid:12) and R∗ = 0.8717+0.0064 −0.0061R(cid:12). Kepler-538b is a 2.215+0.040 Although several thousands of exoplanets have now been detected and characterized, observational biases have led to a paucity of long-period, low-mass exoplanets with measured masses and a cor- responding lag in our understanding of such planets. In this paper we report the mass estimation and characterization of the long-period exoplanet Kepler-538b. This planet orbits a Sun-like star (V = 11.27) with M∗ = 0.892+0.051 −0.034R⊕ sub-Neptune with a period of P = 81.73778 ± 0.00013 days. It is the only known planet in the sys- tem. We collected radial velocity (RV) observations with the High Resolution Echelle Spectrometer (HIRES) on Keck I and High Accuracy Radial velocity Planet Searcher in North hemisphere (HARPS- N) on the Telescopio Nazionale Galileo (TNG). We characterized stellar activity by a Gaussian process with a quasi-periodic kernel applied to our RV and cross-correlation FWHM observations. By simul- taneously modeling Kepler photometry, RV, and FWHM observations, we found a semi-amplitude of K = 1.68+0.39−0.38 m s−1 and a planet mass of Mp = 10.6+2.5−2.4M⊕. Kepler-538b is the smallest planet beyond P = 50 days with an RV mass measurement. The planet likely consists of a significant fraction of ices (dominated by water ice), in addition to rocks/metals, and a small amount of gas. Sophisti- cated modeling techniques such as those used in this paper, combined with future spectrographs with ultra high-precision and stability will be vital for yielding more mass measurements in this poorly un- derstood exoplanet regime. This in turn will improve our understanding of the relationship between planet composition and insolation flux and how the rocky to gaseous transition depends on planetary equilibrium temperature. Keywords: planets and satellites: composition - planets and satellites: detection - planets and satellites: fundamental parameters - planets and satellites: gaseous planets - methods: data analysis - techniques: photometric - techniques: radial velocities † [email protected] ‡ National Science Foundation Graduate Research Fellow (cid:63) Fulbright Fellow 2 NASA Sagan Fellow 1 Astronomy Department, University of California, Berkeley, CA 94720, USA 2 DTU Space, National Space Institute, Technical University of Denmark, Elektrovej 327, DK-2800 Lyngby, Denmark 3 Centre for Star and Planet Formation, Natural History Mu- seum of Denmark & Niels Bohr Institute, University of Copen- hagen, Øster Voldgade 5-7, DK-1350 Copenhagen K., Denmark 4 Astrophysics Group, Cavendish Laboratory, University of Cambridge, J. J. Thomson Avenue, Cambridge CB3 0HE, UK 5 Centre for Exoplanet Science, SUPA, School of Physics and Astronomy, University of St Andrews, St Andrews, KY16 9SS, UK 6 Department of Earth & Planetary Sciences, Harvard Univer- 7 Center for Astrophysics Harvard & Smithsonian, 60 Garden sity, 20 Oxford Street, Cambridge, MA 02138, USA Street, Cambridge, MA 02138, USA 8 Department of Physics and Astronomy, San Francisco State University, San Francisco, CA 94132, USA 9 Department of Physics, Denys Wilkinson Building Keble Road, Oxford, OX1 3RH, UK 10 INAF - Osservatorio Astrofisico di Torino, via Osservatorio 20, I-10025 Pino Torinese, Italy 11 Observatoire de Geneve, Universit`e de Gen´eve, 51 ch. des Maillettes, CH-1290 Sauverny, Switzerland 12 INAF - Fundaci´on Galileo Galilei, Rambla Jos´e Ana Fernan- dez P´erez 7, E-38712 Brena Baja, Spain 13 INAF - Osservatorio Astrofisico di Catania, Via S. Sofia 78, I-95123 Catania, Italy 14 INAF - Osservatorio Astronomico di Palermo, Piazza del Par- 1. INTRODUCTION To date, nearly four thousand exoplanets have been discovered, but over three quarters of them orbit their host star with periods of less than 50 days (NASA Ex- oplanet Archive26; accessed 2019 April 13). However, this is the result of observational biases rather than a feature of the underlying exoplanet population. Bias to short periods is especially strong for the transit method, the most common method of exoplanet detection. Nev- ertheless, Petigura et al. (2018) finds that from 1 to 24 R⊕, the planet occurrence rate either increases or plateaus as a function of period out to many hundreds lamento 1, I-90134 Palermo, Italy 15 INAF - Osservatorio Astronomico di Cagliari, via della Scienza 5, I-09047, Selargius, Italy 16 Department of Astronomy, The University of Texas at Austin, Austin, TX 78712, USA 17 Dipartimento di Fisica e Astronomia "Galileo Galilei," Uni- versita di Padova, Vicolo dell Osservatorio 3, I-35122 Padova, Italy 18 INAF - Osservatorio Astronomico di Padova, Vicolo dellOsser- vatorio 5, I-35122 Padova, Italy 19 INAF - Osservatorio Astronomico di Brera, Via E. Bianchi 46, I-23807 Merate, Italy 20 SUPA, Institute for Astronomy, University of Edinburgh, Royal Observatory, Blackford Hill, Edinburgh EH93HJ, UK 21 Centre for Exoplanet Science, University of Edinburgh, Edin- burgh, EH93FD, UK 26 https://exoplanetarchive.ipac.caltech.edu/ 2 Mayo et al. of days. Therefore, despite the estimated abundance of long-period planets (i.e., planets with periods longer than 50 days27), our understanding of them is still very in- complete. Relative to the short-period population, there are very few long-period exoplanets (particularly in the low-mass regime) with precise and accurate densities and compositions, and even fewer with atmospheric charac- terization. Thus, a larger sample of masses for long-period planets would allow us to address a number of interesting ques- tions. For example, it would allow us to study the rocky to gaseous planet transition and how it depends on stel- lar flux. We could also investigate planet compositions in or near the habitable zone of Sun-like stars. Another interesting feature to study would be the planet radius occurrence gap detected by Fulton et al. (2017) and Fulton & Petigura (2018). Owen & Wu (2017) and Van Eylen et al. (2018) have proposed that photo- evaporation strips planets near their host stars down to the core, thus creating the gap. Lopez & Rice (2018) have investigated the period dependence of the gap posi- tion and Zeng et al. (2017) have analyzed the relationship between gap position and stellar type. More long-period planets, with or without planet masses, would provide new insights into the nature and cause of this radius oc- currence gap. In this paper, we characterize the long-period exo- planet Kepler-538b, the only known planet in the Kepler- 538 system, first validated by Morton et al. (2016). There is a possible second transiting planet candidate with a pe- riod of 117.76 days, but its existence is very much in ques- tion; we briefly discuss this candidate in Section 5.3. We determine the properties of the host star, a G-type star slightly smaller than the Sun. We also determine proper- ties of the exoplanet including the orbital period, mass, radius, and density by modeling transit photometry, ra- dial velocity (RV) data, and stellar activity indices. We find that Kepler-538b is the smallest long-period planet to date with both a measured radius and RV mass. The format of this paper is as follows. In Section 2, we detail our photometric and spectroscopic observations of the planet and its host star. We then discuss stellar pa- rameterization in Section 3 and modeling of photometry and spectroscopy in Section 4. Our results are then pre- sented and discussed in Section 5. Finally, we summarize and conclude our paper in Section 6. 2. OBSERVATIONS Photometric observations of the Kepler-538 system were collected with the Kepler spacecraft (Borucki et al. 2008) across 17 quarters beginning in 2009 May and end- ing in 2013 May. Kepler collected both long-cadence and short-cadence observations of this system. Short- cadence observations (in quarters 3, 7-12 and 17) were collected every 58.89 s, and long cadence observations (in all other quarters) were collected every 1765.5 s (∼ 29.4 minutes). In particular, we used pre-search data con- ditioning (PDC) light curves from these quarters down- loaded from the Mikulski Archive for Space Telescopes. 27 We define long-period planets as exoplanets with periods greater than 50 days. This may seem short relative to planets in our own solar system or many of the multi-year period exoplan- ets already found, but we think it is appropriate, given the relative scarcity of such planets in the known, low-mass planet population. Although Kepler-538 was not validated until Morton et al. (2016), it was flagged as a Kepler Object of Interest well before that. As a result, we have conducted a great deal of spectroscopic follow-up on Kepler-538 since it was identified as a candidate host star by the Kepler mission. First, we collected two spectra with the Tillinghast Re- flector Echelle Spectrograph (TRES; Fur´esz 2008), an R = 44, 000 spectrograph on the 1.5 m Tillinghast reflec- tor at the Fred Lawrence Whipple Observatory (located on Mt. Hopkins, Arizona). These spectra were collected on the nights of 2010 May 28 and 2010 July 5 and had exposure times of 12 and 15 minutes respectively. We also downloaded RVs from 26 spectra collected with the HIRES instrument (Vogt et al. 1994) at the Keck I telescope from 2010 July 25 to 2014 July 11. These spectra were originally collected as part of the Ke- pler Follow-up Observing Program. The standard Cali- fornia Planet Search setup was used (Howard et al. 2010) and the C2 decker was utilized to conduct sky subtrac- tion. Exposure times averaged 1800 s. Finally, we gathered 83 spectra with the High Ac- curacy Radial velocity Planet Searcher in North hemi- sphere (HARPS-N) instrument (Cosentino et al. 2012, Cosentino et al. 2014) on the 3.6 m Telescopio Nazionale Galileo (TNG) on La Palma. These observations were made from 2014 June 20 to 2015 November 7, all with exposure times of 30 minutes. They were collected as part of the HARPS-N Collaboration's Guaranteed Time Observations (GTO) program. Using the technique de- scribed in Malavolta et al. (2017), we confirmed that none of these spectra suffered from Moon contamination. 3. STELLAR CHARACTERIZATION Stellar atmospheric parameters (effective temperature, metallicity, and surface gravity) were determined in two different ways. First, we combined the two TRES spec- tra and used the Stellar Parameter Classification tool (SPC; Buchhave et al. 2012). SPC compares an input spectrum against a library grid of synthetic spectra from Kurucz (1992), interpolating over the library to find the best match as well as uncertainties on the relevant stel- lar parameters. This method provides a measure for the rotational velocity as well. Second, we used ARES+MOOG on the combination of our 83 HARPS-N spectra. More details about this method, based on equivalent widths (EWs), are found in Sousa (2014) and references therein. In short, ARESv2 (Sousa et al. 2015) automatically calculates the EWs of a set of neutral and ionised iron lines (Sousa et al. 2011). These are then used as input in MOOG28 (Sne- den 1973), assuming local thermodynamic equilibrium and using a grid of ATLAS plane-parallel model atmo- spheres (Kurucz 1993). Following Sousa et al. (2011), we added systematic errors in quadrature to our errors. The value for surface gravity was corrected for accuracy following Mortier et al. (2014). The results from SPC and ARES+MOOG agreed well within uncertainties. We then estimated stellar mass, radius, and thus den- sity with the isochrones package, a Python routine for inferring model-based stellar properties from known ob- servations (Morton 2015). We supplied the spectroscopic 28 2017 version: http://www.as.utexas.edu/$\sim$chris/ moog.html Kepler-538b 3 Table 1 Stellar parameters of Kepler-538 Parameter Stellar parameters Unit SPC ARES+MOOG Combined Effective temperature Teff Surface gravity log g Metallicity [m/H] Metallicity [Fe/H] Radius R∗ Mass M∗ Density ρ∗ Distance Age Projected rotational velocity v sin i K g cm−2 dex dex R(cid:12) M(cid:12) ρ(cid:12) pc Gyr km s−1 5547 ± 50 4.51 ± 0.10 −0.03 ± 0.08 ... 0.8707+0.0063 −0.0060 0.925+0.034 −0.036 1.404+0.061 −0.068 156.67+0.71−0.70 3.8+2.1−2.0 1.1 ± 0.5 ... 5522 ± 72 4.55 ± 0.12 −0.15 ± 0.05 0.8727+0.0063 −0.0062 0.870 ± 0.024 1.31 ± 0.052 156.65+0.70−0.68 6.7+1.8−1.6 ... ... ... ... ... 0.8717+0.0064 −0.0061 0.892+0.051 −0.035 1.349+0.089 −0.0716 156.66+0.71−0.69 5.3+2.4−3.0 ... effective temperature, metallicity, the Gaia DR2 paral- lax (Gaia Collaboration et al. 2016, 2018), and multiple photometric magnitudes (B, V, J, H, K, W1, W2, W3, and G) as input. Note that we did not use the surface gravity as an input parameter as this parameter is not well determined spectroscopically (e.g. Mortier et al. 2014). We ran isochrones four times, using the two different sets of spectroscopic parameters and two sets of isochrones, Modules for Experiments in Stellar Astro- physics (MESA) Isochrones and Stellar Tracks (MIST) and Dartmouth29. All four results were consistent, so we followed Mala- volta et al. (2018) and derived our final set of param- eters and uncertainties from the 16th, 50th, and 84th percentile values of the combined posteriors, minimiz- ing systematic biases from using different spectroscopic methods or isochrones. The results of this analysis are listed in Table 1. As a useful check, we find that our estimates of stellar effective temperature, stellar radius, and distance are all within 1σ of the Gaia DR2 revised Kepler stellar param- eters (Berger et al. 2018). 3.1. Consistency with Stellar Activity and Gyrochronology like stellar activity, We used this log R(cid:48) As will be discussed in more detail in later sections, RV observations with both HIRES and HARPS-N yielded log R(cid:48) HK, an indicator of stellar activity. Although log R(cid:48) HK, is time variable, taking an average or median over time is still a useful metric of the general activity level of the star. The median HK with HIRES and HARPS-N was −4.946± 0.035 log R(cid:48) and −5.001 ± 0.027, respectively. The overall log R(cid:48) across both data sets was −4.990 ± 0.034. HK value and the B − V color in- dex30 to estimate the stellar rotation period via Noyes et al. (1984), finding a value of 32.0 ± 1.0 days. Our full model (described in Sections 4.1 and 4.2.3) included the rotation period as a free parameter, which we estimated to be 25.2+6.5−1.2 days, in agreement with the stellar activity predicted rotation period to within 1σ. Further, during our processing of photometric data (see Section 4.1), we produced a periodogram and an auto-correlation func- tion of the photometry. We found signals near 22 and HK 29 The Dartmouth isochrones did not use the G magnitude. 30 determined from https://exofop.ipac.caltech.edu/; accessed 2019 July 29 32 days in the former as well as a weak, broad signal around 20 − 25 days in the latter, all of which are near the activity-inferred rotation period or the rotation pe- riod estimated from our model. We also checked that our estimate of stellar age was consistent with gyrochronology. We found a gyrochrono- logical age for Kepler-538 first by determining the convec- tive turnover timescale from Barnes & Kim (2010) using the B − V color index. Then we used the gyrochrono- logical relation in Barnes (2010) to calculate age from the convective turnover timescale and the rotation period (calculated from our full model). In this way, we deter- mined a stellar age of 3.40+1.86−0.29 Gyr, consistent within 1σ of our isochrone-derived age of 5.3+2.4−3.0 Gyr. 3.2. Possible Binarity of Kepler-538 In order to investigate whether Kepler-538 may be a bi- nary star or have a companion, either of which could have an effect on the dynamics or nature of Kepler-538b, we downloaded all adaptive optics (AO) and speckle data for the star uploaded to https://exofop.ipac.caltech.edu/k2/ before 2019 July 30. The Palomar High Angular Res- olution Observer (PHARO) on the Palomar-5 m tele- scope collected AO observations on 2010 July 1 in J and Ks band; no companions were found between 2" and 5" down to 19th magnitude. The Differential Speckle Sur- vey Instrument (DSSI) on the WIYN-3.5 m telescope collected speckle observations on 23 October 2010 in r and v band; no companions were found between 0".2 and 1".8 down to a contrast of ∆m = 3.6. Finally, the Robo- AO instrument on the Palomar-1.5 m telescope collected an AO observation on 2012 July 28 in the i band; no companions were found between 0".15 and 2".5 down to a contrast of ∆m ≈ 6. In short, there is no evidence of a close stellar companion in any of the AO or speckle data. However, it is worth noting that there is a faint co- moving object 17" from Kepler-538, which Gaia found at approximately the same distance of 157 pc (Gaia Collab- oration et al. 2016, 2018). This means if the two stars are at the same distance, they are separated by 2700±12 au, a large enough separation to negligibly affect the planet. Both objects have good astrometric solutions with Gaia (Lindegren et al. 2018), and their relative motion given by Gaia proper motions is 0.408 ± 0.510 km s−1. How- ever, this relative motion is so slight that we were unable to meaningfully constrain orbital motion. We estimated the mass of the comoving object to be 4 Mayo et al. the NASA Exoplanet Archive31 (accessed 2019 Febru- ary 16) was subtracted from the light curve so that in- transit data would not be clipped or flattened out in the next steps. Next, we flattened the light curve using the lightkurve flatten function, which uses a Savitzky- Golay filter. A window length of 615 or 41 was selected (i.e., 615 or 41 consecutive data points) for short-cadence and long-cadence data respectively, which is approxi- mately three times the ratio between the transit duration and the observation cadence. Then, we clipped outlier data points discrepant from the median flux by more than 5σ. Lastly, we added the transit model from the earlier step back to the light curve. The reduced data can be seen in Fig. 1, plotted in time and also phase- folded to the period of Kepler-538b. We modeled the light curve with the BATMAN Python package (Kreidberg 2015), which is based on the Mandel & Agol (2002) transit model. The model included a base- line offset parameter, a white noise parameter (to allow for instrumental and systematic noise in the data), two quadratic limb-darkening parameters (using the Kipping 2013a parameterization), the transit time (i.e., reference epoch), orbital period, planet radius relative to stellar radius, transit duration, impact parameter, eccentricity, and longitude of periastron. We assumed uniform, Jeffreys, or modified Jeffreys pri- ors for most of the parameters in this model, which are listed in Table 2. A Jeffreys prior is less informative than a uniform prior when the prior range is large and the scale of the parameter is unknown. A modified Jeffreys prior has the following form (Gregory 2007): p(X) = 1 1 X + X0 ln( Xmax+X0 Xmin+X0 ) where Xmin and Xmax are the minimum and maximum prior value and X0 is the location of a knee in the prior. A modified Jeffreys prior behaves like a Jeffreys prior above the knee at X0 and behaves likes a uniform prior below the knee; this is useful when the prior includes zero (creating an asymptote for a conventional Jeffreys prior). A Jeffreys prior is simply a modified Jeffreys prior with the knee at X0 = 0. The only parameter with a different prior was orbital eccentricity. We applied a beta prior to orbital eccentric- ity using the values recommended by Kipping (2013b); we also truncated the prior to exclude e > 0.95. Additionally, we also applied a stellar density prior. This was done given the fact that stellar density can be measured in two distinct ways: from photometry for a transiting exoplanet and from a stellar spectrum com- bined with stellar evolutionary tracks (we used the latter method in Section 3). Specifically, stellar density can be calculated via the following equation (Seager & Mall´en- Ornelas 2003; Sozzetti et al. 2007): (cid:18) a (cid:19)3 R∗ ρ∗ = 3π GP 2 (1) where the orbital period (P ) and the normalized semi- major axis (a/R∗) are exoplanet properties that can be derived from the light curve. We applied a Gaussian 31 https://exoplanetarchive.ipac.caltech.edu/ Figure 1. Transit plot of Kepler-538b. The top subplot is the pre-search data conditioning (PDC) Kepler photometry. The top panel of the bottom subplot shows the phase-folded photometry in and near the transit of Kepler-538b, with the best-fit transit model in orange and binned data in blue. The bottom panel of the bottom subplot shows the photometric residuals after subtracting the best-fit transit model. 0.1169± 0.0075M(cid:12) by applying the photometric relation in Mann et al. (2019) to the Two Micron All-Sky Survey (2MASS) Ks magnitude (Cutri et al. 2003), which gives a total system mass of 1.009± 0.044M(cid:12). With this mass and separation, a circular face-on orbit would have a to- tal relative velocity of 0.576 ± 0.013 km s−1. Thus, both the velocity of a face-on circular orbit and zero velocity are within 1σ of the measured relative velocity. With such weak constraints from Gaia DR2, we cannot rule out a circular orbit at wide separation nor a highly eccen- tric orbit, currently observed at apastron, which brings the companion close enough in to potentially affect the planet. 4. DATA ANALYSIS Our analysis of photometric and spectroscopic data in- cluded a simultaneous fit to both data types. Therefore, we first describe the data reduction process and model components of photometry and spectroscopy separately, then discuss the combined model afterward. 4.1. Photometric Data We cleaned and reduced the photometric Kepler data using the lightkurve Python package (Barentsen et al. 2019). Each quarter was cleaned and reduced separately. For a given quarter, observation times without a corre- sponding flux were removed. Then, a crude light curve model based on the exoplanet parameters reported in Kepler-538b 5 Transit and RV parameters of Kepler-538b Table 2 Unit This Paper Priors Parameter Transit parameters Period P Time of first transit Orbital eccentricity e Longitude of periastron ω Impact parameter b Transit duration t14 Radius ratio Rp/R∗ Quadratic limb-darkening parameter q1 Quadratic limb-darkening parameter q2 Normalized baseline offset Photometric white noise amplitude RV parameters Semi-amplitude K HIRES RV white noise amplitude HARPS-N RV white noise amplitude HARPS-N FWHM white noise amplitude HIRES RV offset amplitude HARPS-N RV offset amplitude HARPS-N FWHM offset amplitude GP RV convective blueshift amplitude Vc GP RV rotation modulation amplitude Vr GP FWHM amplitude Fc GP stellar rotation period P∗ GP inverse harmonic complexity λp GP evolution time-scale λe Derived parameters R⊕ Planet radius Rp System scale a/R∗ ... au Planet semi-major axis a degree Orbital inclination i M⊕ Planet mass Mp ρ⊕ Planet mean density ρp g cm−3 Planet mean density ρp S⊕ Planet insolation flux Sp Planet equilibrium temperature Teq (albedo = 0.3) K Planet equilibrium temperature Teq (albedo = 0.5) K a95% confidence limit bBeta distribution parameter values from Kipping (2013b). cPrior also truncated to exclude e > 0.95. dRotation period uncertainties are highly asymmetric because the posterior includes a large peak at 25 days and a smaller peak at 31 days. prior to the exoplanet-derived stellar density using the density (and corresponding uncertainties) derived from spectra and stellar evolutionary tracks. 4.2. RV Data Our RV analysis of Kepler-538b included not only the RV values determined from our HIRES and HARPS-N spectra, but also a number of indicators of stellar activ- ity estimated from these spectra. For HARPS-N, these included the cross-correlation function (CCF) bisector span inverse slope (hereafter BIS), the CCF full width at half maximum (FWHM), and log R(cid:48) HK. Our data re- day BJD-2454833 ... degree ... hr ... ... ... ppm ppm m s−1 m s−1 m s−1 m s−1 m s−1 m s−1 m s−1 m s−1 m s−1 m s−1 day ... day 81.73778 ± 0.00013 211.6789+0.0010 −0.0011 0.041+0.034 −0.029 (< 0.11)a 140+140−90 0.41+0.10−0.21 6.62+0.21−0.13 0.02329+0.00039 −0.00033 0.164+0.067 −0.042 0.74+0.16−0.22 −2.1+2.7−2.8 112.2+2.5−2.4 −37322.07+0.58−0.73 1.69+0.39−0.38 3.25+0.56−0.48 2.24+0.29−0.27 6.71+0.52−0.46 −0.50+0.78−0.87 6655.4+7.5−8.6 0.86+0.75−0.54 4.0+5.7−3.0 13.3+5.9−4.9 25.2+6.5−1.2 5.2+2.8−2.5 370+200−140 d 2.215+0.040 −0.034 87.5+1.5−1.6 0.3548+0.0066 −0.0068 89.73+0.14−0.06 10.6+2.5−2.4 0.98 ± 0.23 5.4 ± 1.3 5.19+0.31−0.28 380 350 Unif(81.73666,81.73896) Unif(211.6671,211.6901) Beta(0.867,3.03)bc Unif(0,360) Unif(0,1) Unif(0,24) Jeffreys(0.001,1) Unif(0,1) Unif(0,1) Unif(-100,100) ModJeffreys(1,1000,234) ModJeffreys(0.01,10,2.1) ModJeffreys(0,10,2.1) ModJeffreys(0,10,2.1) Jeffreys(0.01,10) Unif(-5,5) Unif(-37330,-37315) Unif(6600,6700) ModJeffreys(0,15,2.1) ModJeffreys(0,15,2.1) Jeffreys(0.01,25) Unif(20,40) Unif(0.25,10) Jeffreys(1,1000) ... ... ... ... ... ... ... ... ... ... duction was performed with the data reduction software (DRS) 3.7 HARPS-N pipeline which applied a G2 stel- lar type mask. For HIRES, RVs are estimated with an iodine cell rather than cross correlation, so log R(cid:48) HK was calculated but not BIS or FWHM. The RV and FWHM observations (and the correspond- ing model fit) can be seen in Fig. 2. Additionally, all RV, FWHM, BIS, and log R(cid:48) HK values are listed in Table 3. There is a clear long-term trend in the FWHM obser- vations (and to a lesser extent in the BIS and log R(cid:48) HK observations). However, we could not determine whether these trends have a stellar or instrumental origin, nor 6 Mayo et al. why there are no similar trends in the RV observations. On the one hand, when we checked three standard stars observed by HARPS-N during the same period of time, only one showed a similar FWHM trend. On the other hand, a FWHM trend in HARPS-N observations was also reported by Benatti et al. (2017) due to a defocus- ing problem, but that issue was corrected in 2014 March, before our first HARPS-N observations began. Still, per- haps a similar but slower and smaller drift affected our observations. We first analyzed our observations with a periodogram, then with a correlation plot, and then constructed a model for our spectroscopic data. 4.2.1. Periodogram Analysis Before modeling our spectroscopic observations, we first investigated the frequency structure of our data. We made a generalized Lomb-Scargle periodogram (Scar- gle 1982; Zechmeister & Kurster 2009) of log R(cid:48) HK, BIS, FWHM, RV, and the window function of the observa- tion time series, all of which can be seen in Fig. 3. log R(cid:48) HK, BIS, and FWHM are indicators of stellar ac- tivity (Queloz et al. 2001; also see Haywood 2015, and references therein). The window function shows how the signals are modified by the time sampling of the mea- surements. The HARPS-N RV periodogram shows a clear peak at 82 days, the orbital period of Kepler-538b; the HIRES RV periodogram shows a weaker signal at the same pe- riod. None of the other periodograms show a similar feature, lending credence to the RV detection of Kepler- 538b. The RV periodograms also exhibit two larger peaks near 0.03− 0.04 days−1, interpreted as the rotational fre- quency. Indeed, as our model fit discussed later in Sec- tion 4.3 and the results in Table 2 will show, both peaks fall within the 1σ confidence region of the stellar rota- tion period. (See Section 4.2.3 for a description of our rotation period estimation.) We also find that the long- term trends observed in the activity indices, combined with the spectral window, affect the periodograms, since a long-term trend is clearly noticeable (see Fig. 2 and Ta- ble 3). We removed these trends and found the resulting periodograms show a peak at the rotational period, but nothing at the orbital period. 4.2.2. Correlation Analysis We also examined correlations between the RV obser- vations and the other stellar activity indices. As can be seen in Fig. 4, there is a slightly stronger correlation between RV and FWHM than between RV and BIS or log R(cid:48) HK. However, there may also be useful information in the correlations between RV and BIS or log R(cid:48) HK. In order to test this, we cross-checked results that included BIS and log R(cid:48) HK in the modeling against those that did not and found consistent results. For this reason, and for the sake of simplicity, in this paper we only report our analysis of RVs in conjunction with FWHM observations. 4.2.3. General RV Modeling Approach In order to model our RV and FWHM observations, we followed the method described in Rajpaul et al. (2015, hereafter R15), which establishes a method to character- ize stellar activity that uses simultaneous regression of distinct data types (with potentially distinct time series). Here we briefly discuss Gaussian process (GP) regression and the novel approach to GPs used by R15. In brief, a GP is a stochastic process that captures the covariance between observations and allows for the mod- eling of correlated noise (Rasmussen & Williams 2006). A GP is specified by a covariance matrix in which the di- agonal elements are the individual observation variances and each off-diagonal element describes the covariance between two observations. The values of the off-diagonal elements are determined by a kernel function, which de- scribes the nature of the correlated noise. GPs provide a great deal of flexibility that has made them an effec- tive tool to account for stellar activity (Haywood et al. 2014). R15 recommended characterizing stellar activity with a quasi-periodic (QP) kernel, which balances phys- ical motivation with simplicity. The QP kernel uses four parameters (commonly called hyperparameters) and de- fines the covariance matrix as follows: KQP(ti, tj) = h2 exp (cid:18) − sin2(cid:0)π(ti − tj)/P∗(cid:1) 2λ2 p (cid:19) , (2) − (ti − tj)2 2λ2 e where ti and tj are observations made at any two times, h is the amplitude hyperparameter (though not a true amplitude, as it incorporates some multiplicative con- stants), P∗ is the period of the variability (i.e., the ro- tation period in the case of stellar activity), λp is the inverse harmonic complexity (a smoothness factor that acts as a proxy for the number of turning points and inflection points per rotation period), and λe is an ex- ponential decay factor (scaling with, though not exactly equal to, the decay timescale of the spots on the star). One of the key insights of R15 is the way in which they related multiple GPs to one another. GP regression can be used on multiple data sets by constructing a co- variance matrix that describes the covariances between two observations of any type. In our case, this means any possible pairing of RV-RV, RV-FWHM, or FWHM- FWHM data points. The following equations (based on equations 13 and 14 from R15) relate RV and FWHM: ∆RV = VcG(t) + Vr G(t) F W HM = FcG(t) (3) (4) Here, G(t) is an underlying GP directly quantifying stellar activity, and Vc, Vr, and Fc are amplitude pa- rameters corresponding to the RV convective blueshift suppression effect, RV rotation modulation, and FWHM signal amplitude (note that this means there are three amplitude parameters instead of the single h parameter expressed in Equation 2). Because RVs and FWHMs re- spond differently to the underlying stellar activity, this approach allows for more rigorous characterization of the stellar activity than methods using only RV observations, which improves the separation of the stellar and plane- tary signals. We followed R15 and simultaneously modeled the HIRES RV data as well as the HARPS-N RV and FWHM data. This included a separate offset parameter and noise parameter (added in quadrature to the uncertain- Kepler-538b 7 Figure 2. Stellar activity and corresponding Gaussian process regression of Kepler-538 (with planetary signal removed). The top subplot shows the HIRES (orange) and HARPS-N (blue) mean-subtracted RV observations and corresponding model fit in the top panel, with residuals in the bottom panel. The black line is the model fit and the gray region is the 1σ confidence interval (drawn from the full posterior distribution). The data points in boxes correspond to the white noise amplitude modeled for each data set. The middle subplot is a zoom in of the top subplot to the latter two campaigns of observations (only the HARPS-N data). The bottom subplot shows the mean-subtracted FWHM times from HARPS-N (matching the time series of the middle panel) and the corresponding model fit in the top panel, residuals in the bottom panel. Note: two RV data points with error bars greater than 5 m s−1 were removed from the plots (but not the underlying model fit). 55005750600062506500675070007250151050510RV (m/s)HIRES RV white noiseHARPS-N RV white noise55005750600062506500675070007250BJD - 245000010010Residuals69007000710072007300151050510RV (m/s)HARPS-N RV white noise69007000710072007300BJD - 2450000505Residuals690070007100720073000.030.020.010.000.01FWHM (km/s)HARPS-N FWHM white noise69007000710072007300BJD - 24500000.010.000.01ResidualsHIRESHARPS-N 8 Mayo et al. ties) for both RV data sets and the FWHM data set (for a total of three offset parameters and three white noise parameters). Finally, the RV reflex motion due to the planet was characterized by a simple five-parameter orbital model: reference epoch, orbital period, reflex mo- tion semi-amplitude, eccentricity, and longitude of peri- astron. Because we conducted a joint fit to both photometry and spectroscopy, all orbital parameters except for reflex motion semi-amplitude are simultaneously used in our photometric model. In other words, reference epoch, or- bital period, eccentricity, and longitude of periastron are used in both the photometric and spectroscopic compo- nents of our full model. For all of the parameters used in the spectroscopic por- tion of the model, we assumed uniform, Jeffreys, or mod- ified Jeffreys priors. The specific types and bounds of the priors are all listed in Table 2. 4.3. Parameter Estimation Overall, our full model included a photometric base- line offset parameter, a photometric white noise param- eter, two quadratic limb-darkening parameters, the im- pact parameter, the transit duration, the planet radius relative to the stellar radius, the reference epoch, the orbital period, eccentricity, longitude of periastron, the reflex motion semi-amplitude, three spectroscopic offset parameters and three spectroscopic white noise param- eters (for HIRES RV, HARPS-N RV, and HARPS-N FWHM), and six GP hyperparameters (two correspond- ing to the two RV semi-amplitudes and one correspond- ing to the FWHM semi-amplitude in Equations 3 and 4, as well as the stellar rotation period, a smoothness factor, and an exponential decay factor). This yielded a total of 24 parameters, all of which are also listed in Table 2. (Note: because we only modeled the detrended and flat- tened photometry, our estimation of the stellar rotation period was derived solely from our spectroscopic data.) We estimated model parameters using MultiNest (Feroz et al. 2009, 2013), a Bayesian inference tool for pa- rameter space exploration, especially well suited for mul- timodal distributions. We used the following MultiNest settings for our parameter estimation: constant efficiency mode, importance nested sampling mode, multimodal mode, sampling efficiency = 0.01, 1000 live points, and evidence tolerance = 0.1. Our full results from this analysis are presented in Ta- ble 2 and discussed in Section 5. Further, the best-fit transit model is plotted against the photometric data in Fig. 1 and the phase-folded, stellar-activity-removed RV observations and model are presented in Fig. 5. We find Kepler-538b to have a mass of Mp = 10.6+2.5−2.4M⊕, a radius of Rp = 2.215+0.040 −0.034R⊕, a mean density of ρp = 0.98 ± 0.23 ρ⊕, and negligible eccentricity (con- sistent with zero, < 0.11 at 95% confidence). Notably, thanks to the Gaia parallax, our uncertainty on the plan- etary radius is less than 2%. For context, the average uncertainty, 0.037R⊕, is only 236 km, approximately the distance between Portland and Seattle32. Finally, we also note that our estimates of transit parameters are all 32 https://www.distancecalculator.net/ from-portland-to-seattle within 1σ of those reported in the original Kepler-538b validation paper (Morton et al. 2016). 4.4. Model Tests In order to confirm the validity of the results from our RV analysis, we conducted a number of tests designed to verify both our method of analysis and its output. These tests included removing our prior knowledge (obtained via transit photometry) of the transit time and period, injecting and recovering synthetic planet signals into the RV data, and removing the GP to model only the planet signal. 4.4.1. Removing the Transit Prior The first test we conducted was to repeat our analysis without any photometric observations, thereby removing the strong photometric constraints on the transit time and orbital period. We refit our model with a prior of BJD-2453833 = Unif(172,252) on transit time, P = Jef- freys(40,120) on orbital period, and the same priors on all other parameters that we previously used in our full anal- ysis. We fit against only RV and FWHM observations, so we did not have any photometric parameters. Our choice of transit time prior was large enough to be naive, but small enough to exclude other transit times modulo some number of orbital periods. Similarly, our choice of orbital period prior was large enough to be naive, but small enough (on the lower end) to prevent overlap with the stellar rotation period of 25 − 30 days. The results were consistent with the full simultaneous fit to spectroscopy and photometry. Of course, the pos- terior distributions on transit time and orbital period were much wider, which is to be expected. Specifically, the transit time was found to be t0 (BJD - 2454833) = 203+14−13 and the period was found to be P = 82.25+0.62−0.74 days. However, all parameters agreed within 1σ of those from the full, simultaneous fit results. Further, all un- certainties (other than those of transit time, period, and eccentricity) were of a similar scale to those from the full model. 4.4.2. Injection Tests The next test we conducted was to introduce a 1.7 m s−1, non-eccentric, sinusoidal planetary signal into the RV data at various periods to see whether the sig- nal could be recovered, whether the measured RV semi- amplitude was accurate, and whether the uncertainties were similar to those for Kepler-538b. We ran four sep- arate model fits with a synthetic planetary signal intro- duced at 60 days, 70 days, 90 days, and 100 days respec- tively. For each data set, we modeled Kepler-538b and the synthetic signal simultaneously, including eccentric- ity in the model for both planets. To reduce computa- tional expenses, we did not model the Kepler photometry for these tests, instead we applied Gaussian priors to the orbital period and transit time of Kepler-538b based on the values from our main results (see Table 2). As for our injected signal, we applied Gaussian priors to transit time and orbital period, centered respectively on the transit time and orbital period of the injected signal, with the same variance on transit time and same fractional vari- ance on orbital period as for Kepler-538b. Finally, priors on semi-amplitude, eccentricity, and longitude of perias- tron were identical to those for Kepler-538b. Kepler-538b 9 Figure 3. Periodograms of the window function (computed from observation times), RV, log R(cid:48) HK , CCF FWHM, and CCF BIS the Kepler-538 system. Subplots in blue are based on HARPS-N observations, subplots in orange are based on HIRES observations, and subplots in pink are based on both HARPS-N and HIRES. The gray region is the 1σ confidence interval of the rotation period of Kepler- 538 (a stellar activity parameter we estimated in our full model). The gray line is the orbital period of Kepler-538b (P = 81.74 days). Lastly, the dashed black lines correspond to various false alarm probabilities.(Note the different y-axis scalings for HIRES.)) 0.050.100.15logR0HK Frequency (d1)0.050.100.15BIS Frequency (d1)0.050.100.15FWHM Frequency (d1)0.050.100.15RV Frequency (d1)0.050.100.15Window Function Frequency (d1)0.10.20.3Power0.20.40.60.20.40.6Power0.10.20.30.10.20.3PowerHIRES + HARPS-NHIRESHARPS-NPlanet orbital periodStellar rotation period (1 region)FAP = 0.1%FAP = 1%FAP = 10% 10 Mayo et al. Figure 5. Kepler-538 RVs (with stellar activity subtracted) as a function of the orbital phase of Kepler-538b. Observations from HARPS-N and HIRES are plotted in blue and orange respectively, and binned data points are plotted in black. Data in the gray regions on each side of the plot are duplicates of the data in the white region. The median model and 1σ confidence interval are plotted as a black line and gray region respectively. Note: two RV data points with error bars greater than 5 m s−1 were removed from the plot (but not the underlying model fit). Another important test we conducted was trying to model the RVs of Kepler-538b without accounting for the stellar activity at all. We did this by simply running the analysis without the GP. If the GP regression adequately accounted for the stellar activity (rather than subsume and weaken the planetary signal), we would expect to recover a similar RV semi-amplitude for the planet when the GP is excluded, as well as either comparable or larger uncertainties. And this is indeed what we find. Without a GP, we found an RV semi-amplitude of K = 2.06+0.49−0.46 m s−1, within 1σ of the semi-amplitude found when a GP was included. Similarly, all other parameters in common be- tween the two model fits agreed to within 1σ, adding confidence to our results. This particular test illustrates that our choice to use a GP to account for stellar activity was sufficient for this system and data set, though not strictly necessary. This may be due to the long evolution time scale of the stel- lar activity and the large difference in periods between stellar rotation and planetary orbital period. However, we cannot rely on favorable stellar features in general, therefore it is best to err on the side of caution and use a sufficiently sophisticated method (e.g. GP regression) to characterize stellar activity signals. 5. RESULTS AND DISCUSSION The results of our stellar characterization and light curve, RV, and FWHM modeling can be found in Ta- bles 1 and 2. After conducting our model fits and running the req- uisite follow-up tests, we found the mass of Kepler-538b to be Mp = 10.6+2.5−2.4M⊕. Combining this with the plan- etary radius of Rp = 2.215+0.040 −0.034R⊕ resulted in a plane- tary density of ρp = 0.98 ± 0.23 ρ⊕, or 5.4 ± 1.3 g cm−3. Owing to its long orbital period, and its location on log R(cid:48) Figure 4. Scatter plots of RV vs. HK , BIS, and FWHM for Kepler-538. The RVs have been mean-subtracted and plotted against the other three data types. Blue data points correspond to HARPS-N observations, orange data points to HIRES. In the top left corner of each panel is the Spearman correlation coefficient between the two data sets, an indicator of nonlinear, monotonic correlation. (The coefficients were calculated using the observation values but not their uncertainties.) In all four model fits, we recovered the semi-amplitude of the injected signal to within 1σ of 1.7 m s−1 (except for the 60d injection test, for which we found a semi- amplitude that was less than 1.7 m s−1 by 1.1σ). Further, the recovered semi-amplitude uncertainty of the injected planets were all on the order of 0.4 − 0.5 m s−1, similar to the error bars on the semi-amplitude of Kepler-538b. Finally, in all four cases, the measured eccentricity of the injected planet was consistent with zero to within 2σ. 4.4.3. Fitting without a GP 5.105.055.004.954.90log(R0HK)rs= 0.180.060.050.040.030.020.01BIS (km/s)rs= 0.191050510RV (m/s)6.646.656.666.676.68FWHM (km/s)rs= 0.290.500.250.000.250.500.751.001.251.50Orbital Phase864202468RV (m/s)HARPS-NHIRES Kepler-538b 11 of them. If Kepler-538b were scattered inward, then its orbital eccentricity could have been higher initially, and then damped to its current value through interactions with the disk when the disk was still around. Alterna- tively, inward migration through planet-disk interactions may be a more likely scenario, since a disk would al- ways keep the planet orbital eccentricity low (Chambers 2018; Morbidelli 2018) and would probably be required to damp any eccentricity from scattering. In summary, Kepler-538b is only the tip of a huge ice- berg, likely representing a class of planets common in our Galaxy, but which are not found in our own solar sys- tem. The absence of planets in between the size of the Earth and Neptune (about four Earth radii) is linked to the formation/presence of a gas giant -- Jupiter (Izidoro et al. 2015; Barbato et al. 2018), and vice versa. To date, very few exoplanets have been found on long- period orbits that also have any kind of mass mea- surements. In fact, according to the NASA Exoplanet Archive33 (accessed 2019 July 31), there are only 10 tran- siting exoplanets (excluding Kepler-538b) with an RV mass measurement and an orbital period greater than 50 days. If we look at other common methods of mass measurement (specifically transit timing variations and dynamical mass measurements of circumbinary planets), that number only increases to 37. Further, most of those planets are quite large, more similar to Jupiter or Saturn in mass and radius than Nep- tune or Earth. Fig. 7 demonstrates where Kepler-538b fits into this sparse region of parameter space. Kepler- 538b is one of the very few small, low-mass planets well characterized to date. As the sample of small, long-period planets with pre- cisely determined masses and densities grows, we will be able to address a number of fundamental questions. For example, what effect does stellar incident flux have on the size and composition of exoplanets? Since most known exoplanets have periods shorter than that of Mer- cury, it is difficult to analyze exoplanet composition and size for incident fluxes comparable to or less than that of Earth. Similarly, is there a relationship between the location or depth of the planet radius occurrence gap de- tected by Fulton et al. (2017) and a planet's mass or com- position? Further characterization of this gap at longer periods would help confirm (or refute) the photoevapora- tion explanation of the gap and therefore provide insights about exoplanet formation. 5.1. Detection of Kepler-538b with Other Methods As methods of detecting exoplanets become more sen- sitive, regions of parameter space accessible to multiple detection methods will grow, and with them the oppor- tunity to more rigorously characterize the planet popu- lation and calibrate detection methods against one an- other. Kepler-538b pushes RV characterization further into the low-mass, long-period planet regime. As a re- sult, it is interesting to explore whether other methods might also be able to characterize such a planet. 33 https://exoplanetarchive.ipac.caltech.edu/. This num- ber was determined by constraining orbital period > 50 days, planet mass < 11MJup, planet mass limit flag = 0 (to remove upper limit results), planet circumbinary flag = 0, planet transit flag = 1, and planet RV flag = 1. Figure 6. Mass-radius diagram of transiting planets with frac- tional mass and radius uncertainties less than 50%. Planet col- ors correspond to orbital period, with short periods in red and long periods (such as Kepler-538b) in blue. Further, except for Kepler-538b, planets with larger fractional mass and radius uncer- tainties are fainter. Venus and Earth are also labeled and plotted in black for reference. Gray lines correspond to planetary compo- sitions (from top to bottom) of 100% H2O, 50% H2O, 25% H2O, 100% MgSiO3, 50% MgSiO3 + 50% Fe, and 100% Fe, respectively (Zeng & Sasselov 2013; Zeng et al. 2016). Kepler-538b lies closest to the 25% H2O composition line. The planet likely consists of a significant fraction of ices (dominated by water ice), in addition to rocks/metals, and a small amount of gas. the mass-radius diagram, Kepler-538b likely consists of a significant fraction of ices (dominated by water ice), in addition to rocks/metals, and a small amount of gas (Zeng et al. 2018). Its host star is slightly less massive than our own Sun. Because the luminosity of a main- sequence star is a strong function of its mass (typically to the power of 3 or 4), the luminosity of the host star Kepler-538 is somewhat less than the Sun. Therefore, the snowline in the disk when this system was formed was closer in, increasing the likelihood for Kepler-538b to accrete ices during its formation. The estimated bulk density of Kepler-538b is compa- rable to that of the Earth. However, this high mean density is partly due to its high mass resulting in more compression of materials under self-gravity. Its uncom- pressed density, as revealed by the mass-radius curves (Zeng & Sasselov 2013; Zeng et al. 2016) in Fig. 6, is consistent with a composition somewhat less dense than pure-rocky and/or Earth-like rocky (1:2 iron/rock mix- ture). One ready explanation is that Kepler-538b is an icy core, which for some reason had not accreted as much gas as our own Uranus or Neptune (both are estimated to have a few up to ten perfect mass of gas). The eccentricity of Kepler-538b is small (less than 0.11 with 95% confidence). However, the planet may still have arisen from a dynamical origin, that is, inward planet migration due to planet-planet gravitational interactions (Raymond et al. 2009). Some planet formation theories have suggested the formation of multiple icy cores in rel- atively adjacent space near the snowline around a host star, increasing the likelihood of dynamical interactions among them and resulting in inward scatterings for some 1234561020Planet Mass (M)0.51.01.52.02.53.0Planet Radius (R)100% H2O50% H2O25% H2O100% MgSiO350% Fe + 50% MgSiO3100% FeEarthVenusKepler-538b10203040506070Period (d) 12 Mayo et al. crolensing sensitivity curve of the Wide Field Infrared Survey Telescope (WFIRST). They estimated that if ev- ery star hosted a planet like Kepler-538b, we could expect WFIRST to detect a microlensing signal from roughly 10-30 such planets during the course of the full mission (see Fig. 9 from Penny et al. 2019). 5.2. Potential for Atmospheric Characterization One interesting question to ask about Kepler-538b is whether or not it may be amenable to atmospheric char- acterization via transmission spectroscopy. The James Web Space Telescope (JWST ; Gardner et al. 2006; Dem- ing et al. 2009; Kalirai 2018) will devote a significant por- tion of its mission to the characterization of exoplanet atmospheres. The spectra shown in Fig. 8 for the at- mosphere of Kepler-538b were generated by the JWST Exoplanet Targeting (JET) code (C. D. Fortenbach & C. D. Dressing 2019, in preparation) assuming five observed transits. This code first takes the observed planet and system parameters (Rp, period, insolation flux, R∗, Tef f , and J-band magnitude) and then derives other key pa- rameters (semi-major axis, Teq, planet surface gravity, planet mass, and transit duration). In this case we used the planet mass already determined in this paper. We also assumed an optimistic low-metallicity (five times so- lar) planetary atmosphere with no clouds. JET then used Exo-Transmit (Kempton et al. 2017) to generate model transmission spectra and used Pandexo (Batalha et al. 2017) to generate simulated instrument spectra. We fo- cused on the Near InfraRed Imager and Slitless Spec- trograph (NIRISS) SOSS-Or1 and NIRSpec G395M in- struments/modes since they are, according to Batalha & Line (2017), best suited for exoplanet transmission spec- troscopy. Finally, the JET code performed a statistical analysis for multiple transits and determined if the sim- ulated instrument spectra fit the model well enough to confirm a detection. Given current estimates of the pre- cision (noise floor) of these JWST instruments (as well as visual inspection of the simulated spectra after five tran- sits in Fig. 8), it would likely be very difficult to detect the Kepler-538b atmosphere even with a large number of transit observations with JWST. Perhaps other next-generation observatories such as the Thirty Meter Telescope (Sanders 2013), the Ex- tremely Large Telescope (Udry et al. 2014), the Giant Magellan Telescope (Johns et al. 2012), or the Large UV/Optical/IR Surveyor (The LUVOIR Team 2018) will be able to make such a project feasible. 5.3. Possibility of a Second Planet in the System Some early versions of the Kepler catalog included a weak transit signal at 117.76 days and labeled it as a planet candidate (K00365.02). However, one early cata- log instead labeled it as a false positive (Mullally et al. 2015) and the final Kepler catalog (DR25; Thompson et al. 2018) did not detect a candidate at that period at all (or even a threshold crossing event, the broadest detection category in the Kepler pipeline). Further, the Kepler False Positive Working Group (Bryson et al. 2017) investigated K00365.02 and could not determine a final disposition; they did however flag the candidate with a "Transit Not Unique False Alarm" flag, meaning "the de- tected transit signal is not obviously different from other Figure 7. Orbital period versus planet radius for all transiting exoplanets with P > 50 days and RV or transit timing variation (TTV) mass measurements. Data for all planets besides Kepler- 538b were retrieved from the NASA Exoplanet Archive (accessed 2019 February 16). Kepler-538b is plotted as a pink circle, all other exoplanets with RV mass measurements are plotted as black cir- cles, one exoplanet (Kepler-117c) has a jointly derived mass from RV and TTV measurements and is plotted as a black square, and exoplanets with only TTV mass measurements are plotted as gray triangles. (Period and radius uncertainties are plotted for all plan- ets, including Kepler-538b, but are smaller than the data points in many cases.) At long periods (P > 50 days), Kepler-538b is the smallest transiting exoplanet with an RV mass measurement, and Kepler-20d is the only such planet with a lower mass (by 0.5M⊕). Overall, there are very few mass measurements for planets in the long-period, small-radius regime of Kepler-538b. To begin with, there is no possibility of detecting an as- trometric signal of Kepler-538b. Perryman et al. (2014), which analyzed the expected planet yield from Gaia as- trometry, found that the expected along-scan accuracy per field of view for Gaia would be σfov = 34.2µas for a star like Kepler-538 (G = 11.67). While they required an astrometric signal of 3σfov for a detection, the astromet- ric signal of Kepler-538b is only 0.095 ± 0.022µas, over 1000 times smaller than this detection threshold. Similarly, a planet like Kepler-538b is very unsuitable for direct imaging. According to the NASA Exoplanet Archive34 (accessed 2019 July 28), there are no directly imaged planets less massive than 2MJup or closer to their host star than 2 au, both of which disqualify Kepler- 538b. Further, direct imaging is well suited for young stars which still host self-luminous planets, but the me- dian estimated age of Kepler-538 is 3.8 Gyr, older than nearly every host star of a directly imaged planet on the NASA Exoplanet Archive (there are only two exceptions, WISEP J121756.91+162640.2 A and Oph 11). Unlike astrometry and direct imaging, Penny et al. (2019) determined that a planet with the mass and semi- major axis of Kepler-538b would be just inside the mi- 34 https://exoplanetarchive.ipac.caltech.edu/ 6080100140200270Orbital Period (d)1.52.03.04.06.010.0Planet Radius (R)Kepler-538bRVRV + TTVTTV Kepler-538b 13 6. SUMMARY AND CONCLUSIONS In this paper, we analyze the Kepler-538 system in order to determine the properties of Kepler-538b, the single, known exoplanet in the system. Kepler-538 is a 0.924M(cid:12), G-type star with a visual magnitude of V = 11.27. We model the Kepler light curve and deter- mine the orbital period of Kepler-538b to be P = 81.74 days and the planetary radius to be Rp = 2.215+0.040 −0.034R⊕ (for reference, 0.037 = 236 km, approximately the dis- tance between Portland and Seattle37). These results are in agreement with previous transit fits. We also determine the planetary mass by accounting for stel- lar activity via a GP regression that uses information from the FWHM and RV observations simultaneously. Our model fit yields a mass estimate for Kepler-538b of Mp = 10.6+2.5−2.4M⊕. Combined, these results show the planet to have a density of ρp = 0.98±0.23ρ⊕ = 5.4±1.3 g cm−3. This suggests a composition and atmosphere somewhere between that of Earth and Neptune, with a significant fraction of ices (dominated by water ice), in addition to rocks/metals, and a small amount of gas (Zeng et al. 2018). To date, there have been very few precise and accu- rate mass measurements of long-period exoplanets. Be- yond 50 days, Kepler-538b is only the 11th transiting exoplanet with an RV mass measurement (NASA Ex- oplanet Archive38; accessed 2019 May 4). Additional, well-constrained mass measurements of long-period plan- ets will improve our understanding of the long-period ex- oplanet population. Beyond that, they will also help to answer questions about the short-period planet popula- tion, such as the nature of the planetary radius occur- rence gap (Fulton et al. 2017) and the effect of stellar flux on exoplanet compositions and atmospheres. With new, next-generation spectrographs such as HPF (Mahadevan et al. 2010, 2014), KPF (Gibson et al. 2016, 2018), EXPRES (Jurgenson et al. 2016), ESPRESSO (M´egevand et al. 2010), and NEID (Schwab et al. 2016) coming online now or in the near future, our ability to characterize long-period exoplanets will only improve. Better data will require more advanced analysis methods to extract as much information as possible. The methods used in this paper, such as GP regression, injection tests, and simultaneous modeling of RV observations and stel- lar activity indices, are valuable tools that strengthen the analysis of spectroscopic data, improve exoplanet char- acterization, and therefore better our understanding of the exoplanet population as a whole. A.W.M. is supported by the NSF Graduate Research Fellowship grant No. DGE 1752814. V.M.R. thanks the Royal Astronomical Society and Emmanuel College, Cambridge, for financial support. 37 https://www.distancecalculator.net/ from-portland-to-seattle 38 https://exoplanetarchive.ipac.caltech.edu/ This number was determined by constraining Orbital Period > 50 days, Planet Mass < 11MJup, Planet Mass Limit Flag = 0 (to remove upper limit results), Planet Circumbinary Flag = 0, Planet Transit Flag = 1, and Planet RV Flag = 1. Figure 8. A simulated transmission spectrum of Kepler-538b with five transits observed with JWST. The model spectrum, with low metallicity (five times solar) and no clouds, is shown as a gray line. The black data points are the simulated instrument spectra, using NIRISS SOSS-Or1 (0.81-2.81 µm) and NIRSpec G395M (2.87-5.18 µm). signals in the flux light curve."35 The radius of K00365.02 was reported on the NASA Exoplanet Archive as 0.62+0.10−0.03R⊕. Assuming a pure iron composition and using Zeng & Sasselov (2013) and Zeng et al. (2016) yields an upper limit mass of 0.37+0.25−0.05M⊕ and an upper limit semi-amplitude of 5.3+3.4−0.8 cm s−1, well below the detection threshold for HARPS-N, HIRES, or any other spectrograph. However, for the sake of rigor, we also ran a two planet model for Kepler-538b and K0035.02 on our RV and FWHM data (similar to our main model). Instead of jointly modeling photometry, we applied period and transit time priors on Kepler-538b and K00365.02 (the former based on our final results, the latter determined from the NASA Exoplanet Archive36; accessed 31 July 2019). Our results showed an RV semi- amplitude at 117.76 days of K = 0.26+0.28−0.18 m s−1, negli- gible and consistent with zero at less than 1.5σ. of Kepler-538b and Additionally, in or near a first-order mean K00365.02 are not motion resonance (or second-order, for that matter), so we do not expect a large, detectable transit timing variation (TTV) signal on Kepler-538b either (Lithwick et al. 2012). Indeed, the NASA Exoplanet Archive (accessed 31 July 2019) does not report a TTV flag for Kepler-538b. As a result, with an unverified transit signal, a negligible RV signal, and an apparently negli- gible TTV signal, the existence of K00365.02 remains inconclusive. the periods 35 https://exoplanetarchive.ipac.caltech.edu/\docs/API_ fpwg_columns.html 36 https://exoplanetarchive.ipac.caltech.edu/ 14 Mayo et al. C.D.D. acknowledges K2 Guest Observer 80NSSC19K0099. support program through from the NASA grant This work was performed in part under contract with the California Institute of Technology (Caltech)/Jet Propulsion Laboratory (JPL) funded by NASA through the Sagan Fellowship Program executed by the NASA Exoplanet Science Institute (R.D.H.) We acknowledge the support by INAF/Frontiera through the "Progetti Premiali" funding scheme of the Italian Ministry of Education, University, and Research. This publication was made possible through the sup- port of a grant from the John Templeton Foundation. The opinions expressed in this publication are those of the authors and do not necessarily reflect the views of the John Templeton Foundation. This work was supported in part by a grant from the Carlsberg Foundation. This paper includes data collected by the Kepler mis- sion. Funding for the Kepler mission is provided by the NASA Science Mission directorate. Some of the data presented in this paper were ob- tained from the Mikulski Archive for Space Telescopes (MAST). STScI is operated by the Association of Univer- sities for Research in Astronomy, Inc., under NASA con- tract NAS5 -- 26555. Support for MAST for non -- Hubble Space Telescope data is provided by the NASA Office of Space Science via grant NNX13AC07G and by other grants and contracts. Some of the data presented herein were obtained at the W. M. Keck Observatory (which is operated as a scien- tific partnership among Caltech, UC, and NASA). The authors wish to recognize and acknowledge the very sig- nificant cultural role and reverence that the summit of Maunakea has always had within the indigenous Hawai- ian community. We are most fortunate to have the op- portunity to conduct observations from this mountain. We would like to thank the HIRES observers who car- ried out the RV observations with Keck. We are grateful to Howard Isaacson for valuable proof- reading and suggestions as well as providing access to HIRES log R(cid:48) HK observations. Based on observations made with the Italian Telesco- pio Nazionale Galileo (TNG) operated by the Fundaci´on Galileo Galilei (FGG) of the Istituto Nazionale di As- trofisica (INAF) at the Observatorio del Roque de los Muchachos (La Palma, Canary Islands, Spain). The HARPS-N project has been funded by the Prodex Program of the Swiss Space Office (SSO), the Harvard University Origins of Life Initiative (HUOLI), the Scot- tish Universities Physics Alliance (SUPA), the University of Geneva, the Smithsonian Astrophysical Observatory (SAO), and the Italian National Astrophysical Institute (INAF), the University of St Andrews, Queens Univer- sity Belfast, and the University of Edinburgh. of data Space Agency This work has made use from the (ESA) mission Gaia European (https://www.cosmos.esa.int/gaia), processed by the Gaia Data Processing and Analysis Consortium (DPAC; https://www.cosmos.esa.int/web/gaia/ dpac/consortium). Funding for the DPAC has been provided by national in particular the institutions participating in the Gaia Multilateral Agreement. institutions, Facilities: Kepler, FLWO:1.5 m (TRES), TNG: (HARPS-N), Keck:I (HIRES), Gaia, NASA Exoplanet Archive, ADS, MAST Software: MultiNest (Feroz et al. 2009, 2013), Exo-Transmit (Kempton et al. 2017), Pandexo (Batalha et al. 2017), JET (C. D. Fortenbach & C. D. Dress- ing 2019, in preparation), isochrones (Morton 2015), lightkurve (Barentsen et al. 2019), BATMAN (Kreidberg 2015) REFERENCES Barbato, D., Sozzetti, A., Desidera, S., et al. 2018, A&A, 615, A175 Barentsen, G., Hedges, C., Vincius, Z., et al. 2019, KeplerGO/lightkurve: Lightkurve v1.0b29, doi:10.5281/zenodo.2565212 Barnes, S. A. 2010, ApJ, 722, 222 Barnes, S. A., & Kim, Y.-C. 2010, ApJ, 721, 675 Batalha, N. E., & Line, M. R. 2017, AJ, 153, 151 Batalha, N. E., Mandell, A., Pontoppidan, K., et al. 2017, PASP, 129, 064501 Benatti, S., Desidera, S., Damasso, M., et al. 2017, A&A, 599, A90 Berger, T. A., Huber, D., Gaidos, E., & van Saders, J. L. 2018, ApJ, 866, 99 Borucki, W., Koch, D., Basri, G., et al. 2008, in IAU Symposium, Vol. 249, Exoplanets: Detection, Formation and Dynamics, ed. Y.-S. Sun, S. Ferraz-Mello, & J.-L. Zhou, 17 -- 24 Bryson, S. T., Abdul-Masih, M., Batalha, N., et al. 2017, The Kepler Certified False Positive Table, Tech. rep. Buchhave, L. A., Latham, D. W., Johansen, A., et al. 2012, Nature, 486, 375 Chambers, J. 2018, ApJ, 865, 30 Cosentino, R., Lovis, C., Pepe, F., et al. 2012, in Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series, Vol. 8446, Ground-based and Airborne Instrumentation for Astronomy IV, 84461V Cosentino, R., Lovis, C., Pepe, F., et al. 2014, in Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series, Vol. 9147, Ground-based and Airborne Instrumentation for Astronomy V, 91478C Cutri, R. M., Skrutskie, M. F., van Dyk, S., et al. 2003, VizieR Online Data Catalog, 2246 Deming, D., Seager, S., Winn, J., et al. 2009, PASP, 121, 952 Feroz, F., Hobson, M. P., & Bridges, M. 2009, MNRAS, 398, 1601 Feroz, F., Hobson, M. P., Cameron, E., & Pettitt, A. N. 2013, ArXiv e-prints, arXiv:1306.2144 Fur´esz, G. 2008, PhD thesis, University of Szeged, Hungary Fulton, B. J., & Petigura, E. A. 2018, AJ, 156, 264 Fulton, B. J., Petigura, E. A., Howard, A. W., et al. 2017, AJ, 154, 109 Gaia Collaboration, Prusti, T., de Bruijne, J. H. J., et al. 2016, A&A, 595, A1 Gaia Collaboration, Brown, A. G. A., Vallenari, A., et al. 2018, A&A, 616, A1 Gardner, J. P., Mather, J. C., Clampin, M., et al. 2006, Space Sci. Rev., 123, 485 Gibson, S. R., Howard, A. W., Marcy, G. W., et al. 2016, in Proc. SPIE, Vol. 9908, Ground-based and Airborne Instrumentation for Astronomy VI, 990870 Gibson, S. R., Howard, A. W., Roy, A., et al. 2018, in Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series, Vol. 10702, Ground-based and Airborne Instrumentation for Astronomy VII, 107025X Gregory, P. C. 2007, MNRAS, 381, 1607 Haywood, R. D. 2015, PhD thesis, University of St Andrews, doi:10.5281/zenodo.35161 Haywood, R. D., Collier Cameron, A., Queloz, D., et al. 2014, MNRAS, 443, 2517 Howard, A. W., Johnson, J. A., Marcy, G. W., et al. 2010, ApJ, 721, 1467 Izidoro, A., Raymond, S. N., Morbidelli, A., Hersant, F., & Pierens, A. 2015, ApJ, 800, L22 Kepler-538b 15 Johns, M., McCarthy, P., Raybould, K., et al. 2012, in Proc. SPIE, Vol. 8444, Ground-based and Airborne Telescopes IV, 84441H Jurgenson, C., Fischer, D., McCracken, T., et al. 2016, in Proc. SPIE, Vol. 9908, Ground-based and Airborne Instrumentation for Astronomy VI, 99086T Kalirai, J. 2018, Contemporary Physics, 59, 251 Kempton, E. M.-R., Lupu, R., Owusu-Asare, A., Slough, P., & Cale, B. 2017, PASP, 129, 044402 Kipping, D. M. 2013a, MNRAS, 435, 2152 -- . 2013b, MNRAS, 434, L51 Kreidberg, L. 2015, PASP, 127, 1161 Kurucz, R. 1993, ATLAS9 Stellar Atmosphere Programs and 2 km/s grid. Kurucz CD-ROM No. 13. Cambridge, Mass.: Smithsonian Astrophysical Observatory, 1993., 13 Kurucz, R. L. 1992, in IAU Symposium, Vol. 149, The Stellar Populations of Galaxies, ed. B. Barbuy & A. Renzini, 225 Lindegren, L., Hern´andez, J., Bombrun, A., et al. 2018, A&A, 616, A2 Lithwick, Y., Xie, J., & Wu, Y. 2012, ApJ, 761, 122 Lopez, E. D., & Rice, K. 2018, MNRAS, 479, 5303 Mahadevan, S., Ramsey, L., Wright, J., et al. 2010, in Proc. SPIE, Vol. 7735, Ground-based and Airborne Instrumentation for Astronomy III, 77356X Mahadevan, S., Ramsey, L. W., Terrien, R., et al. 2014, in Proc. SPIE, Vol. 9147, Ground-based and Airborne Instrumentation for Astronomy V, 91471G Malavolta, L., Borsato, L., Granata, V., et al. 2017, AJ, 153, 224 Malavolta, L., Mayo, A. W., Louden, T., et al. 2018, AJ, 155, 107 Mandel, K., & Agol, E. 2002, ApJ, 580, L171 Mann, A. W., Dupuy, T., Kraus, A. L., et al. 2019, ApJ, 871, 63 M´egevand, D., Herreros, J.-M., Zerbi, F., et al. 2010, in Proc. SPIE, Vol. 7735, Ground-based and Airborne Instrumentation for Astronomy III, 77354Y Morbidelli, A. 2018, arXiv e-prints, arXiv:1803.06708 Mortier, A., Sousa, S. G., Adibekyan, V. Z., Brandao, I. M., & Santos, N. C. 2014, A&A, 572, A95 Morton, T. D. 2015, isochrones: Stellar model grid package, Astrophysics Source Code Library, ascl:1503.010 Morton, T. D., Bryson, S. T., Coughlin, J. L., et al. 2016, ApJ, 822, 86 Mullally, F., Coughlin, J. L., Thompson, S. E., et al. 2015, ApJS, 217, 31 Noyes, R. W., Hartmann, L. W., Baliunas, S. L., Duncan, D. K., & Vaughan, A. H. 1984, ApJ, 279, 763 Owen, J. E., & Wu, Y. 2017, ApJ, 847, 29 Penny, M. T., Gaudi, B. S., Kerins, E., et al. 2019, ApJS, 241, 3 Perryman, M., Hartman, J., Bakos, G. ´A., & Lindegren, L. 2014, ApJ, 797, 14 Petigura, E. A., Marcy, G. W., Winn, J. N., et al. 2018, AJ, 155, 89 Queloz, D., Henry, G. W., Sivan, J. P., et al. 2001, A&A, 379, 279 Rajpaul, V., Aigrain, S., Osborne, M. A., Reece, S., & Roberts, S. 2015, MNRAS, 452, 2269 Rasmussen, C. E., & Williams, C. K. I. 2006, Gaussian Processes for Machine Learning Raymond, S. N., Barnes, R., Veras, D., et al. 2009, ApJ, 696, L98 Sanders, G. H. 2013, Journal of Astrophysics and Astronomy, 34, 81 Scargle, J. D. 1982, ApJ, 263, 835 Schwab, C., Rakich, A., Gong, Q., et al. 2016, in Proc. SPIE, Vol. 9908, Ground-based and Airborne Instrumentation for Astronomy VI, 99087H Seager, S., & Mall´en-Ornelas, G. 2003, ApJ, 585, 1038 Sneden, C. A. 1973, PhD thesis, THE UNIVERSITY OF TEXAS AT AUSTIN. Sousa, S. G. 2014, ARES + MOOG: A Practical Overview of an Equivalent Width (EW) Method to Derive Stellar Parameters, ed. E. Niemczura, B. Smalley, & W. Pych, 297 -- 310 Sousa, S. G., Santos, N. C., Israelian, G., et al. 2011, A&A, 526, A99+ Sousa, S. G., Santos, N. C., Mortier, A., et al. 2015, A&A, 576, A94 Sozzetti, A., Torres, G., Charbonneau, D., et al. 2007, ApJ, 664, 1190 The LUVOIR Team. 2018, arXiv e-prints, arXiv:1809.09668 Thompson, S. E., Coughlin, J. L., Hoffman, K., et al. 2018, ApJS, 235, 38 Udry, S., Lovis, C., Bouchy, F., et al. 2014, ArXiv e-prints, arXiv:1412.1048 Van Eylen, V., Agentoft, C., Lundkvist, M. S., et al. 2018, MNRAS, 479, 4786 Vogt, S. S., Allen, S. L., Bigelow, B. C., et al. 1994, in Proc. SPIE, Vol. 2198, Instrumentation in Astronomy VIII, ed. D. L. Crawford & E. R. Craine, 362 Zechmeister, M., & Kurster, M. 2009, A&A, 496, 577 Zeng, L., Jacobsen, S. B., & Sasselov, D. D. 2017, Research Notes of the American Astronomical Society, 1, 32 Zeng, L., Jacobsen, S. B., Sasselov, D. D., & Vand erburg, A. 2018, MNRAS, 479, 5567 Zeng, L., & Sasselov, D. 2013, PASP, 125, 227 Zeng, L., Sasselov, D. D., & Jacobsen, S. B. 2016, ApJ, 819, 127 16 BJD Mayo et al. Table 3 RV observations and activity indicators, determined from the DRS. RV RV error (m s−1) FWHM (km s−1) BIS (km s−1) log10(R(cid:48) HK ) log10(R(cid:48) HK ) error Instrument (m s−1) −8.78 −1.24 −0.83 −5.95 4.58 −3.56 −1.33 −5.98 −6.02 −2.53 0.44 6.63 −0.23 1.05 −2.55 −3.55 2.47 5.47 0.39 7.76 5.79 1.27 −0.32 2.73 −0.74 2.59 2455402.854339 2455414.971547 2455486.859621 2455544.719659 2455760.087400 2455796.934228 2455797.920566 2455799.056115 2456114.931149 2456133.896429 2456147.919325 2456163.912379 2456164.801818 2456166.047782 2456166.759374 2456167.990464 2456451.100822 2456483.086583 2456486.833662 2456488.822611 2456494.987645 2456506.780605 2456507.968496 2456532.877110 2456830.887055 2456850.049952 2456828.616553 −37327.64 2456828.651774 −37320.76 2456829.664594 −37319.24 2456830.665375 −37319.62 2456831.690035 −37319.77 2456832.615999 −37314.68 2456833.672301 −37322.65 2456834.581042 −37315.83 2456834.677908 −37321.70 2456835.587887 −37318.54 2456845.576470 −37322.26 2456846.662015 −37327.96 2456847.656794 −37321.88 2456848.652903 −37327.73 2456849.657878 −37326.37 2456850.660745 −37324.98 2456851.654237 −37323.89 2456852.655703 −37316.98 2456853.657053 −37318.56 2456865.684262 −37320.32 2456866.681774 −37323.10 2456883.639193 −37324.45 2456884.647365 −37324.61 2456885.644031 −37322.71 2456886.642561 −37346.88 2456887.651622 −37322.33 2456888.580937 −37321.19 2456889.585275 −37324.05 2456903.541993 −37318.65 2456919.514886 −37322.96 2456922.547287 −37323.20 2456923.501548 −37320.74 2456924.510113 −37318.66 2456936.514073 −37326.01 2456939.418861 −37323.19 2456969.402685 −37323.75 2457106.734166 −37320.49 1.32 1.33 1.50 2.14 1.39 1.25 1.22 1.48 1.27 1.31 1.31 1.31 1.30 1.37 1.44 1.51 1.48 1.94 1.31 1.19 1.29 1.21 1.23 1.14 1.39 1.26 6.07 1.50 1.56 1.83 1.70 1.80 2.12 2.01 1.98 2.07 1.44 2.10 2.67 1.68 2.17 2.17 1.66 2.55 1.62 1.67 3.48 1.89 1.84 1.86 11.81 1.83 2.74 2.15 1.41 2.44 1.67 1.62 2.43 1.72 1.36 3.10 2.77 - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - 6.63443 −0.02220 6.66839 −0.03107 6.66376 −0.03206 6.67197 −0.03353 6.66691 −0.03495 6.66854 −0.03182 6.67240 −0.03957 6.66876 −0.03718 6.66877 −0.03961 6.65839 −0.02603 6.66668 −0.03079 6.65639 −0.02891 6.65969 −0.03820 6.66082 −0.03667 6.66712 −0.02913 6.66481 −0.03397 6.66590 −0.03297 6.67237 −0.03670 6.67355 −0.03209 6.66629 −0.02961 6.66279 −0.04100 6.65864 −0.03904 6.66539 −0.02953 6.66503 −0.03266 6.66764 −0.08202 6.67033 −0.03390 6.65114 −0.02926 6.65998 −0.03761 6.66569 −0.03568 6.66569 −0.03738 6.65621 −0.04098 6.66346 −0.03370 6.65831 −0.03765 6.65525 −0.03770 6.65561 −0.03499 6.65044 −0.03308 6.66677 −0.03337 −5.078 −5.003 −4.985 −4.971 −4.982 −4.974 −4.972 −5.005 −4.944 −4.933 −4.967 −4.922 −4.924 −4.927 −4.913 −4.912 −4.953 −4.943 −4.938 −4.926 −4.926 −4.931 −4.947 −4.940 −4.962 −4.954 −4.9522 −4.9819 −4.9788 −4.9882 −4.9634 −4.9975 −4.9553 −4.9764 −4.9768 −5.0090 −4.9739 −4.9885 −4.9938 −4.9900 −4.9558 −4.9575 −4.9625 −4.9994 −4.9818 −4.9818 −4.9629 −5.0004 −4.9982 −4.9753 −5.0609 −4.9883 −4.9577 −4.9716 −4.9851 −4.9735 −5.0131 −5.0026 −4.9763 −4.9864 −4.9986 −5.0094 −4.9801 - - - - - - - - - - - - - - - - - - - - - - - - - - 0.0699 0.0106 0.0111 0.0146 0.0128 0.0150 0.0182 0.0169 0.0169 0.0197 0.0097 0.0184 0.0273 0.0131 0.0182 0.0182 0.0121 0.0260 0.0122 0.0130 0.0388 0.0164 0.0166 0.0159 0.2193 0.0152 0.0266 0.0189 0.0095 0.0224 0.0146 0.0127 0.0237 0.0143 0.0093 0.0377 0.0289 HIRES HIRES HIRES HIRES HIRES HIRES HIRES HIRES HIRES HIRES HIRES HIRES HIRES HIRES HIRES HIRES HIRES HIRES HIRES HIRES HIRES HIRES HIRES HIRES HIRES HIRES HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N Kepler-538b Table 3 -- Continued 17 RV error (m s−1) 1.51 1.87 1.61 2.87 12.94 2.14 2.03 1.70 1.60 1.75 1.58 1.81 2.35 1.51 1.52 1.58 2.63 1.61 1.57 1.67 1.44 1.65 1.37 1.78 3.17 3.63 3.17 2.58 1.67 2.72 2.51 3.34 2.41 2.36 1.88 2.04 2.21 1.69 1.36 1.44 1.46 1.47 1.79 2.40 1.89 1.64 BIS FWHM (km s−1) (km s−1) 6.64918 −0.03136 6.64875 −0.03660 6.64485 −0.03692 6.64814 −0.04127 6.66453 −0.05998 6.65282 −0.03562 6.65589 −0.04059 6.65201 −0.04164 6.64953 −0.03798 6.65033 −0.03871 6.64797 −0.03663 6.64944 −0.03550 6.65649 −0.03450 6.65258 −0.03701 6.65510 −0.03815 6.65099 −0.03864 6.64860 −0.03852 6.64984 −0.03343 6.65434 −0.03606 6.65251 −0.03217 6.65760 −0.03912 6.65239 −0.03455 6.64834 −0.04333 6.65155 −0.04076 6.63767 −0.04077 6.64970 −0.03534 6.66790 −0.02615 6.65109 −0.03404 6.66113 −0.03348 6.66677 −0.02712 6.65780 −0.02953 6.66154 −0.02550 6.66627 −0.03812 6.64870 −0.04351 6.65391 −0.03681 6.64591 −0.03502 6.64810 −0.04496 6.65385 −0.03696 6.65014 −0.03809 6.65177 −0.03459 6.64860 −0.04486 6.65573 −0.03632 6.65019 −0.03970 6.64231 −0.03293 6.65257 −0.03583 6.65030 −0.04224 log10(R(cid:48) HK ) log10(R(cid:48) HK ) error Instrument −5.0282 −5.0338 −5.0174 −5.0063 −4.9161 −4.9989 −5.0130 −5.0095 −5.0160 −5.0016 −5.0014 −4.9890 −5.0265 −5.0096 −5.0237 −5.0010 −5.0649 −5.0211 −5.0243 −5.0374 −5.0097 −4.9930 −5.0237 −4.9982 −5.0049 −5.0279 −5.0750 −5.0105 −5.0084 −4.9272 −5.0054 −4.9889 −5.0093 −4.9894 −4.9651 −5.0136 −5.0215 −5.0211 −5.0301 −5.0336 −5.0183 −5.0121 −5.0308 −5.0135 −5.0053 −5.0085 0.0120 0.0181 0.0141 0.0321 0.1624 0.0202 0.0185 0.0140 0.0127 0.0146 0.0122 0.0145 0.0237 0.0115 0.0118 0.0123 0.0316 0.0135 0.0128 0.0147 0.0107 0.0132 0.0101 0.0150 0.0375 0.0478 0.0424 0.0277 0.0136 0.0251 0.0266 0.0385 0.0260 0.0239 0.0149 0.0193 0.0224 0.0145 0.0099 0.0110 0.0109 0.0111 0.0159 0.0243 0.0167 0.0131 HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N HARPS-N RV BJD (m s−1) 2457116.717298 −37324.15 2457118.706394 −37327.91 2457121.726137 −37324.71 2457153.685174 −37323.84 2457156.714776 −37324.15 2457159.642662 −37323.22 2457160.638323 −37320.99 2457161.626357 −37324.63 2457180.658376 −37322.50 2457181.686408 −37322.02 2457182.670828 −37322.60 2457183.652886 −37321.34 2457184.643705 −37324.94 2457185.662466 −37324.52 2457186.662672 −37325.81 2457188.679310 −37328.26 2457189.672084 −37322.40 2457190.685669 −37325.59 2457191.685746 −37323.07 2457192.684342 −37324.65 2457193.684869 −37323.90 2457195.594752 −37318.06 2457221.626801 −37324.84 2457222.569536 −37324.99 2457223.579194 −37322.72 2457225.522395 −37319.94 2457226.582966 −37317.08 2457226.606265 −37321.98 2457227.627333 −37316.52 2457228.630703 −37320.68 2457229.584812 −37316.53 2457230.528273 −37320.18 2457254.631919 −37325.16 2457256.398580 −37314.12 2457257.413714 −37315.46 2457267.435503 −37324.84 2457268.492540 −37322.86 2457269.418733 −37320.32 2457270.407599 −37321.14 2457271.408119 −37325.01 2457273.426969 −37326.84 2457301.384627 −37319.51 2457302.383904 −37318.58 2457321.426080 −37318.14 2457330.417736 −37318.77 2457334.397358 −37321.22
1008.2209
1
1008
2010-08-12T20:05:10
Trans-Neptunian Objects with Hubble Space Telescope ACS/WFC
[ "astro-ph.EP" ]
We introduce a novel search technique that can identify trans-neptunian objects in three to five exposures of a pointing within a single Hubble Space Telescope orbit. The process is fast enough to allow the discovery of candidates soon after the data are available. This allows sufficient time to schedule follow up observations with HST within a month. We report the discovery of 14 slow-moving objects found within 5\circ of the ecliptic in archival data taken with the Wide Field Channel of the Advanced Camera for Surveys. The luminosity function of these objects is consistent with previous ground-based and space-based results. We show evidence that the size distribution of both high and low inclination populations is similar for objects smaller than 100 km, as expected from collisional evolution models, while their size distribution differ for brighter objects. We suggest the two populations formed in different parts of the protoplanetary disk and after being dynamically mixed have collisionally evolved together. Among the objects discovered there is an equal mass binary with an angular separation ~ 0."53.
astro-ph.EP
astro-ph
Accepted to ApJ: August 10, 2010 Trans-neptunian Objects with Hubble Space Telescope ACS/WFC1 Cesar I. Fuentes2,3, Matthew J. Holman2, David E. Trilling3, Pavlos Protopapas2 ABSTRACT We introduce a novel search technique that can identify trans-neptunian ob- jects in three to five exposures of a pointing within a single Hubble Space Telescope orbit. The process is fast enough to allow the discovery of candidates soon after the data are available. This allows sufficient time to schedule follow up obser- vations with HST within a month. We report the discovery of 14 slow-moving objects found within 5◦ of the ecliptic in archival data taken with the Wide Field Channel of the Advanced Camera for Surveys. The luminosity function of these objects is consistent with previous ground-based and space-based results. We show evidence that the size distribution of both high and low inclination populations is similar for objects smaller than 100 km , as expected from colli- sional evolution models, while their size distribution differ for brighter objects. We suggest the two populations formed in different parts of the protoplane- tary disk and after being dynamically mixed have collisionally evolved together. Among the objects discovered there is an equal mass binary with an angular separation ∼ 0.(cid:48)(cid:48)53. Subject headings: Kuiper Belt -- Solar System: formation 1. Introduction Trans-neptunian objects (TNOs) represent the leftovers of the same planetesimals from which the planets in the solar system formed. These offer a unique opportunity for testing 1Based on observations made with the NASA/ESA Hubble Space Telescope, obtained from the Data Archive at the Space Telescope Science Institute, which is operated by the Association of Universities for Research in Astronomy, Inc., under NASA contract NAS 5-26555. These observations are associated with program 11778. 2Harvard-Smithsonian Center for Astrophysics, 60 Garden Street, Cambridge, MA 02138, USA; [email protected] 3Department of Physics and Astronomy, Northern Arizona University, PO Box 6010, Flagstaff, AZ 86011 -- 2 -- theories of the growth and collisional history of planetesimals and the dynamical evolution of the giant planets (Kenyon & Bromley 2004; Morbidelli et al. 2008). The study of the orbital distribution of TNOs has shown the existence of at least two distinct dynamical populations (Levison & Stern 2001; Brown 2001) with different colors (Doressoundiram et al. 2008) and size distributions (Bernstein et al. 2004; Fuentes & Holman 2008). Most of what is known about TNOs is based on follow-up studies of the brightest objects (Brown 2008). The bias toward analysis of brighter objects can be seen in challenging obser- vations like lightcurves and binarity fraction. This is even more apparent for spectroscopic observations and albedo measurements, which are available for only ∼ 30 objects (Stans- berry et al. 2008; Brucker et al. 2009), among which the smallest is over 130 km in diameter. This is explained by the relative faintness of outer solar system bodies and the difficulty of tracking them after discovery. Observations made several months and even years apart are needed to secure accurate orbits. In general the fainter the object the more demanding the observing conditions necessary to detect and track it. Despite the challenge, a great deal of effort has been dedicated to searching for faint TNOs (Chiang & Brown 1999; Gladman et al. 2001; Allen et al. 2002; Bernstein et al. 2004; Petit et al. 2006; Fraser et al. 2008; Fuentes & Holman 2008; Fraser & Kavelaars 2009; Fuentes et al. 2009). These surveys have concentrated near the ecliptic, where the sky plane density of objects is largest. Elaborate observational techniques have been developed to extend the sensitivity of these surveys. Usually a compromise is reached between the sky coverage and magnitude depth of these resource intensive techniques. This results in "pencil beam" searches that concentrate on a limited region of the sky. The results produced are statistically calibrated and provide a precise assessment of the TNO sky plane density. However these surveys typically obtain short arcs, yielding imprecise information about TNO orbits. These studies have extended our understanding of the TNO size distribution to tens of km in diameter. In the deepest survey to date, reaching a limit of R ∼ 28.5, Bernstein et al. (2004) recognized a break using HST data at R ∼ 25. At bright magnitudes the luminosity function was consistent with the power law behavior surveys carried out from the ground had measured. However those searches claimed the luminosity function of bright objects could be extended up to a magnitude R ∼ 26 (Gladman et al. 2001; Petit et al. 2006). The controversy was settled when Fuentes & Holman (2008) corroborated the existence of the break. That work had the advantage of being a single survey with the sky coverage and magnitude depth to be sensitive to R ∼ 25.5 objects and obtained a statistically significant result that did not rely in the combination of fields observed under various conditions. Deeper ground based searches have been able to narrow the gap between ground and space based -- 3 -- surveys by coadding data taken over an entire night (Fraser & Kavelaars 2009; Fuentes et al. 2009). Brown (2001) determined that the TNO inclination distribution was well fit by the sum of a narrow and wide gaussian distribution. Bernstein et al. (2004) used the somewhat arbitrary value of i = 5◦ to differentiate between hot and cold objects and recognized different size distributions for both populations. They determined the size distribution of hot objects had a shallower slope than that of cold objects for objects larger and smaller than the break. Observationally most large objects are hot and most small objects are cold. However, this result was based on a few objects smaller than the break, especially on the three cold TNOs found by Bernstein et al. (2004). A simple definition for hot and cold objects is useful for pencil beam surveys where the constraint on the orbits is not precise. From the ground not much more than a rate of motion on the sky is obtained from a night's worth of observation. The large uncertainties associated with the distance and inclination estimated under the assumption of a circular orbit could eventually bias the analysis. A survey able to find faint objects (R ∼ 26) and provide accurate constrains on the distance and inclination could show if indeed there is a difference in the size distribution of high and low inclination objects. The most basic information that can be extracted from a set of TNO discoveries is the luminosity function. If the albedo is assumed and the distance to each object can be esti- mated, the size distribution is obtained. With further information about the trajectory of an object we can estimate its inclination, which can be used as a proxy for dynamical excitation. Ground based detections provide a very short arc that gives us limited information about the distance if the degeneracy between the object's velocity and parallactic motion cannot be disentangled. The HST has the advantage of not being affected by atmospheric seeing, achieving very precise astrometric measurements. Also, its orbital motion about the Earth adds extra par- allax to the observations. For Solar System bodies this helps in unraveling the contributions of the Earth's parallax and the object's intrinsic motion, allowing precise orbital estimates, even when not observing at opposition. The objective of our investigation was to find faint TNOs with acceptable orbital uncer- tainties to further constrain the size distribution of the hot and cold populations. For this we defined a limited, well characterized search for moving objects. Our search is sensitive to R ∼ 26 and is able to constrain the distance and inclination of the objects discovered. In Section §2 we present a summary of the data selection and acquisition. The characteriza- tion of the search algorithm is done by sampling a control population, described in §3. The -- 4 -- detection pipeline is described in detail in §4, and the detection efficiency is explained in §5. Results and analysis of the data appear in §6, where special emphasis is given to testing the capabilities of HST in finding the correct orbital information. We discuss the significance of our findings in §7. Our conclusions appear in §8. 2. Data Objects in the TNO realm (∼ 42 AU ) exhibit parallactic motion of ∼ 3 (cid:48)(cid:48) h−1 when observed at opposition, mainly due to Earth's translation. Depending on the resolution and data quality of the observations TNOs are readily identified by this motion if two or more images of the same field are taken with an adequate interval between exposures. This parallactic motion implies that if the shutter is kept open for a time longer than it takes a TNO to move beyond its PSF, the image will trail. If observed at opposition, an image of a typical TNO will take ∼ 10 min to traverse the PSF of a ground based image (FWHM ∼ 0.(cid:48)(cid:48)5) while only 1 min in an image taken with the Wide Field Channel of the Advanced Camera for Surveys (ACS/WFC, FWHM ∼ 0.(cid:48)(cid:48)05). We focused our search on data taken with ACS/WFC, the largest field of view camera on HST (202(cid:48)(cid:48) × 202(cid:48)(cid:48) or 0.003 deg2.) Bernstein et al. (2004) coadded tens of ACS/WFC exposures to reach a sensitivity of R ∼ 28.5. However, of the three objects they discovered two of them were detected in each individual image, and the faintest (R = 27.8) exhibited a lightcurve that made it visible in a fraction of the exposures. The latest results for the TNO luminosity function (Fuentes et al. 2009; Fraser & Kavelaars 2009) indicate that the sky density of TNOs brighter than R = 27 on the ecliptic is 0.5 per ACS/WFC field. The lack of the degrading effect of the atmosphere compensates for the relatively small size of HST's 2.4m mirror. The archive provides numerous data of different targets, with different filters and science goals. The ACS/WFC data provided by the Space Telescope Science Institute (STScI) is quite homogeneous in its format which allows us to build software that can apply a standard processing procedure for all data considered in this project. Most exposure times are ∼ 500s in order to maximize the open shutter time. In addition, it is customary that longer observations be divided in a number of shorter exposures, allowing a median rejection of cosmic rays. This, typically three or more exposures of a field are obtained in sequence. Access to the HST's electronic archive is provided by the Multimission Archive at STScI (MAST) (archive.stsci.edu). -- 5 -- 2.1. Field Selection We considered observations obtained within 5◦ of the ecliptic, where the sky density of TNOs is highest, as their orbits are concentrated near the ecliptic (Brown 2001). Figure 1 shows the distribution of fields we considered. It is common for ground based surveys to prioritize fields located at opposition. This maximizes the parallactic motion with respect to the object's intrinsic velocity, allowing a reasonable 10 − 20% uncertainty estimate on the distance if a circular orbit is assumed. It also permits a clear distinction of nearer, main belt asteroids from TNOs. Given the superior resolution of ACS/WFC data and the extra parallax derived from the motion of HST itself it is not necessary to observe at opposition to constrain the distance to a moving object without having to rely on a circular orbit. For this reason, we did not restrict out attention to a specific range of solar elongation. We consider images of the same field taken within the same HST orbit as part of a pointing. Only pointings that had a total open shutter time of over 1,000 seconds with three or more images were considered. These images are typically taken within half an HST orbital period, ∼ 48 min. A total of 150 pointings were recognized as satisfactory for this project. We specifically excluded the many observations taken for the work by Bernstein et al. (2004), as those were previously searched for TNOs. STScI makes available data in any step of the reduction process. We selected flat-fielded images that had not been undistorted or combined. We used the distortion corrections and PSF models for various filters provided by Anderson & King (2000). The filters that we considered for this work are summarized in Table 1. 2.2. Astrometric Solution We search for Solar System objects that change position between images over the time span of the pointing. The best astrometry possible is necessary to obtain a precise trajectory for TNOs. The astrometric accuracy provided by the archive's calibration is only as good as the astrometric precision of the HST guide star catalog provided for the "astrometric reference". However, the differential astrometry can be as accurate as the HST's spatial resolution. Instead of combining data taken in different pointings, we took advantage of the pre- cise differential astrometry between images obtained during the same orbit of the HST. ACS/WFC provides exquisite resolution (∼ 50 mas) and a stable and nearly constant PSF -- 6 -- Fig. 1. -- Map of the sky in J2000 coordinates. The 10◦-wide ecliptic band we chose to select our pointings from is shown in blue. The location of all targets considered is plotted as yellow triangles. Many pointings are superimposed on top of each other. (Anderson & King 2000) across the field of view. There is, however a significant large scale distortion, that needs to be accounted for before detections in different images can be com- pared with each other. In order to obtain a consistent astrometric solution for all the images in a single pointing we used the distortion solution in the software developed by Anderson & King (2000) that considers filter-dependent distortions at the pixel level. We defined the first image in the sequence as the "astrometric reference". All images were searched for sources and their positions transformed to the undistorted frame. All images in the pointing are registered to the reference by matching common sources. All detections are transformed to J2000 coordinates via the astrometric information in the refer- ence image provided by STScI. In this way there is a unique transformation from a detection in any image to J2000 coordinates. This transformations from the image array to sky co- ordinates is readily inverted to be used during the implanting of the control population, to be discussed in the next section. For static sources the uncertainty in the position was very close to the 50 mas advertised by the ACS/WFC's documentation. -- 7 -- 3. Control Population Our moving object detection method requires the identification of an object in at least three different images. There are various reasons for an object in the field to go undetected. Being too faint is the most common. Chance alignment with a background star or cosmic ray (CR) will also reduce the chances of finding an object. Chip gaps and bad pixels should be taken into account also when considering the detection efficiency. We measure these and other unknown effects with the use of a control population that covers the range of observational characteristics the TNO population is expected to exhibit. We implant our control population in the original flat-fielded images, before any distor- tion correction is applied. Since these objects go through the pipeline with the original data, anything that would affect our ability to detect real faint objects will also affect our ability to detect the objects that were implanted. During the visual examination phase, to be discussed later, the operator is presented with thousands of candidate moving objects. This provides a constant stream of objects moving in TNO-like trajectories; real objects are indistinguishable from the control popula- tion. For the detection of new objects the most important characteristics to be modeled are the brightness and rate of motion distribution for the synthetic population. 3.1. Apparent Motion The control population also provides a test for the reduction pipeline, the recovery software, and the visual examination. The analysis pipeline, including the human interaction, will be successful if it discovers real objects that do not look special among the control population. This means the control population should be accurate and look like real TNOs. To avoid being biased toward finding exactly what we expect, based on what we already know about TNOs, we require a control population that spans all realistic properties (for example, orbits, colors, lightcurves, and binarity.) However, for simplicity, we considered only single (not binary TNOs), showing no brightness variation in a ∼ 40 min interval and with normal TNO colors (See subsection 6.1.) We only considered bound orbits. In order to have both an accurate and inclusive TNO control population, we considered two different parameterizations. We first used the Keplerian orbital elements of an object to produce ephemerides. This allowed us to produce a distribution of orbital parameters similar to that of TNOs. The second one was based on the Bernstein & Khushalani (2000) elements which considers a cartesian grid centered on the position of the observer at the time of observation. These elements are closely related to, and therefore a better measure -- 8 -- of, the observational rates of motion. For this reason this method is more inclusive. Equal number of objects, usually 200, produced with each approach were implanted in each point- ing. Ephemerides produced by these two methods using independent pieces of software, a variation of Bernstein & Khushalani (2000)'s Orbfit and a custom made integrator, are used by the procedure that inserts synthetic objects in the original images. 3.2. Brightness Distribution For any given pointing we considered a uniform distribution in the instrumental mag- nitude. The faint end of this distribution was selected based on the reported instrumental zeropoint and exposure times of the images. We selected the magnitude distribution so that it would yield ∼ 50 detected objects per pointing, enough objects to sample the efficiency function of each field. The magnitude range spans 2.5 magnitudes, and the faintest object was chosen to be half a magnitude fainter than the faintest object that should appear as a 1-sigma detection in an individual image. As objects will trail over the course of an integration, our software computes the object's position at the beginning, middle and end of the exposure based on its orbital parameters and the position of the HST at the time of the exposure. We fit a 2nd-degree polynomial to this motion and then subdivide that motion in 1-pixel increments. We then divide the object's flux by the number of positions and insert a normalized PSF model at each position for that particular filter (Anderson & King 2000). Based on the position on the array we can also correct the brightness of a source as another effect of the geometrical distortion. Since the photometric uncertainty of the objects that we are interested in are background-noise limited, no additional noise was added to trailed PSFs. 4. Detection of Moving Objects The usual strategy for finding TNOs that are detectable in single images relies in testing all correlations between detections across images that are consistent with a TNO orbit. From the ground the image quality is such that exposure times of some minutes can be used before trailing is an issue. Then the we search for correlations between point-like sources that move from image to image. For observations taken over a single night of observation the algorithm takes the list of detections and finds subsets that follow a straight line with a constant rate. The diffraction limited resolution with respect to the ground and the apparent motion induced by HST orbiting the Earth imply that TNOs detections will be trailed in typical -- 9 -- exposure times (500 s). This trailing spreads an object's flux over a larger number of pixels, which for background limited observations significantly decreases the likelihood of finding a faint moving object. For this project we have taken a distinct approach that takes advantage of this apparent difficulty. Since all TNO detections will be trailed to some degree, analyzing a single set of detections (the centroid of the trails) is not the optimal method. This trailing motivates the overall strategy of our survey. We explore the range of orbital parameters consistent with the TNO region and keep those that produce significantly different trails, keeping them as test trails. Then, the search for sources in each image is optimized to select objects that show the particular test trail. Sources are then correlated in the same way ground-based observations are, and considering motion from image to image that is consistent with the test trail considered. 4.1. Detection using Optimized Kernel Search Searching for all possible orbits in the trans-neptunian space requires an algorithm capable of sampling the complete set of observational features that real TNOs could exhibit. This usually translates into a set of possible rates in R.A. and Dec. that are surveyed with a rate resolution finer than that set by FWHM/∆t, where ∆t is the time span of the observations. For HST data this is a bit more complex than ground based observations. The extra parallax due to the motion of HST around the Earth is ∼ 0.(cid:48)(cid:48)4, significantly larger than the astrometric uncertainty (0.(cid:48)(cid:48)05) implying some structure in a single detection can be identified even for TNOs. This motivated us to use an optimized kernel search. Instead of taking the point-source catalog for every image and searching those for position correlations consistent with any orbit we consider a set of orbits and search the images for detections consistent with them. Each search is performed on the convolution of the original image and a kernel designed to match the signature of an object with the orbit being surveyed. This has the advantage of lowering the number of artifacts while increasing our sensitivity to moving objects. The use of kernels does not affect the photometry as it is only used for detecting sources. The kernels were computed on the fly in the same way a moving object is implanted, fitting a 2nd-degree polynomial to its motion and implanting a set of 1-pixel separated PSFs on that track. The detection was performed using SExtractor (Bertin & Arnouts 1996) that has built in the use of kernels. The number of orbits considered depends on the kernels; if two orbits produce kernels that differ by less than a pixel in all images then only one is used. For every orbit considered a catalog is generated and fed into our search algorithm. -- 10 -- This takes the same set of orbital parameters used to create the kernel and constructs a shift matrix δij, that indicates by how much would an object with the current orbit in image i move from image j. Every detection in all images is at considered as a possible moving object detection, and all other detections are tested for a possible link with it. Detections are linked by proximity to the putative new position using a threshold equivalent to the astrometric precision. Only links of three or more detections are considered viable moving objects. We then filter solutions so that any detection belongs only to one possible moving object, using the astrometric error of the orbital solution with respect to the detections as the parameter to rank these links. The result is a list of candidate objects, each characterized by a set of detections. 4.2. Visual Examination At this stage the list of candidate objects is presented to a human operator to distinguish moving objects from spurious detections. Up to this point the pipeline is fully automatic with no step in the reduction process requiring the input of an operator. The list produced is the pipeline's best guess for which detections appear to be related by a plausible orbit. However, no matter how efficient the processing might be, the whole pipeline relies on the positional information derived by Sextractor. It is nearly impossible to avoid chance align- ment of spurious detections or poorly subtracted cosmic rays, for example, and to program an automatic selection algorithm that could flag these events would be even harder. The human brain is incredibly good at finding patterns. We make use of this fact by presenting the detections as a pattern recognition problem to a human operator. Each candidate is represented by an animated postage stamp of the area around its location in the original image and the one that was CR-removed, both with the detections clearly marked. Both images are embedded in an webpage that gives the option of flagging the object as moving object or as an artifact. Information about the detections are also made available to the observer. It usually takes ∼ 3 min to a trained operator to flag all objects in a field as moving or artifact. On average the operator is presented with ∼ 100 objects per pointing and nearly half of the detections that go through the human filter were recognized as artifacts. These usually correspond to: chance alignment of cosmic-rays (readily recognized due to their poor fit and for appearing much brighter than reported), extended objects elongated in the direction of the ecliptic (galaxies, saturated stars' wings), etc. -- 11 -- It is only now that the list of selected objects is compared to that of implanted ones. Those that are related to a synthetic object are used to characterize the detection efficiency of our method, and those that are "real" moving objects are flagged for constructing the luminosity function. In all pointings considered in this project we recovered over 5,000 fake objects, many times more than the 14 real objects discovered. The fact that real objects (apart from a binary) were indistinguishable from the implanted objects is a sign that the search is well described by our control population. 5. Detection Efficiency After the list of implanted objects is revealed and we correlate it with that of the objects found, the next step is computing the efficiency function. The likelihood of obtaining a particular set of objects from the control population depends on the efficiency function η(R). This likelihood function Lη has the form: η(Ri) × N−(cid:89) (cid:20) R − R50 j=1 erfc (cid:21) N +(cid:89) i=1 A 2 Lη = η(R) = [1 − η(Rj)] , 2 w (1) (2) the probability of finding a set of objects (1, . . . , N +) and of not finding the complement (1, . . . , N−). The parameters in η are the maximum efficiency (A), the magnitude at which the detection probability equals half that of the maximum (R50) and the width of the decline in probability (w). We search for those values that maximize Lη. In general each pointing may be considered as an independent survey with its own detection efficiency and an area equal to ACS/WFC's field of view. However, this is only true for uncorrelated observations, where there is no chance of finding the same object in distinct pointings. Since we are using archival data, there are many consecutive observations of the same field that were included in our survey, where the probability of "discovering" an object twice is non-negligible. If this effect is not considered appropriately we will over estimate the area surveyed and consequently underestimate the TNO luminosity function. This problem may be solved if we know how many of the real TNOs would move from pointing to pointing. This requires that we have an accurate model for the orbital distribution of the TNO population, since different distributions will yield different levels of "contamina- tion". The main variable determining how much an object appears to move is its distance to the observer. We used a distribution that resembles the heliocentric distribution in Fuentes -- 12 -- & Holman (2008). The density of trial objects will determine the statistical significance of the overlapping areas between pointings. We could use a single population with an accurate orbital distribution and a high density for the entire area of the sky. However, as the field of view of all our targets is negligible compared to the area from which they were selected a density of 10 per pointing would yield ∼ 108 objects for which orbits and ephemeris would have to be computed. Additionally, a different control would be necessary to test the detection efficiency of uncommon, but physically plausible orbits. Instead we chose to separate the problem into detection efficiency and effective area. The detection efficiency is well sampled for every pointing, as described above. In order to account for the area that was observed by more than one pointing we find all intersections between related pointings. The area that intersects two pointings corresponds to the fraction of the area in a field where a TNO could have been detected twice. Since we are dealing with moving objects we need to take into account the orbit distribution of the real TNO population. We use a swarm of fake bodies in each pointing to estimate the overlap. We first identify plausibly correlated pointings by their observing time and location, 64 such sets were found. We then created a large population of mock orbits (1,000) with similar characteristics to the real TNO population in each one of those pointings and computed how many fell in the field of view of each other. The result becomes a bit more complicated when we consider that an object could be in 3 or more of those pointings, each with its own efficiency function. In our survey we had a maximum of 5 pointings that were correlated and could identify and precisely account for all intersection areas that were surveyed more than once. The effective area (Ωeff) in Fig. 4 has over 2,000 parameters and was computed as shown in Eq. 3. (cid:88) N(cid:88) ηS = 1 −(cid:89) S∈Pi i=1 Ωeff = ΩSηS (1 − ηo) , (3) (4) o∈S where the sum indexed i is carried out over all N sets of related pointings. If a set has ni pointings, then there are 2ni possible combinations of overlapping fields or subsets S in its power set Pi. Each one of those subsets represent an area ΩS that was surveyed with a detection efficiency ηS. The detection efficiency of the subset is the probability of being detected in any of the pointings in it. The computation of ΩS is provided by the fake bodies. It is the same fraction of a field's area as the fraction of objects created in any of the pointings in S that end up in all pointings in S. Given the large number of fields and filters considered, and despite the many consider- ations the shape of the effective area does not vary too sharply compared to the total area and can be approximated as a function with only four parameters: -- 13 -- (cid:20)R − R25 (cid:21) 2 w1 (cid:20)R − R25 (cid:21) 2 w2 Ωeff(R) ≈ A 4 erfc erfc , (5) where the maximum effective area A = 0.28±0.01 deg2, the magnitude at which the detection efficiency is 25% of its maximum R25 = 26.5 ± 0.1, and the width of the decline in efficiency is parameterized as w1 = 0.78± 0.3 and w2 = 0.31± 0.3. In similar surveys it is conventional to define the magnitude at which Ωeff(R) is half its maximum, which for our survey is R50 = 26.14. The maximum effective area 0.28 deg2 is comparable to the total area surveyed 0.45 deg2. These differ due to the many fields that effectively sampled the same objects. The effective area is shown in the top panel of Figure 4. After the real objects are recognized and some members of the control population are identified among the detected objects we analyze the photometry and astrometry of each one of those detections. We construct the efficiency function and luminosity function. The orbital constraint on every object is also investigated to understand the uncertainties and possible degeneracies imposed by the data. 6. Analysis 6.1. Photometry After sources are detected SExtractor (Bertin & Arnouts 1996) is used to obtain pho- tometry. As was discussed in §2 TNOs will shift their position during the exposure and their shape will be elongated in the direction of motion. We selected the AUTO flux mea- surement since it is the most appropriate for extended objects. It uses an elliptical aperture which is computed for every detected source. Instrumental magnitudes are provided for each discovered object in Table 2. For the range of magnitudes that is relevant for this study the photometric uncertainty is dominated by the noise in the background illumination. This uncertainty is also computed by SExtractor and is in good agreement with the deviation between different images and with the error for implanted objects. The photometric accuracy depends mainly on the background brightness and the filter used. The suite of filters that we considered for this project is presented in Table 1. Transfor- mations between ACS/WFC magnitudes and UBVRI standard magnitudes were computed based on (Sirianni et al. 2005; Jordi et al. 2006) and considered typical colors for TNOs -- 14 -- based on Doressoundiram et al. (2008) (V-R = 0.6; R-I = 0.6; B-R = 1.6) 6.2. Orbital Information Though we use the position of the objects in each image and its trail over a single exposure to find orbits that are consistent with an objects' motion, we obtain a tighter constraint if we simulate the images themselves. We use the stability of HST's PSF, and its angular resolution to find the set of orbital parameters that are consistent with the data and in this way provide accurate uncertainty estimates for them. Our ability to constrain the range of orbital parameters for a given object is greatly improved by the motion of the telescope during a pointing. The extra parallax and precise astrometry provided by HST allow us to better disentangle the parallactic and proper motion of an object during the exposure. The motion of HST will produce a parallax for any motion perpendicular to the ecliptic that will be evident as a curved path in the image. For fields at low ecliptic latitude the component along the ecliptic is largest and changes with time, as the target "rises", "transits", and "sets" with respect to Earth throughout the pointing. On images with equal exposure time this will manifest as a set of streaks with different lengths. We run a Markov Chain Monte Carlo (MCMC) simulation where the function to min- imize is the residual on the objects' image. To compute the χ2 we consider a rectangular region around each detection that depends on the shape of the trail and the uncertainties in the data. We parameterize this function on variables related to the observed motion on the sky. We constrain the number of parameters we fit for to those that affect the orbit of the objects, in order to speed the convergence of the Markov Chain, hence fluxes are taken from SExtractor. We use the best "test orbit" from the automatic process as the starting point to define the section of the images where the model and residuals will be computed. This allows the inclusion of all images in the pointing, regardless of whether SExtractor found the object in every image or of any error in the position of the detections. The utility and success of MCMC is greatly increased if we are able to find a transfor- mation to a set of orthogonal variables, where the effect on the target likelihood produced by a small change in one dimension is decoupled from changes in others. Using Keplerian elements is not the most appropriate choice of variables since a small change in one element would affect others if the position of the object at a given time is to remain constant, making our method very inefficient. For this part of the analysis we considered the parameteriza- tion and routines developed by Bernstein & Khushalani (2000). These consider a cartesian coordinate system centered on the observer that points toward the center of the first im- -- 15 -- age. Though its variables are fairly independent when describing the parallactic motion of a TNO, we took into consideration a few modifications to the parameters. These changes of variable were chosen to ensure a smooth transition between different areas of the parameter space that yield similar trajectories. Since a change in distance also changes the travel time we were forced to include a shift in the object's relative position in every step so that the Markov Chain would not get stuck updating all other parameters every time the distance changed. The only constraint we imposed on trial orbits was that they were bound and that the velocity along the line of sight was zero, a good approximation given the short arc and the large distance to these objects. In Figure 2 we show the postage stamps around one of the objects hst11 for an iteration in the Markov chain. 6.3. Binarity In our sample of 14, there is only one object that is readily recognized as binary (See Fig. 3). The separation between the components is δα = 0.(cid:48)(cid:48)53 ± 0.05 and their magnitudes are: 23.6 ± 0.3 and 23.7 ± 0.3 respectively, making this a very likely equal-mass binary. The characteristics of this binary are quite common among the binary TNO population. Its absolute solar system magnitude H ∼ 6.8 (a proxy for size) and inclination (i ∼ 3.5) place it among many other binaries in (Noll et al. 2008, Figure 2). Fig. 3 shows an obvious binary, however limits on the binary fraction are difficult to obtain, given that we did not calibrate our search for binary detection. No binary control population was implanted and for this reason we are unaware of our efficiency at detecting them as a function of separation, brightness ratio or orbit. Nevertheless, having one detection we can only place a 7% lower limit on the fraction of wide, equal brightness binaries among the faint TNO population. 6.4. Size distribution Once the distance and magnitude are measured we can compute the size of each body, assuming a value for the albedo. If we further assume all objects are roughly at the same distance, the luminosity function can be written as a function of size, as shown in Figure 4 where we assume the distance d = 42 AU and the albedo p = 0.07 (Stansberry et al. 2008). Introduced in Bernstein et al. (2004), the Double Power Law (DPL) is a handy functional -- 16 -- form for the density of objects as a function of R magnitude that considers a break in the size distribution. The parameters are σ23 or the surface density of objects with R = 23, α1 and α2 or the slopes of the power law behavior of the luminosity function for the brightest and smallest objects, and Req is the magnitude where the behavior changes from that of small to that of large sizes. The previous best fit to the cumulative size distribution, that combined the results of all surveys listed in Fuentes et al. (2009) is shown in Figure 4 in red. We also considered our 14 objects together with the many surveys that provide detailed information about their calibration (See Table 2 in Fuentes & Holman (2008) and Fuentes et al. (2009); Fraser & Kavelaars (2009).) The total area surveyed or effective area in all those surveys is plotted in the top panel of Figure 5. We consider only objects that were discovered at magnitudes brighter than the magnitude at which their respective surveys' detection efficiency fell below 15% of the maximum efficiency. Using the orbital information provided in those surveys, we define the hot and cold populations as those with inclinations larger and smaller than 5◦. A caveat about some of the surveys that provide inclination and distance information with a ∼ 24-hr arc at opposition is that they can only compute a rate of motion on the sky. Fuentes et al. (2009) included only the rate of motion for every object, from which we computed the distance and inclination. We analyzed the likelihood function for all these observations given the effective surveyed area following the MCMC analysis described in Fuentes & Holman (2008). The likelihood function is plot against two of the DPL variables in each of the panels in Figure 8 for the hot, cold and all objects (top, middle and lower panel, respectively). The constraints on the DPL parameters, computed on the likelihood of each parameter marginalized over all others is: α1 = 0.89 ± 0.10, α2 = 0.29 ± 0.06, Σ23 = 1.61 ± 0.11, Req = 23.8 ± 0.3 for all objects considered, α1 = 0.70 ± 0.10, α2 = 0.30 ± 0.07, Σ23 = 0.93 ± 0.03, Req = 24.1 ± 0.7 for hot objects and α1 = 0.80 ± 0.08, α2 = 0.21 ± 0.09, Σ23 = 0.92 ± 0.02, Req = 24.2 ± 0.4 for cold ones. 7. Discussion Our search can be compared to the targeted use of HST by Bernstein et al. (2004) where 6 fields where imaged ∼ 20 times each to find the faintest TNOs possible. Though that search was significantly more sensitive to faint objects due to a careful selection of the fields, the coaddition of signal and the use of a wide-filter, it only focused on six independent -- 17 -- fields. That group found 3 objects, two of which remain the faintest TNOs ever imaged. This work would have been able to detect the two brightest of those objects in any pointing that satisfied our criteria. A more targeted survey, with better selection of filters and pointings at the stationary point, where objects won't trail as much, would be much more efficient at finding TNOs than our archival search. Of the total area imaged, 0.45 deg2 corresponding to the 150 pointings that were ana- lyzed, the effective area for this survey is only 0.3 deg2, as shown in the top panel of Figure 4. This is due to background sources, cosmic ray confusion, and any other feature that would completely or partially prevent us from detecting an object in at least three images. The main cause for this reduced survey area is, however, the existence of pointings of the same field taken is succession, which effectively increases the chances of detecting an object in that field but decreases the area of our survey by re-observing a field, where the same objects are visible, many times. The luminosity function of the 14 objects discovered in this survey is presented in the lower panel in Figure 4. The effective area is also plotted in the top panel to put the significance of each detection in context. The previous best fit to the TNO population is plot in green, the best fit for all surveys (including this ones) is also plotted in red. As we can see our survey follows precisely the expectations derived from previous work. There are two bright objects R < 23 among the 14, which indicates a higher density than expected. The statistical significance of this deviation is low. We took most calibrated surveys for TNOs in the literature (See §6.4), along with this one, to construct a effective survey area and luminosity function for all these objects, shown in Figure 5. There are over 400 TNOs included, which allows a precise constraint on the luminosity function. The two faintest objects beyond R = 27.5 were discovered with HST (Bernstein et al. 2004), and we see that ground based surveys are already sensitive to R = 27 TNOs. The exquisite astrometric precision in HST data enables us to measure a TNO's distance and inclination with a few percent uncertainty, even from an arc as short as 40 min. Although an object's motion depends on the solar elongation at which the observations were taken, the extra motion due to HST's orbit helps disentangle the objects' parallax from its proper motion along the ecliptic. Though this is a sample of only 14 objects, the distances found are in good agreement with Kavelaars et al. (2008). Dynamically cold objects are constrained to 35¡d¡50 AU, while hot objects do not cluster at a particular distance (See Table 2). Using the inclination information in Table 2 we separate objects into the hot and cold dynamical classes, for this we use a simple i = 5◦ cutoff. Applying this filter to all surveys we -- 18 -- compute the luminosity function of hot and cold objects (shown in Figure 6 and Figure 7 re- spectively.) We see that the bright-end and faint-end slopes for the hot and cold populations are similar. This is best seen in the MCMC posterior probability for the luminosity function DPL parameters (See Figure 8.) The hot and cold luminosity function constraints on α2, the faint-end slope, are consistent with each other. However, there is a significant deviation at the bright end of the dynamical hot population between the best fit model and the luminos- ity function which indicates that the size distribution of bright hot objects is shallower than that of the cold fraction, the largest and brightest objects tend to be dynamically excited, consistent with the results of Levison & Stern (2001); Bernstein et al. (2004). Our lack of data to contrain the bright end of the TNO luminosity function is explained as most of the objects in our analysis come from well characterized pencil beam surveys where only a few large objects are present. For the same reason, our constraints are more significant for the faint end of the luminosity function. This result is in contradiction with claims that the luminosity function of hot and cold objects differs for small bodies (Bernstein et al. 2004; Fuentes & Holman 2008). However, there is an explanation for this difference. For a long time the only TNOs fainter than R ∼ 26 were those found by Bernstein et al. (2004), all of them cold. This lack of faint hot objects allowed for extremely flat slopes for smaller sizes. By including the deeper surveys of Fuentes et al. (2009) and Fraser & Kavelaars (2009), which have detected several high inclination objects, the non-detection of R > 27 hot objects is less significant. However, these deep ground based surveys rely on short arcs and must constrain their orbits to be circular to compute a distance and inclination. This increases the probability of contamination between hot and cold objects. This problem is less severe with detections obtained from HST. In our survey we see roughly equal number of cold and hot objects, which is consistent with the ground based inclinations being accurate. As was discussed earlier, the size distribution is intimately related to the luminosity function. If we disregard the distance estimate and assume all objects are located at 42 AU and have a 7% albedo the transformation is direct and corresponds to the top axis in Figure 5. The break magnitude that marks the transition between the bright and the faint slope luminosity function becomes then a break in the size distribution. Such a break is expected from the collisional evolution the TNO population has undergone since these objects formed (Kenyon & Bromley 2004; Kenyon et al. 2008; Pan & Sari 2005). We find the location of such break to be consistent for both hot and cold populations Deq ∼ 100 km . In the model of Pan & Sari (2005) such a large size corresponds to the largest object that has been disrupted in the age of the Solar System. We note that this results relies on an assumed distance and albedo for all objects, something that we know is inaccurate for distances. The albedo is -- 19 -- also likely to be different as there seems to be a correlation with size and color (Stansberry et al. 2008). The results discussed in this section constrain the TNO size distribution at low ecliptic latitudes. As we cover more of the sky, sampling the TNO population away from the ecliptic we will measure density of objects as a function of latitude, as well as the relative proportion of hot vs. cold objects. We shall be able to include more objects in our analysis and hence put better constraints on the location of the break in the size distribution, for different dynamical families. Among the objects found there is a wide-separation binary. Since we did not consider the possibility of binaries in our control population, we cannot directly extract the statistical significance of our measurement. However, our result allows us to set a limit on the binary fraction of 7+13−2 %, in excellent agreement with the current limits for different population of 5-20% Noll et al. (2008). For the couple of objects with two different filter observations we put them in context of the Hainaut & Delsanti (2002) database, as shown in Figure 10. The two objects seem to fall right in what is expected for classicals, which is consistent with their inclination estimates: i ∼ 10◦ and 5◦. For at least these two objects the assumption V-R=0.6 is justified. The objects presented in this paper typically cannot be followed up. The short ob- servation arc, and the long time elapsed since the date the observations were taken imply an uncertainty in the position too large to recover them in current observations. However, the orbital estimates are accurate enough to grant an uncertainty ellipse that fits within an ACS/WFC fields within a month after the discovery observation (See Fig. 9.) This opens the possibility of a slightly different observing strategy, where data are processed as they become available, and follow-up observations can be scheduled quickly. 8. Conclusions We have successfully completed a search for TNOs within 5 degrees of the ecliptic using archival data taken with the ACS/WFC camera aboard HST. The data span 6 years. Of the 150 pointings analyzed 14 objects were found, yielding roughly 1 object per 10 pointings. This suggests that there are possibly hundreds of new TNOs with exquisite astrometry and photometry still hidden in the ACS/WFC archive at higher ecliptic latitudes. We have proven our ability to detect and characterize these even with data intended for completely different purposes, where most of the filters and strategies used for these observations are sub-optimal for detecting TNOs. -- 20 -- Given the excellent astrometric precision of the images, it is possible to use observations taken in a single pointing of HST to predict the position of a TNO a month later and have the uncertainty ellipse fit within the field of view of ACS. This, coupled to the fast turnover of data from our pipeline may yield many viable candidates for follow-up observations. A detection a month later would allow us to collect a significant set of small TNOs with accurate orbits, opening the possibility for detailed observations with present and future instruments like JWST. Binaries are only detected as such if the separation of the components on the plane of the sky can be resolved. The trailing of the binary imposes further constraints to the fraction of time a given binary would be recognized, which is specially problematic for HST data where trailing of TNOs is more prominent. Previous searches for binaries using HST have surveyed known TNOs and tracked the telescope to counteract their motion. The one relatively faint (R ∼ 23) binary discovered in this project illustrates the successful detection of a trailed binary. This opens the possibility of constraining the rate of binaries as a function of size from the same survey in which the TNOs are discovered. The recognition of a low and high inclination population among TNOs (Brown 2001) has been interpreted as the existence of two dynamically distinct set of objects. Those with higher inclinations or excited (hot) and those with lower inclinations and not excited (cold). Bernstein et al. (2004) found that these two, defined by their inclination had different size distributions. There is also evidence that they have different colors Doressoundiram et al. (2008). The cold population's size distribution is steeper than that of the hot population for objects larger than ∼ 100 km Levison & Stern (2001). In this paper we show that the location of the break in the size distribution and its slope for objects fainter than it for cold and hot objects are consistent with each other. This is compatible with the theory of collisional evolution, where a population with a given steep power law size distribution gets collisionally grinded as time progresses (Pan & Sari 2005; Kenyon & Bromley 2004; Kenyon et al. 2008). The slope of the size distribution of small objects in those models is constant since it is given by the steady state of collisions. The location of the break is also consistent for both populations, but Deq ∼ 100 km is larger than what theories expect. This difference could be better understood if any albedo dependence with size is first investigated for smaller objects. At present we can only extrapolate the apparent correlations between size and color, and inclination and albedo from measurements performed only on the largest, brightest TNOs. By chance, multifilter observations of two TNOs were obtained, yielding V-I colors for two objects. We expect to find more of these serendipitous color observations for other faint TNOs in intensively surveyed fields away from the ecliptic. -- 21 -- The advent of the Wide Field Camera 3 (WFC3) installed in May 2009 opens the possibility for extending this work to the Near-IR and to a larger fraction of the data collected by HST. The prospects of such observations would allow extending surface studies to small objects. As we continue the analysis of HST archival data to higher ecliptic latitudes we will start sampling an area of the sky that has only been surveyed for brighter (R (cid:46) 21 (Trujillo & Brown 2003). When the whole archive is searched we shall take the depth and resolution of pencil beam searches to the whole sky. Support for program 11778 was provided by NASA through a grant from the Space Telescope Science Institute, which is operated by the Association of Universities for Research in Astronomy, Inc., under NASA contract NAS 5-26555. REFERENCES Allen, R. L., Bernstein, G. M., & Malhotra, R. 2002, AJ, 124, 2949 Anderson, J., & King, I. R. 2000, PASP, 112, 1360 Bernstein, G., & Khushalani, B. 2000, AJ, 120, 3323 Bernstein, G. M., Trilling, D. E., Allen, R. L., Brown, M. E., Holman, M., & Malhotra, R. 2004, AJ, 128, 1364 Bertin, E., & Arnouts, S. 1996, A&AS, 117, 393 Brown, M. E. 2001, AJ, 121, 2804 -- . 2008, The Largest Kuiper Belt Objects, ed. Barucci, M. A., Boehnhardt, H., Cruikshank, D. P., & Morbidelli, A., 335 -- 344 Brucker, M. J., Grundy, W. M., Stansberry, J. A., Spencer, J. R., Sheppard, S. S., Chiang, E. I., & Buie, M. W. 2009, Icarus, 201, 284 Chiang, E. I., & Brown, M. E. 1999, AJ, 118, 1411 Doressoundiram, A., Boehnhardt, H., Tegler, S. C., & Trujillo, C. 2008, Color Properties and Trends of the Transneptunian Objects, ed. M. A. Barucci, H. Boehnhardt, D. P. Cruiksank, & A. Morbidelli, 91 -- 104 -- 22 -- Fraser, W. C., & Kavelaars, J. J. 2009, AJ, 137, 72 Fraser, W. C., et al. 2008, Icarus, 195, 827 Fuentes, C. I., George, M. R., & Holman, M. J. 2009, ApJ, 696, 91 Fuentes, C. I., & Holman, M. J. 2008, AJ, 136, 83 Gladman, B., Kavelaars, J. J., Petit, J.-M., Morbidelli, A., Holman, M. J., & Loredo, T. 2001, AJ, 122, 1051 Hainaut, O. R., & Delsanti, A. C. 2002, A&A, 389, 641 Jordi, K., Grebel, E. K., & Ammon, K. 2006, A&A, 460, 339 Kavelaars, J., Jones, L., Gladman, B., Parker, J. W., & Petit, J. 2008, The Orbital and Spa- tial Distribution of the Kuiper Belt, ed. Barucci, M. A., Boehnhardt, H., Cruikshank, D. P., & Morbidelli, A., 59 -- 69 Kenyon, S. J., & Bromley, B. C. 2004, AJ, 128, 1916 Kenyon, S. J., Bromley, B. C., O'Brien, D. P., & Davis, D. R. 2008, Formation and Collisional Evolution of Kuiper Belt Objects (The Solar System Beyond Neptune), 293 -- 313 Levison, H. F., & Stern, S. A. 2001, AJ, 121, 1730 Morbidelli, A., Levison, H. F., & Gomes, R. 2008, The Dynamical Structure of the Kuiper Belt and Its Primordial Origin, ed. Barucci, M. A., Boehnhardt, H., Cruikshank, D. P., & Morbidelli, A., 275 -- 292 Noll, K. S., Grundy, W. M., Chiang, E. I., Margot, J., & Kern, S. D. 2008, Binaries in the Kuiper Belt, ed. Barucci, M. A., Boehnhardt, H., Cruikshank, D. P., & Morbidelli, A. , 345 -- 363 Pan, M., & Sari, R. 2005, Icarus, 173, 342 Petit, J.-M., Holman, M. J., Gladman, B. J., Kavelaars, J. J., Scholl, H., & Loredo, T. J. 2006, MNRAS, 365, 429 Sirianni, M., et al. 2005, PASP, 117, 1049 Stansberry, J., Grundy, W., Brown, M., Cruikshank, D., Spencer, J., Trilling, D., & Margot, J.-L. 2008, Physical Properties of Kuiper Belt and Centaur Objects: Constraints from the Spitzer Space Telescope (The Solar System Beyond Neptune), 161 -- 179 -- 23 -- Trujillo, C. A., & Brown, M. E. 2003, Earth Moon and Planets, 92, 99 This preprint was prepared with the AAS LATEX macros v5.2. -- 24 -- Table 1. Photometric conversion F ilter Description zeropoint R − F ilter F435W Johnson B F475W SDSS g' F555W Johnson V F606W Broad V F625W SDSS r' F775W SDSS i' F814W Broad I F850LP SDSS z' 25.17 25.77 25.69 26.67 26.23 26.42 26.80 25.95 −1.02 −0.54 −0.66 −0.61 −1.03 −0.65 −0.69 +0.32 Note. -- HST filter name, equivalent standard name, and their respective zeropoint. The transformation to R assumes TNO colors (V-R = 0.6; R-I = 0.6; B-R = 1.6.) -- 25 -- Fig. 2. -- Images around the location of the found object hst11 in each of the three images in the pointing where it was found. Each row shows the data after cosmic ray processing, the model and the residuals for an acceptable trial sampled during the MCMC minimization of the residuals. Note that we did not fit for the fluxes but took them from SExtractor. -- 26 -- Fig. 3. -- A postage stamp sequence of cosmic ray corrected images around the position of hst5. No distortion correction has been applied to these images. The detections that were linked by the search algorithm are shown as magenta circles. The component closer to the background galaxy has a F814W magnitude of 23.6 ± 0.3 and the other one 23.7 ± 0.3. The separation is δα = 0.(cid:48)(cid:48)53 ± 0.01, which at a distance of 42.9 ± 0.6 AU gives a lower limit to their physical separation a > 165, 000 ± 2, 000 km . -- 27 -- Fig. 4. -- The top panel shows the effective area surveyed in this paper as a function of R magnitude in blue. While that function depends on the efficiency function and shared area of every pointing in this survey we can simplify the ∼ 2, 000 parameters (See Equation (3)) into the 4-parameters of the function plot in dashed orange (See Equation (5)). The lower panel shows the luminosity function of objects found in this survey, normalized by the effective area at each magnitude. The best model in Fuentes et al. (2009) is overplotted in green, while the best model for all surveys, including this one, is shown in red. The gray area represents the area enclosed by the 1 − σ confidence region for all surveys. The lower set of shaded areas represent the 1 − σ confidence limits for the cumulative function of the hot (red) and cold (yellow) population. -- 28 -- Fig. 5. -- The top panel shows in blue the effective area for all surveys considered in this paper as a function of R magnitude, normalized by the effective area at each magnitude. The lower panel shows the luminosity function of TNOs in all surveys considered, normalized by the effective area at each magnitude. The best model in Fuentes et al. (2009) is overplotted in green, while the best model for all surveys, including this one, is shown in red. The gray area corresponds to the 1 − σ confidence region given for all objects. The same confidence regions are given for hot and cold objects, in red and yellow respectively. -- 29 -- Fig. 6. -- Same as Figure 5 but only for objects deemed dynamically cold, i ≤ 5◦ -- 30 -- Fig. 7. -- Same as Figure 5 but only for objects deemed dynamically hot, i > 5◦ -- 31 -- Fig. 8. -- Probability density for the surface number density σ(R). The parameters are those in the double power law model. In each panel the painted areas represent the 3, 2 and 1-σ confidence regions. The panels on the left show the probability distribution as a function of the power law exponents for bright and faint objects (α1, α2). Panels on the right axis show the likelihood of the brightness at which the luminosity function changes slope, (Req) and the density of objects at R = 23 (Σ23). The red crosses show the 1 − σ confidence region for each parameter when the probability density has been marginalized over all other variables. The details of this likelihood analysis can be found in Fuentes & Holman (2008) and references therein. The bottom panel shows the results for all objects in all surveys considered, the most likely value for the parameters is: α1 = 0.89 ± 0.10, α2 = 0.29 ± 0.06, Σ23 = 1.61 ± 0.11, Req = 23.8 ± 0.3. The middle panel shows only objects considered as hot or excited, selected for having i > 5◦, the best parameters are: α1 = 0.70 ± 0.10, α2 = 0.30 ± 0.07, Σ23 = 0.93 ± 0.03, Req = 24.1 ± 0.7. The top panel corresponds to cold objects (i ≤ 5◦), where the most likely solution is α1 = 0.80 ± 0.08, α2 = 0.21 ± 0.09, Σ23 = 0.92 ± 0.02, Req = 24.2 ± 0.4. -- 32 -- Fig. 9. -- 1-sigma (green) and 99% (red) uncertainty in the position of a particular ob- ject (120-deg solar elongation), 30-days after discovery. The field of view of ACS/WFC is overplotted as a black square. -- 33 -- Fig. 10. -- Colors for our two objects along with previously known TNOs in the mboss database (Hainaut & Delsanti 2002) (red: classicals and plutinos; blue: scattered and cen- taurs; yellow: trojans; purple: long and short period comets.) The black points correspond to hst13 and hst17 respectively, where the V-R color for both is assumed to be 0.6 and the uncertainties are the same as those in V-I. -- 34 -- b i ] g e d [ b d ] U A [ 4 . 0 ± 9 . 2 2 . 0 ± 5 . 3 5 . 4 ± 9 . 6 4 . 3 1 ± 7 . 2 6 9 . 1 1 ± 7 . 4 1 4 . 6 1 ± 4 . 8 1 4 . 1 2 ± 6 . 8 2 9 . 6 ± 5 . 5 0 1 5 . 0 ± 9 . 4 0 . 4 6 ± 1 . 3 9 1 . 1 ± 4 . 8 0 . 2 ± 9 . 2 7 1 5 . 1 ± 6 . 3 3 3 . 0 ± 5 . 3 0 . 1 ± 6 . 2 3 . 6 ± 3 . 3 7 1 8 . 5 ± 4 . 9 4 . 1 ± 5 . 6 3 6 . 0 ± 9 . 2 4 5 . 9 ± 5 . 1 4 9 . 1 ± 8 . 3 5 3 . 3 3 ± 1 . 9 6 1 . 7 ± 2 . 8 2 5 . 1 ± 0 . 9 7 3 . 3 ± 5 . 5 8 5 . 2 ± 7 . 4 3 5 . 4 ± 9 . 1 3 1 9 . 1 ± 0 . 4 4 3 . 3 ± 2 . 6 5 0 . 1 ± 1 . 5 3 7 . 1 ± 6 . 3 5 4 . 1 ± 5 . 8 4 2 . 4 ± 6 . 7 5 5 . 2 ± 6 . 4 4 s t c e j b O d n u o F . 2 e l b a T g n a p p o H R a m a r e t l i F c e D A R D J M e m a N ] g e d [ 5 . 0 6 9 . 0 6 1 1 . 3 7 8 . 2 6 1 6 . 2 9 2 . 4 8 9 . 0 6 1 8 . 8 4 1 0 . 8 5 1 0 . 8 5 1 0 . 5 7 5 . 7 7 7 . 3 4 1 3 . 5 1 1 0 . 0 1 8 . 6 3 . 0 1 0 . 9 3 . 8 5 . 1 1 5 . 6 2 . 6 8 . 7 3 . 5 5 . 9 4 . 8 4 . 0 1 5 . 7 6 . 8 9 . 7 2 . 9 1 . 0 ± 0 . 5 2 3 . 0 ± 5 . 2 2 3 . 0 ± 9 . 5 2 8 . 0 ± 7 . 5 2 4 . 0 ± 1 . 6 2 2 . 0 ± 4 . 5 2 3 . 0 ± 9 . 4 2 1 . 0 ± 6 . 2 2 2 . 0 ± 9 . 5 2 1 . 0 ± 3 . 5 2 4 . 0 ± 3 . 5 2 1 . 0 ± 2 . 4 2 2 . 0 ± 9 . 4 2 7 . 5 2 1 . 3 2 6 . 6 2 4 . 5 2 7 . 6 2 1 . 6 2 9 . 5 2 6 . 3 2 6 . 6 2 W 5 7 7 F W 4 1 8 F W 4 1 8 F P L 0 5 8 F W 5 5 5 F W 4 1 8 F W 5 2 6 F W 5 3 4 F W 4 1 8 F 9 . 5 2 : 9 . 5 2 W 4 1 8 F W : 6 0 6 F 9 . 5 2 3 . 5 2 6 . 5 2 W 4 1 8 F W 5 2 6 F W 4 1 8 F 0 1 7 . 6 4 : 9 5 : 5 9 0 . 1 1 : 4 5 : 6 8 1 . 9 0 : 2 1 3 1 − 7 − : 3 7 0 6 . 4 3 2 9 9 . 4 3 8 8 2 . 7 2 : : : 2 5 2 5 2 5 : : : 6 1 6 1 6 1 2 1 9 . 5 1 : 2 5 : 1 1 4 7 . 2 0 0 6 3 . 1 0 7 2 5 . 6 4 0 1 1 . 7 2 : : : : 6 0 5 1 8 3 9 2 : : : : 8 6 5 . 6 4 9 5 5 . 6 4 6 1 7 . 6 4 : : : 1 2 1 2 1 2 1 1 2 2 2 2 2 1 : : : 3 3 3 5 7 7 0 7 4 1 . 8 1 3 3 5 8 7 1 0 4 6 . 5 8 5 3 5 5 3 0 1 1 3 . 7 6 8 3 5 5 7 7 8 4 3 4 . 8 3 8 3 5 8 5 5 5 4 1 9 . 4 6 9 3 5 8 5 4 0 6 0 . 6 5 9 3 5 3 3 8 8 9 . 4 3 0 4 5 7 0 6 . 3 4 9 8 7 . 1 1 3 9 1 . 0 4 : : : 8 4 3 5 1 5 : : : 2 0 0 9 1 3 . 0 5 : 3 1 : 2 1 7 8 2 9 0 7 0 . 6 8 7 3 5 4 6 8 . 8 1 5 8 3 . 4 3 : : 9 3 9 3 : : 0 0 6 4 5 1 2 5 5 . 9 8 9 3 5 6 2 5 3 1 2 . 8 8 9 3 5 6 2 2 . 3 2 : 3 3 : 1 8 1 . 1 0 : 2 2 7 8 9 . 5 2 : 6 0 : : 9 − 0 1 2 1 − 2 3 5 . 5 0 8 1 3 . 5 1 : : 5 5 4 5 : : 1 2 3 1 3 9 4 . 3 0 : 7 4 : 9 9 7 2 9 3 1 . 2 4 0 4 5 1 4 7 7 3 8 . 4 6 7 2 5 2 6 7 6 2 7 7 . 2 2 4 2 5 c 5 t s h 4 t s h 6 t s h 7 t s h 8 t s h 9 t s h d 0 1 t s h 1 1 t s h 2 1 t s h d 3 1 t s h 4 1 t s h 5 1 t s h d 6 1 t s h 1 . 0 ± 1 . 5 2 8 . 5 2 : 8 . 5 2 W 4 1 8 F W : 5 5 5 F 2 8 2 . 0 4 : 0 0 : 4 1 − 8 8 1 . 3 5 : 4 1 : 2 2 4 3 0 7 3 . 1 4 9 2 5 e 7 1 t s h e h T . s n o i t c e t e d t s r fi e h t r o f n e v i g s n o i t i s o P . s e i t r e p o r p c i r t e m o r t s a d n a c i r t e m o t o h p r i e h t h t i w n w o h s e r a k r o w s i h t n i d n u o f s t c e j b o l l A -- . e t o N . ) 0 0 0 2 i n a l a h s u h K & n i e t s n r e B ( e d o c t fi b r O e h t y b n e v i g n o i t a z i r e t e m a r a p a h t i w C M C M a m o r f d e t a m i t s e e r e w i n o i t a n i l c n i d n a d e c n a t s i d c i r t n e c y r a b n o i t c n u f a , ∆ d g o l 5 + V = H e d u t i n g a m m e t s y S r a l o S e h T . t n e r e ff d i s i s n o i t a v r e s b o e h t f o h c o p e e h t , d l e fi e m a s e h t n i d e r e v o c s i d e r e w s t c e j b o e m o s h g u o h T . 6 . 0 s i s t c e j b o l l a r o f r o l o c R V e h t - t a h t d n a l l a m s s i e l g n a e s a h p e h t i g n m u s s a d e t u p m o c s i , ∆ r e v r e s b o e h t o t e c n a t s i d e h t d n a d e d u t i n g a m V e h t f o d n o c e s e h t f o s t l u s e r e h t e d u l c n i t o n o d e w h g u o h t , s t i b r o h t o b g n i t t fi n o i t u l o s e m a s e h t d n u o f e W . s g n i t n i o p e v i t u c e s n o c o w t n i d n u o f s a w t c e j b o . e n o e d a r g o r p e h t n a h t r e v r e s b o e h t m o r f e c n a t s i d r e g r a l a t a s y a w l a s i t i ; w o r d n o c e s a n i d e t n e s e r p s i n o i t u l o s e d a r g o r t e r . 3 . g i F n i n w o h s y r a n b i e h t s i s i h T c e h T d s i h T e . r o t c e t e d e h t f o e g d e e h t n o t h g i r s i s n o i t c e t e d e h t f o e n o s a g n i t n i o p . n w o h s e r a s e d u t i n g a m l a t n e m u r t s n i i g n d n o p s e r r o c d n a t s i l r e t l fi a , d e s u e r a s r e t l fi e l p i t l u m n e h W . s k a e p h t o b t r o p e r e w e l b i s s o p e r a s n o i t u l o s e d a r g o r t e r d n a e d a r g o r p n e h W a b
1604.03511
2
1604
2016-04-21T01:31:58
Directed Energy Missions for Planetary Defense
[ "astro-ph.EP", "astro-ph.IM", "physics.pop-ph" ]
Directed energy for planetary defense is now a viable option and is superior in many ways to other proposed technologies, being able to defend the Earth against all known threats. This paper presents basic ideas behind a directed energy planetary defense system that utilizes laser ablation of an asteroid to impart a deflecting force on the target. A conceptual philosophy called DE-STAR, which stands for Directed Energy System for Targeting of Asteroids and exploRation, is an orbiting stand-off system, which has been described in other papers. This paper describes a smaller, stand-on system known as DE-STARLITE as a reduced-scale version of DE-STAR. Both share the same basic heritage of a directed energy array that heats the surface of the target to the point of high surface vapor pressure that causes significant mass ejection thus forming an ejection plume of material from the target that acts as a rocket to deflect the object. This is generally classified as laser ablation. DE-STARLITE uses conventional propellant for launch to LEO and then ion engines to propel the spacecraft from LEO to the near-Earth asteroid (NEA). During laser ablation, the asteroid itself provides the propellant source material; thus a very modest spacecraft can deflect an asteroid much larger than would be possible with a system of similar mission mass using ion beam deflection (IBD) or a gravity tractor. DE- STARLITE is capable of deflecting an Apophis-class (325 m diameter) asteroid with a 1- to 15-year targeting time (laser on time) depending on the system design. The mission fits within the rough mission parameters of the Asteroid Redirect Mission (ARM) program in terms of mass and size. DE-STARLITE also has much greater capability for planetary defense than current proposals and is readily scalable to match the threat. It can deflect all known threats with sufficient warning.
astro-ph.EP
astro-ph
To be published in: Advances in Space Research (ASR) Journal - Elsevier 2016 Submitted November 3, 2014 Directed Energy Missions for Planetary Defense Philip Lubina, Gary B. Hughesb, Mike Eskenazic, Kelly Kosmoa, Isabella E. Johanssona , Janelle Griswolda, Mark Pryord, Hugh O'Neille, Peter Meinholda, Jonathon Suena, Jordan Rileya, Qicheng Zhanga, Kevin Walshf, Carl Melisg, Miikka Kangasa, Caio Mottaa, and Travis Brashearsa Corresponding author for questions: [email protected] aPhysics Department, University of California, Santa Barbara, CA 93106, USA bStatistics Department, California Polytechnic State University, San Luis Obispo, CA 93407, USA cATK Space, 600 Pine Avenue, Goleta, CA 93117, USA dVorticy Inc, San Diego, CA 92121, USA ePhysics Department, Ventura College, Ventura, CA 93003, USA fSouthwest Research Institute, Boulder, CO 80302, USA gCenter for Astrophysics and Space Sciences, UC San Diego, San Diego, CA 92093, USA For more information and related articles, videos and talks see: http://www.deepspace.ucsb.edu/projects/directed-energy-planetary-defense ABSTRACT Directed energy for planetary defense is now a viable option and is superior in many ways to other proposed technologies, being able to defend the Earth against all known threats. This paper presents basic ideas behind a directed energy planetary defense system that utilizes laser ablation of an asteroid to impart a deflecting force on the target. A conceptual philosophy called DE-STAR, which stands for Directed Energy System for Targeting of Asteroids and exploRation, is an orbiting stand-off system, which has been described in other papers. This paper describes a smaller, stand-on system known as DE-STARLITE as a reduced-scale version of DE-STAR. Both share the same basic heritage of a directed energy array that heats the surface of the target to the point of high surface vapor pressure that causes significant mass ejection thus forming an ejection plume of material from the target that acts as a rocket to deflect the object. This is generally classified as laser ablation. DE-STARLITE uses conventional propellant for launch to LEO and then ion engines to propel the spacecraft from LEO to the near-Earth asteroid (NEA). During laser ablation, the asteroid itself provides the propellant source material; thus a very modest spacecraft can deflect an asteroid much larger than would be possible with a system of similar mission mass using ion beam deflection (IBD) or a gravity tractor. DE- STARLITE is capable of deflecting an Apophis-class (325 m diameter) asteroid with a 1- to 15-year targeting time (laser on time) depending on the system design. The mission fits within the rough mission parameters of the Asteroid Redirect Mission (ARM) program in terms of mass and size. DE-STARLITE also has much greater capability for planetary defense than current proposals and is readily scalable to match the threat. It can deflect all known threats with sufficient warning. Keywords: Directed Energy, Laser Phased Array, Planetary Defense, DE-STAR, DE-STARLITE 1.1 DE-STAR and DE-STARLITE 1. INTRODUCTION While implementing a realistic directed energy planetary defense system may have seemed preposterous as little as a decade ago, recent technological developments allow serious consideration of such a system. The critical items such as phase locked laser amplifiers and lightweight photovoltaic deployable arrays are becoming increasingly more efficient and lower in mass. The necessary technology now exists to build such a system that will considerably enhance our ability to augment or enhance other methods to fulfill the need for planetary defense against asteroids that pose a threat of impacting Earth. This paper primarily focuses on a design for a stand-on directed energy planetary defense system called DE- STARLITE. DE-STARLITE is a stand-on system, i.e., it is designed to be delivered to a position that is nearby a threatening asteroid with a modest spacecraft and then work slowly on the threat to change its orbit. DE-STARLITE is suitable for mitigating targets that are many hundreds of meters in diameter and whose orbit is known to be a threat long before projected impact. DE-STARLITE is one component of a more far-reaching philosophy for directed energy planetary defense. A future orbiting system is envisioned for stand-off planetary defense. The conceptual system is called DE-STAR, for Directed Energy System for Targeting of Asteroids and exploRation. Fluctuations in the Earth's atmosphere significantly hinder ground-based directed energy systems; thus, deploying a directed energy system above Earth's atmosphere eliminates such disturbances, as the interplanetary medium is not substantial enough to significantly affect the coherent beam. DE-STAR is discussed extensively in other papers (Lubin and Hughes, 2015; Kosmo et al., 2015; Lubin et al., 2014). The broader DE-STAR system is not discussed in depth in this paper, which will focus on DE- STARLITE. 1.2 General Concepts for Orbit Deflection Residents near Chelyabinsk, Russia experienced the detrimental effects of a collision with a near-Earth asteroid (NEA) on 15 February 2013 as a ~20 m object penetrated the atmosphere above that city (Popova et al., 2013). The effective yield from this object was approximately 1/2 Megaton TNT equivalent (Mt), or that of a large strategic warhead. The 1908 Tunguska event, also over Russia, is estimated to have had a yield of approximately 15 Mt and had the potential to kill millions of people had it come down over a large city (Garshnek et al., 2000). Asteroid impacts pose a clear threat and future advancement to minimize this threat requires effective mitigation strategies. A wide array of concepts for asteroid deflection has been proposed. Several detailed surveys of threat mitigation strategies are available in the literature, including Sanchez-Quartielles et al. (2007), Belton et al. (2004), Gritzner and Kahle (2004), and Morrison et al. (2002). Currently proposed diversion strategies can be broadly generalized into six categories. (1a) Kinetic impactors, without explosive charges. An expendable spacecraft would be sent to intercept the threatening object. Direct impact could break the asteroid apart (Melosh and Ryan, 1997), and/or modify the object's orbit through momentum transfer. The energy of the impact could be enhanced via retrograde approach, e.g. McInnes (2004). (1b) Kinetic impactors, with explosive charges. Momentum transfer using an expendable spacecraft could also be enhanced using an explosive charge, such as a nuclear weapon, e.g. Koenig and Chyba (2007). (2) Gradual orbit deflection by surface albedo alteration. The albedo of an object could be changed using paint, e.g. Hyland et al. (2010). As the albedo is altered, a change in the object's Yarkovsky thermal drag would gradually shift the object's orbit. Similar approaches seek to create an artificial Yarkovsky effect, e.g. Vasile and Maddock (2010). (3) Ion beam deflection (IBD) or ion beam shepherd (IBS) where high speed ions, such as the type used for ion thrusters, are directed at the asteroid from a nearby spacecraft, to push on asteroid and thus deflect it. (Bombardelli et al., 2016; Brophy, 2015; Bombardelli, et al., 2013; Bombardelli and Peláez, 2011). (4) Direct motive force, such as by mounting a thruster directly to the object. Thrusters could include chemical propellants, solar or nuclear powered electric drives, or ion engines (Walker et al., 2005). (5) Indirect orbit alteration, such as gravity tractors. A spacecraft with sufficient mass would be positioned near the object, and maintain a fixed station with respect to the object using onboard propulsion. Gravitational attraction would tug the object toward the spacecraft, and gradually modify the object's orbit (Mazanek, et al., 2015; Wie, 2008; Wie, 2007; McInnes, 2007; Schweickart et al., 2006; Lu and Love, 2005). (6) Expulsion of surface material such as by robotic mining. A robot on the surface of an asteroid would repeatedly eject material from the asteroid. The reaction force when material is ejected affects the object's trajectory (Olds et al., 2007). (7) Vaporization of surface material. Like robotic mining, vaporization on the surface of an object continually ejects the vaporized material, creating a reactionary force that pushes the object into a new path. Vaporization can be accomplished by solar concentrators (Vasile and Maddock, 2010), lasers deployed from the ground (Phipps, 2010), or lasers deployed on spacecraft stationed near the asteroid (Maddock et al., 2007; Park and Mazenek, 2005; Gibbings et al., 2013; Phipps and Michaelis, 1994; Campbell, 2000; Vasile et al., 2013). One study (Kahle et al., 2006) envisioned a single large reflector mounted on a spacecraft traveling alongside an asteroid. The idea was expanded to a formation of spacecraft orbiting in the vicinity of the asteroid, each equipped with a smaller concentrator assembly capable of focusing solar power onto an asteroid at distances near ~1 km (Vasile and Maddock, 2010). 1.3 DE-STARLITE The DE-STARLITE mission design, which is detailed in this paper, utilizes the same technologies and laser system as the larger standoff directed energy system. Namely, DE-STAR is a modular phased array of lasers that heat the surface of potentially hazardous asteroids to approximately 3000 K, a temperature sufficient to vaporize all known constituent materials. Mass ejection due to vaporization causes a reactionary force large enough to alter the asteroid's orbital trajectory and thus mitigate the risk of impact. Each DE-STAR system is characterized by the log of its linear size (Lubin et al., 2014). DE-STARLITE is basically a DE-STAR 0, consisting of a laser phased array on the order of 1 m in diameter. DE-STARLITE utilizes deployable photovoltaic arrays to power the system. A conceptual design is illustrated in Fig. 1. Figure 1. Artistic rendering of a deployed DE-STARLITE spacecraft deflecting an asteroid. The spacecraft is outfitted with two 15 m diameter MegaFlex PV Arrays, a z-folded radiator deployed up and down, a laser array mounted on a gimbal at the front, and ion engines at the back. The laser array can be either a phased array or a parallel non phased array. A baseline mission includes 1 m total aperture, with a goal to produce a 10 cm spot from a distance of 10 km. This is an artistic rendering only, demonstrating the overall concept. Note that the optimal thrust vector for orbit deflection is generally parallel or anti-parallel. One advantage of this approach is the ability to target the asteroid from a significant distance, mitigating the effects of the ejecta on the spacecraft. In addition to the ion engines shown in the rear of the spacecraft, there are small ion engines on the sides of the spacecraft for station keeping and maneuvering toward or away from the asteroid. DE-STARLITE fits into the same basic launch vehicle and mass envelope as the current Asteroid Redirect Mission (ARM) block 1 program, which is designed to capture a 5-10 m diameter asteroid; however, DE-STARLITE is designed to be a true planetary defense system capable of redirecting large asteroids. It has been designed to use the same ion engines as the ARM program and the same PV system, though due to the reduced mass of DE-STARLITE, a much larger PV array can be deployed within the SLS block 1 mass allocation (70 tons to LEO) if desired. The scaling to megawatt class systems is discussed below. This paper will focus on a 100 kW (electrical) baseline DE-STARLITE as a feasible and fundable option that could pave the way for the ultimate long-term goal of a full standoff planetary defense system. Larger systems are also discussed. This paper details the design of the main elements of the spacecraft, namely, the photovoltaic panels, ion engines, laser array, and radiator as well as the parameters of the launch vehicles under consideration, and details the deflection capabilities of the system. 2. DESIGN The objective is to design a system that will enable a spacecraft with a 1 m to 4.5 m diameter laser phased array to arrive at an NEA (Near Earth Asteroid) and deflect it from its potentially hazardous trajectory. The laser phased array is detailed in section 2.3, along with a lower risk potential fallback-a close packed focal plane array of fiber lasers. The propulsion for the LEO to NEA portion of the DE-STARLITE mission is made possible with a high-power solar electric propulsion (SEP) system (Brophy and Muirhead, 2011). The solar PV arrays, detailed in Section 2.1, convert power from the sun to provide system power. PV panels will originally be stowed for launch and will deploy upon reaching low-Earth orbit to provide a required 100 kW electrical power from two 15 m diameter ATK MegaFlex panels. Even larger power is possible within the launch mass and shroud sizes available as discussed in detailed below. The system will utilize ion engines (detailed in section 2.2) to propel the spacecraft from LEO to an NEA, as proposed in JPL's ARM program. The system aims to stay within the same mass and launch constraints as ARM and use much of the same propulsion technology. The laser efficiency determines the laser power obtained from the PV arrays; 35 kW of laser power would be produced at 35% efficiency and 50 kW at 50%.. The 35 kW estimate is based on the current efficiency (35%) of existing technology of the baseline Ytterbium laser amplifiers and thus provides for the worst case, while the 50 kW estimate is based on near term technological improvement within the next 5 years. A passive cooling radiator with z-folded arrays will be used to reject waste heat and maintain the temperature at an operational 300 K. The basic design principle is to utilize a cylindrical bus with the lateral center of gravity close to the centerline (Kosmo et al., 2014). PV panels will be stowed at the back of the bus until deployment, and the hexagonal laser array will be mounted on a gimbal at the front of the spacecraft (Fig. 2). Radiator panels will deploy up and down (perpendicular to the bus) and will rotate about their axis so as to remain perpendicular to the sun in order to maximize radiator efficiency. Ion engines are located at the back of the spacecraft. Critical components are outlined in sections 2.1 through 2.4. Figure 2. Conceptual design of the deployed spacecraft with two 15 m PV arrays that produce 50 kW each at the beginning of life for a total of 100 kW electrical, ion engines at the back, and the laser array pointed directly at the viewer. A 2 m diameter laser phased array is shown with 19 elements, each of which is 1-3 kW optical output. A 2 m diameter optical system is one of the possibilities for DE-STARLITE. More elements are easily added to allow for scaling to larger power levels. A 1 - 4.5 m diameter is feasible; no additional deflection comes from the larger optic, just additional range from the target. 2.1 Photovoltaic Panels Two 15 m diameter MegaFlex PV arrays, manufactured by ATK Aerospace Systems in Goleta, CA, will be used to obtain the baselined 100 kW power solution. Extensive testing has been conducted on MegaFlex technology and the MegaFlex arrays have a high TRL (Murphy et al., 2014). Fig. 3 shows the PV array design and implementation. Figure 3. Detailed design of deployed MegaFlex array. Image courtesy ATK (Murphy et al., 2014). 2.2 Ion Engines The spacecraft carrying DE-STARLITE will utilize ion propulsion to spiral out from LEO to its target (Kosmo et al., 2015). Ion engines are proposed for DE-STARLITE because they are between five and ten times more efficient than engines using conventional chemical propellants, depending on the type of ion engine. 2.3 Laser Array The objective of the laser directed energy system is to project a large enough flux onto the surface of a near- Earth asteroid (via a highly focused coherent beam) to heat the surface to a temperature that exceeds the vaporization point of constituent materials, namely rock, as depicted in Fig. 4. This requires temperatures that depend on the material, but are typically around 2000-3000 K, or a flux in excess of 107 W/m2. A reactionary thrust due to mass ejection will divert the asteroid's trajectory (Lubin et al., 2014). To produce a great enough flux, the system must have both adequate beam convergence and sufficient power. From a distance of 10 km, a spot size on the asteroid of 10 cm provides enough flux to vaporize (sublimate) rock (Hughes et al., 2014). Optical aperture size, pointing control and jitter, and efficacy of adaptive optics techniques are several critical factors that affect beam convergence. As mentioned, the optical power output of the laser is projected to be between 35 kW and 70 kW, depending on technological advancements in laser amplifier efficiency in the coming years. Currently the amplifiers are about 35% efficient but it is expected they will exceed 50% within five years. Similar requirements are sought by power beaming systems (Mankins, 1997; Lin 2002). For the optional (non-phase-locked) fiber focal plane array the lasers are even more efficient and already exceed 50%. Any power level in this range will work for the purpose of this mission, but higher efficiency allows for more thrust on the target for a given electrical input as well as for smaller radiators and hence lower mission mass. 7 MegaFlex Space Power Workshop 2014 MegaFlex Design Overview - deployed Main Panels Composite Spar Gore Assembly Spar Latch Hub Assembly Stack Release Root Hinge and Latch Motor and Spool Gorelet Tiedowns 10-m MDU wing design leveraged CRS Sub-Assys have flight qualification TRL Extension Hinges (deployed) (stowed) Figure 4. LEFT: Physics based simulation of asteroid laser ablation (Lubin et al., 2014). MIDDLE: Simulation showing one spot from the baseline phased array on the target at sufficient temperature to cause ablation. RIGHT: Multi-beam simulation depicting 19 beams on the target from an optional choice of a close packed fiber laser focal plane array. The proposed baseline optical system consists of 19 individual optical elements in a phased array. A single element concept is shown in Fig. 5. A significant benefit of utilizing an array of phase-locked laser amplifiers is that it is completely modular and thus scalable to much larger systems, and allows for a greater range than would a close packed array with a single optic. Focusing and beam steering are achieved by controlling the relative phase of individual laser elements. Rough pointing of the array to the target is determined by spacecraft attitude control and gimbal pointing of the optics. Laser tips behind each optical element are mounted on 6-axis micro-positioner hexapod; lateral movement of the laser tips behind each lens provides intermediate pointing adjustment for individual array elements. Each fiber tip is supported on the hexapod and can be augmented with a z-axis rapid position controller if needed. It is not clear if this is needed currently. Precision beam steering is accomplished by coordinated phase modulation across the array by z- position control of the fiber tips as well as by electronic phase modulation. Each fiber is fed with a phase-controllable laser amplifier. Phase feedback from in front of the lens array to each phase controller provides a signal for beam formation adjustment (spot focus). Phase alignment is maintained to within lambda/10 1-sigma RMS across the entire array, assuming adequate phase controller system response (Hughes et al., 2014). Figure 5. Single element of laser phased array, showing fiber-tip actuator for mid-level pointing control and rough phase alignment. The DE-STARLITE spacecraft will be capable of outfitting a laser array with an aperture between 1 m and 4.5 m. Changing the aperture would not change the power of the laser; thus the solar arrays and radiator could be of the same dimensions, regardless. The difference in mass between the 1 m array and the 4.5 m array is not significant enough to pose new constraints. The benefit of a larger aperture is that the range of the laser from the target for a given power scales with the linear size of the optical aperture. For example, the range of a 3 m array is three times greater than that of a 1 m array. The benefit of having a longer range is that it allows the spacecraft to remain clear of the debris of the ejected material. The debris flux (kg/m2 s) that hits the spacecraft drops as the square of the distance to the target. However, the total amount of particle debris (kg/s) on the optic is independent of distance since the range is proportional to the optic diameter, and the area of the optic is proportional to the square of the diameter. The main drawback of the larger aperture is a higher associated cost. The decision, thus, is dependent on funding and other mission specifics. Even sub-meter diameter optics are feasible if needed for specific missions. The laser array will be placed on a gimbal to eliminate any potential issues with fuel usage in maneuvering the spacecraft, as depicted in Fig. 6. Further, it will allow for much greater flexibility in mission execution. This is imperative because the laser will have to raster scan the asteroid in order to maximize thrust, prevent burn through, and de-spin the asteroid if needed. Though much of this can be done with electronic steering, using a gimbal will be more energetically efficient than pointing the spacecraft. A gimbal would also be beneficial in the event that the spacecraft needs to orbit the target. Further, the added flexibility due to the gimbal mitigates risk by allowing the system to target smaller pieces of the asteroid that may get dislodged and pose a threat to the spacecraft. The gimbal will allow for two degrees of freedom because the angular orientation around the boresight of the spot on the asteroid is not a significant concern. This will be cheaper, easier to manufacture, and lighter than a system with greater degrees of freedom. (a) (b) (c) Figure 6. Mounted hexagonal laser phased array with a baseline of 19 elements depicted: (a) at 45 degrees, (b) face on, and (c) from the back. 2.4 Secondary Optical Arrangement If necessary, a fallback option is to implement a hexagonal close packed focal plane array of laser fibers with a conventional optic such as a reflecting telescope instead of a phased array. A conceptual diagram is shown in Fig. 7. This system would consist of 19 circular fibers, each 25 μm in diameter with a sheath (cladding) around the inner core. The cladding will be 37.5 μm thick so that the center-to-center spacing of adjacent laser fibers is 100 μm. The thickness of the cladding may be increased if power leakage and cross talk is an issue. As with the phased array design, each fiber is attached to an amplifier; however, the fibers are close packed in the focal plane and utilize one larger hexapod and the lasers are NOT phase locked for simplicity. The close packed array will produce 19 individual spots on the target, separated center to center by the ratio of the target distance to optical size times the fiber spacing in the focal plane. For the baseline of 1.5 kW per amplifier, each fiber will illuminate the target with a spot diameter of approximately 12 mm and a center to center spacing of approximately 50 mm; however, this can be changed depending on the optical design. This option carries a lower risk, higher initial TRL, lower cost, and can also be implemented more rapidly. In addition, it requires the spacecraft to be significantly closer to the target than would be required with a phased array. The plan is to pursue both the phased array and the close packed array, and down select depending on specific mission parameters. Figure 7. LEFT: Comparison view of mounted laser phased array and close packed array. RIGHT: Hexagonal close packed focal plane array of 19 laser fibers in the focal plane. Laser fibers have a diameter of 25 μm, the cladding around each of which is 37.5 μm thick. 2.5 Radiator Thermal radiators are essential to spacecraft design so as to minimize incident radiation and maintain the spacecraft and its components at a functional temperature. The efficiency of the radiator can be determined by equation 1: (1) where is the emittance of the surface, is the Stephan-Boltzmann constant, T is the temperature, is the heat rejected, A is the area, and F is the flux (Aaron, 2002). The baseline radiator will be coated in AZ-93 white paint, which has a high emittance of 0.91± 0.02 (or conservatively, 0.89) and a low alpha, as it only absorbs 14 -16% of incident sunlight on the spacecraft. The goal is to maintain a temperature of 300 K, as both the laser and onboard control electronics are operational at this temperature. At this temperature, the radiator can reject an idealized outward flux of 408 W/m2. When taking into account the incident radiation, using a solar constant of 1362 W/m2 and a maximum 16% absorptance, the net flux of energy across the surface of the radiator is approximately 190 W/m2. The baseline is to prevent direct solar illumination of the radiator. The area of the radiator must be determined by thermal analysis, and is dependent on the desired operating temperature, heating from the environment, interactions with other surfaces of the spacecraft (e.g., solar arrays), and the highest estimate (worst case) satellite waste heat. The waste heat in this case is dependent on the efficiency of the laser amplifiers-35% or50% as mentioned. The worst-case estimate (35% efficiency) requires 65 kW to be rejected as waste heat for a 100 kW electrical input assuming virtually all the power goes to the laser (which is approximately correct during laser firing). The required area A can be easily determined: (2) where Fnet is the net outward flux and is the heat rejected. Given these parameters, the maximum required area of the radiator is ~341 m2 for a 35% efficient laser amplifier. For a 50% efficient laser, a radiator area of ~262 m2 is required. We assume that either a pumped liquid cooling loop or an advanced heatpipe would be used to transfer the heat from the laser to the radiator as is currently done now in the other uses of these laser amplifiers. rejected = AFnet A passive cooling z-folded radiator consisting of two deployable panels will be used in order to provide a sufficient surface area over which to emit the waste heat generated by the system. Each panel z-folds out into six segments, each of which further folds out into two additional segments, making 18 segments in total for each panel. The panels will rotate about their axes to maximize efficiency by remaining perpendicular to the sun and by radiating out of both sides. Each segment will be 2.2 m by 2.2 m, yielding a total area of 348 m2. Note that the radiators radiate out of both sides and that there are two radiator panels. These values are approximate; a more detailed radiator design would be required as part of an overall mission design. We would expect that, by the time of any mission start, significant increases in laser efficiency will have been achieved, thus reducing the required radiator size. Sun shades may also be used to limit solar absorption and thus allow for greater efficiency. The current mass to power ratio for radiators is about 25 kg/kW for the ARM system. 3. LAUNCH SYSTEM The objective is to assess which launch vehicle is the most feasible and will provide the greatest performance given the mission directives of DE-STARLITE. The launch systems in consideration are Atlas V 551, Space Launch System (SLS) Block 1, Falcon Heavy, or Delta IV Heavy. These are likewise the launch systems in consideration for JPL's Asteroid Redirect Mission, which calls for a payload of comparable parameters (Brophy and Muirhead, 2013). Table 1. Parameters of various launch vehicles in consideration for DE-STARLITE Parameter Atlas V 551 SLS Block 1 Falcon Heavy Delta IV Heavy Payload Mass to LEO Cost per unit mass to send into LEO Diameter of Payload Fairing Status 18 500 kg $13 200 / kg 70 000 kg $18 700 / kg 53 000 kg $1 890 / kg 5.4 m Flight proven 8.4 m Development- First Expected Flight: 2017 5.2 m Development- First Expected flight: 2015 28 790 kg $13 000 / kg 5 m Flight proven The DE-STARLITE spacecraft will fit within the payload fairing of any of the proposed launch systems, as depicted in Fig. 8. As is evident from the data in Table I, the SLS Block 1 has the highest capabilities, though also requires the highest cost. The Falcon Heavy demands the smallest cost per unit mass, and has capabilities between that of the Atlas V and SLS Block 1. While the Atlas V 551 and Delta IV Heavy have previously undergone successful missions, the SLS Block 1 and Falcon Heavy are projected to be flight-proven within the timescale of the DE- STARLITE mission. As with the Asteroid Redirect Mission, it is possible to compensate for the lower capabilities of the Atlas V by using the SEP system to spiral out of Earth's orbit and escape from Earth using Lunar Gravity Assist (LGA); however, this process of spiraling out and using LGA will take an additional 1 – 1.5 years of flight. All of these factors must be taken into consideration to choose the most effective launch system for the DE-STARLITE mission. Figure 8. Stowed view of the DE-STARLITE spacecraft, in a 5 m fairing. 4. EXTENSIBILITY TO MEGAWATT SCALES Both the laser array and the PV arrays are easily extended to larger power levels. The mass per unit power of the laser amplifiers is about 5 kg/kW currently with a strong push to bring this down to 1 kg/kW in the next five years. Similarly the PV is about 7 kg/kW, or similar to the laser amplifiers. Interestingly, it is the radiator panels that are the most difficult to scale up, at about 25 kg/kW. This is an area that needs work, though in all simulations for mission masses the assumption is 25 kW (radiated) for the radiator panels. ATK has to scale their existing 10 m diameter design to push the PV arrays to 30 m diameter which will yield about 225 kW per manufactured unit, or 450 kW per pair and still fit in an SLS PF1B 8.4 m diameter fairing. Fig. 9 and Fig. 10 show the scaling and deployment of the PV arrays to larger sizes for various launch vehicles. Even larger sizes into the megawatt range can be anticipated in the future. Figure 9. Solar PV ATK Megflex arrays extended to 30 m diameter and 225 kW per panel for a total of 450 kW per pair. The 30 m diameter panel fits into the SLS fairing. Extension to the megawatt class could be accomplished with multiple units of these or possible extension to larger diameters. Figure 10. Deployment scenario for 30 m diameter 450 kW (pair) of ATK Megaflex panels from packing in an SLS fairing. Radiators are also shown as one possible option. Solar arrays are depicted larger than the emitter array, allowing for less efficient PV conversion. 5. ORBITAL DEFLECTION CAPABILITIES 5.1 How magnitude and duration of applied thrust influence miss distance When an asteroid is exposed to the DE-STARLITE laser, the temperature [K] and flux [W/m2] on the target asteroid must approach sufficiently high levels in order for significant ablation to occur, targeting a temperature on the order of 3000 K and a flux of >107 W/m2. This causes direct evaporation of the asteroid at the spot of contact. Evaporation at the spot produces a vaporization plume thrust [N] that can be used to change the asteroid's orbit and effectively deflect asteroids from colliding with Earth. A miss distance of at least two Earth radii (12742 km) is required to eliminate the threat of collision. The orbital deflection depends on the duration, magnitude, and direction of the applied thrust. Previous results describe analytical and semi‐analytical treatments of orbital deflection. Colombo et al. (2009) use a semi-analytical approach to describe the motion of an asteroid subject to a low‐thrust action with a thrust magnitude inversely proportional to the square of distance from the Sun. A simple low‐thrust formula that shows the dependency of miss distance on t2 is given in Scheeres and Schweickart (2004). Several previous papers explore optimal strategies for deflection Earth-approaching asteroids, including Conway (2001), Carusi et al. (2002), and Vasile and Colombo (2008). Colombo et al. ﴾2009﴿ provides a detailed derivation of the analytical formulae for orbit deflection as well as a comparison with full numerical simulations for different types of orbit. Zuiani et al. ﴾2012﴿ use first-order perturbation solutions of the accelerated motion to calculate an accurate deflection in the case of laser ablation. Bombardelli et al. (2011) develop asymptotic solutions for the deflection of asteroids with low thrust propulsion, including an analytical solution of the miss distance on the b‐plane. In this paper, we utilize a three-body simulation (accounting for the gravitational effects of the Earth, the sun, and the target asteroid) to analyze how the applied thrust and the laser-active time impact the miss distance (Zhang et al., 2016; Zhang et al., 2015-1; Zhang et al., 2015-2). In order to determine the orbital deflection, ∆x, of an asteroid that is being acted on over a period of time, t, an approximation that is commonly used in orbital mechanics was used as a comparison. The detailed numerical simulation is compared to the approximation of multiplying by 3 the naive distance achieved by accelerating and coasting a system that is not a bound gravitational system. Hence the orbital deflection is compared to: ∆xapprox = 3(0.5atactive 2 + atactivetcoast) (3) where a is the acceleration caused by the plume thrust, tactive is the time the laser is active, and tcoast is the coast time (typically zero). The reason this is done is because this approximation is often used for preliminary mission design. Note that compared to an impactor the deflection for the laser case (and other constant-force systems) scales quadratically with the time while the impactor case scales linearly. In reality, the actual orbital deflections are more complex. The numerical simulations were performed in a rotating frame, where the thrust was pointed both along and against the velocity vector for comparison. Many dozens of orbital simulations were analyzed. The first data set shown compares the laser-active time to the miss distance for a given thrust acting on targets of varying diameter. This paper focuses on the 325 m diameter asteroid case, as this is approximately the size of Apophis-a well-known possible threat. Computations have also been done for 20 to 1000 m asteroids under many mission scenarios. The same code is used to analyze the IBD, gravity tractor and impactor (impulse) cases to which DE-STARLITE are compared. A sample of the results for the 325 m asteroid case is displayed in Fig. 11. It is evident that the factor-of-three approximation is indeed only an approximation and in some cases fails badly. Further, for a given period of time, the required force is compared to the miss distance for targets of varying diameter. Simulations were run for targets of diameters between 20 m and 1000 m for varying mission parameters. Again, a sample of this data is displayed in Fig. 12 for a 325 m asteroid. As is evident in Fig. 11 and Fig. 12 applying a thrust either parallel or antiparallel to the motion of the asteroid produces an equivalent deflection (but misses on opposite side of the Earth's orbit). Numerical analysis suggests that applying a thrust of 2.3 N (produced by ~100 kW electrical power @ 35% efficiency) over a period of 15 years will allow a 325 m asteroid to miss the Earth by two Earth radii. In contrast, if the laser were active for ten years, it would require approximately 5 N of force (produced by ~200 kW electrical power@ 35% efficiency) to deflect a 325 m asteroid by two Earth radii, and if it were only active for five years, it would require nearly 20 N of thrust (produced by ~870 kW electrical power@ 35% efficiency) to produce a comparable result. Figure 11. Plots of miss distance vs. laser-active time for an Apophis-sized asteroid (325 m) subject to a thrust of 2 N, 4 N, and 6 N, respectively. These thrusts are achievable with systems of approximately 20, 40 and 60 kWoptical or about 50, 100 and 150 kWelectrical. Note that the analytic approximation and the detailed numerical modeling results approximately agree, but not in detail. This varies on a case-by-case basis, which is the point of being aware of the limitations of analytic approximations. See our other related figures in this paper. Figure 12. Plots of miss distance vs. thrust for an Apophis-sized asteroid (325 m) over a period of five years, ten years, and fifteen years, respectively. It is clear that operating over a longer period of time (longer warning time) greatly simplifies the system requirements in terms of thrust needed and power required. Note again the comparison between the analytic and numerical techniques. 5.2 How optics diameter, distance from the target, and laser power effect the flux on the target For a laser phased array of a given power, a larger aperture enables the spacecraft to be further from the target while still producing sufficient ablation. The ratio of power to spot area must remain >107 W/m2 in order for significant ablation to occur for high temperature compounds. Comets take much less flux due to their high volatility. With 35 kW of optical output, a laser phased array with a 4.5 m aperture can provide sufficient ablation from an approximate distance of 125 km, whereas a 35% efficient laser phased array with a 1 m aperture must be within 28 km of the target. Increased efficiency of the laser amplifiers will provide for even greater range. At 50% efficiency, with 50 kW of optical output, a 4.5 m laser array will have a range of approximately 150 km, while a 1 m array will have a range of roughly 33 km. Several cases are shown in Fig. 13. Figure 13. LEFT: Plot of laser power vs. range for various apertures. For the phased array, the relevant laser power is the sum of all the laser power, while for the close packed fiber array it is the power of EACH individual laser. RIGHT: Spot size vs. range for various apertures. The spot size on the target is approximately independent of whether a phased array or a close packed fiber array is used. It basically only depends on the aperture size and wavelength for a diffraction limited system. For the close packed array alternative, the product of the optical diameter and the ratio of the spot sigma to the focal plane sigma provides an estimate for the target range. If a 1 m diameter optical system is implemented with a 2 mm spot sigma and 4 μm focal plane sigma, a range of 0.5 km can be achieved. Given the same parameters, a 4.5 m optical system will have a range of 2.3 km-much closer than with the phased array. 5.3 Comparison of efficiency of laser ablation versus ion beam deflection (IBD) Ion beam deflection (IBD) and ion beam shepherding (IBS) is an alternative approach to achieve asteroid orbital deflection in which an ion beam / neutralized beam is used to push against the asteroid. In using this approach, the spacecraft must provide twice as much thrust as would otherwise be necessary to deflect the asteroid a desired distance. Half of the thrust is lost in station keeping in order to keep the spacecraft stable, as the spacecraft must push towards or away from the asteroid with an equal amount of thrust. The mass required (Δm) for a desired impulse (Δp) is determined by the following succession of equations. The force F is given by: where vrel is the exhaust velocity. The desired impulse is which can be solved to find the mass required for a desired impulse, The mass of propellant needed to produce an impulse of Δp on an asteroid using ion beam deflection is given by equation 7: (4) (5) (6) F=dmdtVrelDp=FDt=dmdtvrelDt=(Dm)vrelDm=DpVrel (7) where the factor of two accounts for half of the thrust being used to stabilize the spacecraft. Using the IBD approach, the magnetically shielded Hall thruster is essentially reduced from 40 µN/Welec to 20 µN/Welec effective on the asteroid. Using Xenon as a propellant, the exhaust speed is effectively 30 km/s for a Hall effect thruster with an Isp of 3000 s. As a baseline example, to deflect an Apophis-sized asteroid (325 m) a miss distance of two Earth radii requires 2.3 N of thrust over a period of 15 years. According to equations 4-7, this action would require ~72300 kg of Xenon. An additional 5% of this mass is required to account for the tanks needed to hold the propellant, thus totaling 75900 kg. If gridded ion thrusters with an Isp of 6000 s are employed, which may be the case for a dedicated IBD mission, the mass of propellant required to produce a given thrust would be cut in half as the exhaust velocity is twice as great, though the power requirements roughly double in this case. The ion thrusters need to be chosen to have a small enough divergence angle and be close enough to the asteroid so most of the ions hit it. While a significant amount of propellant is required to deflect an asteroid using the IBD method, no extra propellant is necessary after rendezvous with the asteroid (which uses a small amount of propellant from LEO) for the laser ablation case. A significant benefit of using laser ablation is that the asteroid is propelled by the ejection of its own material and thus the asteroid is itself the fuel. The mass of the laser array and the increased mass of the radiator (because the ion engines have a greater efficiency than the laser amplifiers) required for the laser ablation approach are of a far lesser magnitude than the additional mass of fuel needed for the IBD approach. Laser ablation is therefore proposed to be a more mass efficient mechanism by which to deflect an asteroid. A laser ablation system such as the proposed DE-STARLITE system is much lower in mass than an equivalent IBD system. The following data presents preliminary mass estimates that should be treated as such. With roughly 7 tons to LEO, a DE-STARLITE spacecraft could accommodate 100 kW of electrical input, which corresponds to 50 kW of optical output (assuming 50% laser amplifier efficiency). The resulting effective thrust on the target is 2.5 N, given an effective thrust per electrical watt of 25 μN/Welec. This assumes 50% laser efficiency, 70% of the optical power in the central spot (encircled energy) and 80 μN/Woptical for the laser-asteroid coupling. Our simulations predict up to 5 times this amount; however, a conservative lower value is assumed here. Applying this amount of thrust over a period of 15 years would result in the orbital deflection of a 325 m asteroid by over two Earth radii. To produce the same result using IBD requires ~79900 kg (75900 kg for the propellant and tanks, as described above, as well as approximately 4000 kg for the dry mass of the spacecraft) if using magnetically shielded Hall effect thrusters, or ~42000 kg (approximately 38000 kg for propellant and tanks in addition to the dry mass of the spacecraft) if using gridded ion thrusters with an Isp of 6000 s. Note the higher ISP (6000 s) ion engine requires about twice the power of the lower ISP (3000 s) ARM engine for the same thrust. A trade study needs to be done to optimize this. In comparing the systems, the extra ion engine fuel needed to deflect an asteroid could be instead used to massively increase the PV arrays and thus provide even more power to the laser system. Given 14 tons to LEO, a spacecraft utilizing IBD with Hall effect thrusters is estimated to be capable of outfitting 40 kW of electrical input, whereas a spacecraft utilizing laser ablation would be capable of supporting approximately 380 kW of electrical power by redistributing the mass. This corresponds to 190 kW of optical output for laser amplifiers operating at 50% efficiency, or a thrust on the target of approximately 9.5 N-enough thrust to deflect a 250 m asteroid by a miss distance of two Earth radii over a period of five years. Thus with the same mass as a spacecraft that would be used to capture a 5-10 m asteroid, a system using laser ablation could protect the Earth from catastrophic devastation. For a given power, the mass of the spacecraft utilizing laser ablation is approximately independent of the diameter of the target asteroid. Though the laser must be active for more time to deflect an asteroid of larger diameter, it does not require more mass to do so. In contrast, the mass of a spacecraft utilizing IBD increases as the cube of the asteroid diameter in order to accommodate more propellant to provide sufficient integrated thrust. The deflection time in both IBD and laser ablation (as well as most other approaches) increases with the cube of the asteroid diameter. Increasing the power output of the system will decrease the required warning time for a target of a given diameter because it will lessen the required laser-active time. Several warning-time scenarios are depicted in Fig. 14. The DE- STARLITE mission calls for a baseline of 100 kW electrical, though this could be increased while staying within the mission parameters in order to decrease the required warning time. As shown above there is already a path to 30 m diameter class PV arrays that would yield about 450 kW of electrical power per pair. This path is consistent with the launch capabilities of the launch vehicles under consideration for the DE-STARLITE mission. Fig. 15 shows the estimate mission mass (at LEO) for various power scenarios while Fig. 16 shows the required laser on time vs. asteroid diameter. As is readily seen directed energy is extremely effective for even large targets with modest exposure times. Dm=2DpVrel Figure 14. LEFT: Warning time versus asteroid diameter to produce a two Earth radii deflection for various system and optical powers assuming a laser amplifier efficiency of 50%, an 8.5 year build and travel time and an asteroid density of 2 g/cc. RIGHT: Asteroid diameter vs. spacecraft mass (left axis) for the IBD case (utilizing both magnetically shielded Hall effect thrusters with an Isp of 3000 s, and gridded ion thrusters with an Isp of 6000s) and for laser ablation, as well as asteroid diameter vs. the required warning time for a laser ablation system with 100 kW electrical power (right axis). For an equivalent warning time, the IBD case with an Isp of 3000 s requires ~125 kW electrical power, and the IBD case with an Isp of 6000 s requires ~250 kW electrical power. The same parameters (8.5 year build and travel time, 50% efficient laser amplifiers, 2g/cc and 2 Earth radii miss distance) are assumed. Note that the 8.5 year build and travel time is assumed for a spacecraft using ion engines with an Isp of 3000 s; the travel time may be decreased with ion engines of greater specific impulse and efficiency. Figure 15. Spacecraft mass at LEO vs. electrical power available to system. Assumptions include the nominal ATK MegaFlex mass of about 7 kg/kW electrical, 5 kg/kW (optical power) laser amplifiers and 25 kg/kW (radiated) for the radiators and other nominal system bus parameters including ion engines and Xe fuel for LEO to asteroid. For reference an SLS Block 1 is spec'd at 70 metric tons to LEO and a Block 2 is 130 tons to LEO. Optical power from the laser is 1/2 of the electrical power for 50% efficiency laser amplifiers. Other electrical losses due to conversion and additional systems need to be included in a full analysis. 5.4 Impactor Comparison Figure 16. Laser deflection time needed to achieve 2 Earth radii miss distance vs. electrical power available assuming 50% amplifier efficiency and 80 μN/Wopt coupling efficiency. Note that this is the laser on time not the warning time. The warning times in the figures above assume a build and travel time. The time shown here is the time the laser is actually on. The asteroid density is assumed to be 2000 kg/m3. Here we discuss the case of using an impactor (ramming asteroid) vs. using a laser. As a common metric we use the launch mass as a common element for both cases–i.e., for the same launch mass, what can each system do? For a simplistic analysis the impactor delivers a large impulse or momentum transfer to deflect the target (integrated force - time in units of Ns). This momentum transfer imparts a change in the speed ΔV of the asteroid equals Δpimpactor /M where M is the mass of the asteroid. Δpimpactor is the impulse delivered at a time τ before (if un-deflected) impact. Δpimpactor = β·mv where m is the spacecraft mass, v is the relative closing speed between the spacecraft and asteroid and β is an enhancement factor due to asteroid mass ejection from the impact. The enhancement factor is a much debated term that is a complex function of asteroid material properties at the impact site, geometry of impact, speed of impact. In general one assumes pure inelastic collision (β =1) to be conservative as any a priori β is generally not going to be known for a given target. We will assume later that β =1 but it is important to keep this enhancement possibility in mind to be fair. The change of speed is thus: The deflection distance at the Earth is approximately: (9) where the factor of 3 is the same approximation used from orbital dynamics but as we have shown in several of our papers it is not always a good approximation (Zhang et al., 2015-1, Zhang et al., 2015-2) We use it here for analytic purposes for simplicity and because it is often used in mission planetary defense planning exercises. We also note that for impactors there can be an additional effect do the mass ejection upon impact. This depends on the asteroid materials, and the specifics of the impact. Δximpactor = 3 ΔV·τimpact = 3· β·v·τimpact (m/M) ΔV = β·mv/M = β·v (m/M) (8) Note that the miss distance Δximpactor is linearly proportional to the spacecraft or impactor mass (m), the closing speed (v) and time to impact τ and inversely proportional to the asteroid mass M. Note that the asteroid mass M is proportional to the cube of the asteroid diameter D. The momentum change (impulse delivered) is largely independent of Δximpactor = 3 ΔV· τimpact = 18·m·v·τ / (πρD3) the asteroid mass and only depends on the spacecraft mass (m) and the closing speed (v). For a homogeneous asteroid of density ρ then miss distance is: (10) Since the asteroid is moving rapidly with typical speeds of 5-40 km/s we can simplify this to assume the spacecraft is simply in the way of the asteroid (inelastic billiard ball) and thus the speed of the spacecraft relation to the earth is of lesser importance. This of course depends on the specifics of the asteroid orbit (closing from the front vs. the back of the asteroid orbit). Essentially then it is the mass of the spacecraft that is critical to maximize. Once the space craft is launched to LEO it is assumed that ion engines will be used to allow a larger fraction of the launch mass to survive until impact to maximize the impulse. Since the miss distance is proportional to the inverse cube of the asteroid diameter, and the spacecraft mass is limited by the launcher capability, the only free parameter is the time to impact τ. Thus the miss distance is: In other words, the miss distance is proportional to: (12) For the case of directed energy the equivalent miss distance (using the same factor of 3 approximation for the effects of orbital mechanics) is (Chesley and Chodas, 2002): where: laser = 3/2 (a·τ) τlaser = 3/2 ΔV· τlaser = 3/2 (F/M) τ2 laser /(πρD3 ) Δximpactor = 3·ΔV· τimpact = 3· τimpact ·Δp/M = 18·m·v· τimpact /(πρD3) laser /M = 1/2·3· Δplaser · τlaser /M = 9 α P · τ2 Δximpactor ~m·v·τ·D-3 Δxlaser =3·1/2·a· τ2 =3/2 F· τ2 (11) (13) laser a = acceleration imparted due to the laser plume thrust F = laser plume thrust = α P  P = laser power α = laser plume thrust coupling coefficient – thrust per optical watt  = beam efficiency factor – fraction of beam that is in central spot ρ = asteroid density D = asteroid diameter M = asteroid mass = πρD3/6 τlaser = laser ablation time (laser on time) – assumed to be on the entire time before impact and after rendezvous Δplaser = F τlaser = α P  τlaser Δxlaser = 1/2·3·τlaser ·Δplaser /M Δximpactor = 3· τimpact·Δpimpactor /M , or: Δxlaser = 1/2 Δximpactor laser while the impact deflection is proportional to τimpact. This is Note that the laser deflection Δxlaser is proportional to τ2 important as the deflection grows quadratically with time for the laser and linearly with time for the impactor. We assume the laser thrust is constant and the asteroid mass changes very little due to the mass loss from ablation and that the laser plume thrust is proportional to the laser power. See our other papers on the detailed modeling for this. For simplicity we assume α ~80 µN/Woptical based on our conservative laboratory measurements (Brashears et al., 2015). These estimates are consistent with other published results (Gibbings et al., 2013). Note that for the case of directed energy or any constant force (such as ion engines, gravity tractors, etc.) the miss distance: (14) Where while for the impulse delivery (effectively instantaneously at a time τimpact before impact) for the same overall delta momentum delivered to the asteroid is: (15) for the same Δp and τ. Again this is for the simplistic assumption of the factor of 3 to approximate the orbital mechanics effects. The question now becomes, "For a given launch mass which is more effective – impactor or laser?" If we set the miss distance to be the same Δximpactor = Δxlaser, then we can compare the laser-on time to impact-time, both before nominal Earth impact. We have which gives: or (18) Note that the ratios of times τimpact / τlaser grows linearly with τlaser so that the time ratio depends on the specifics of the case and not just on the fixed system parameters α, P, , β, m, v. The real situation is far more complex than the 3 x delta approximations many times and depends on the specifics of the asteroid orbit and mission parameter as shown clearly in Fig. 17. We assume an SLS Block 1 launch of 70000 kg to LEO. For high Isp ion engines of 3000 s (Hall effect thrusters Δxlaser = 1/2·3· τlaser ·Δplaser /M = Δximpactor = 3· τimpact ·Δpimpactor /M τimpact / τlaser = 1/2 Δplaser /Δpimpactor = 1/2 α P  τlaser/ β·mv τimpact = 1/2 α P  τ2 laser/ β·mv (16) (17) baselined for ARM) or 6000 s (gridded ion) a large fraction of the LEO mass will make it to the asteroid. We can show that for the same mass limited launch the laser ablation system takes much less time to deflect the asteroid. This is a critical point. Using the 3x delta analytic approximation we would conclude that for the same launch mass, but this time used for DE-STARLITE we would be able to launch a 1-2 MW laser system and for many scenarios this will be far more effective than an impactor of the same mass. The details of the particular orbits are important but we can draw some basic conclusions. Assuming 60000 kg makes it out to the asteroid and with a closing speed of 10 km/s, the impactor impulse is 6x10 8 Ns. Fig. 17 shows that for this same 70000 kg SLS Block 1 to LEO, we could launch a 1 MW optical power laser delivering ~60 N of thrust on the asteroid for an assumed laser coupling coefficient α ~80 µN/W optical with an assumed beam efficiency in the central spot of 0.7. To get the same deflection in the same time to impact as the impactor, we need the laser system to deliver twice the momentum as the impactor since the impactor delivers the momentum change essentially instantaneously while the laser delivers it slowly over the entire time the laser is on. Hence, we need 1.2x109 N s. At 60 N of laser plume thrust this would require a time τ = 1.2x109 N s/60 N = 2x107 s or about 7 months of laser exposure vs. using the same mass impactor which requires 10 years preemptive hit before Earth impact to obtain the same 2 Earth radii miss. This time ratio depends on the specific of the asteroid orbit. Other differences for real systems are that typical impactor missions need more than one to make sure the impulse was delivered properly and that the asteroid orbital control with an impactor can be quite uncertain. For any real threat, multiple backups would be prudent. Figure 17. Miss distance vs. impulse delivery time before impact for 1 GN s impulse using an impactor on an Apophis class asteroid (325 m diameter) using both analytic approximations as well as detailed numerical simulations (Zhang et al., 2015-1; Zhang et al., 2015-2; Zhang et al., 2016). This impactor mass is somewhat larger than an SLS Block 1 can deliver. A miss distance of 2 Earth radii (typ. min acceptable) would require interdiction about 10 years before impact using this impactor. The seemingly unusual behavior of miss distance vs. time of impactor hit from the full simulation is due to resonance effects from the multiple orbits. It is clear the 3Δv approximation is not always accurate, and can be very misleading in some cases. 6. DIRECTED ENERGY AND ASTEROID ROTATION 6.1 Understanding Rotation Periods All asteroids rotate, but generally quite slowly for larger one. A complete picture of rotation properties is not available, but from the limited data collected on the rotation of larger bodies and the break up speed it is estimated that asteroids in the 0.1-1 km class typically rotate no faster than once per several hours as seen in Fig. 18. Results of detailed observation indicate the rotation properties for more than 6000 significantly rotating asteroids and conclude that fast rotation is not an issue in general for larger asteroids (>150 m) as they are typically gravitational bound rubble piles (Walsh et al., 2012) and for these the maximum rotation is independent of diameter and only depends on density ρ, with an angular speed ω, and rotation period τ given by: (19) 423G,3(cid:160)(cid:160)G(cid:160) (20) Estimated densities are in the range of ~2 [g/cc] yielding a minimum rotation period of about 2.3 hours. This is clearly seen in Fig. 18. Figure 18. Measured rotation period of ~6000 asteroids. A distribution of measured asteroid rotation rates, notice the very sharp cutoff at just above 2 hours for larger diameter asteroids. Data from Minor Planet Center (Harris, 1998). The superfast rotators, those at the lower left with periods < 2.2 hours and D < 0.1 km are likely molecularly bound and form a distinct population. The cutoff in rotation periods is observed to be remarkably sharp (see Fig. 18), and lies very close to 2 hours for asteroids of diameters greater than approximately 150 m, consistent with equation 20. Some smaller asteroids can rotate faster as they can have a tighter binding than purely gravitational (such as an iron meteorite) but these are relatively rare. Even fast rotating asteroids can be dealt with since the mass ejection begins so quickly after the laser is turned on. As is seen in the transient thermal simulations below, the mass ejection and hence thrust begin within about 1 second. It is largely a flux issue so that for the same flux at any distance the mass ejection remains at this rate. This is assuming an asteroid consisting of solid SiO2, which is extremely conservative. Loss is included to mimic the absorption qualities of asteroids, which are very absorptive having typical reflection coefficients around 5-10%. Thus, a rotating asteroid with this rate (1 hour) poses little problem. More interesting perhaps would be an attempt to spin up (or down) an asteroid depending on beam placement as discussed below. This is discussed in detail on one of our recent papers (Griswold et al., 2015) along with laboratory measurements we made that show how effectively we can de-spin asteroids. 7. THERMAL ANALYSIS AND CURRENT MODELS 7.1 Comparison of Thermal Models The thrust produced by DE-STARLITE on an asteroid is calculated using three different modeling approaches, of increasing complexity and realism. Results from the three analyses are compared, which all yield consistent answers. The basic equations are derived from energy conservation: Power in (laser) = Power out (radiation + mass ejection) + Where U= Asteroid internal energy and is effectively from conduction. 41/21/21.1910[g/cc]s3.3hr](cid:160)(cid:160) independent of diamet,er.[dUdtdUdt In the steady state = specific heat [J/kg-K], = Laser flux [W/m2] - (in), = Radiation flux [W/m2] - (out), = Thermal conduction [W/m2] - (in), and = Ejecta flux [W/m2] - (out). Assuming , then: Or Locally: (21) (22) (23) (24) (25) (26) (27) Where K is the thermal conductivity (which can be position and temperature dependent) and is the mass ejection flux [kg/m2-s], and Hv is the heat of vaporization [J/kg]. The heat of fusion, Hf, is included for relevant cases. The heat of fusion is sometimes referred to the heat of sublimation as is sometimes the case for compounds in vacuum. Hf is typically a small fraction of Hv. The mass ejection flux is shown in equation 28which uses vapor pressure. Where: (28) The models vapor pressure for each element and compound is determined using a semi analytic form known as Antoine coefficients A, B and C in equation 29. (29) Where A, B and C are unique per element and compound. Hence: (30) A Gaussian profile is assumed for the laser as an approximation shown in equation 30 where the Gaussian laser power is , and r is the distance from the spot center. In the approximation where the spot is small compared to the asteroid, the equation becomes: In the dynamic case, it is possible to solve for transient heat flow by: (31) (32) dUdt0,dUPPoutindht(cid:160)witvU(cid:160)(cid:160)cdvv(cid:160)Were(cid:160)(cid:160)hcLFcondFradFejectaFinradEjectacondPPPPradEjectacoLndFFF)(cid:160)dA(cid:160)n0F(cid:160)(LradEjectacondFFFFdV0LradEjectacondFFFF4rad(cid:160)FTnAB/(TC)1/21/2EjectavveFenM(2RT)10n(cid:160)HH(cid:160)cond4radEjectav(cid:160)(cid:160)FKT,FT,F(cid:160)Hand(cid:160)(cid:160)e(cid:160)e1/21/2evhevhM(PP)(cid:160)M(2RT)(PP)2MRTe(cid:160)vheM Molar mass kg/molP Vapor pressure PaP Ambient vapor pressure 0 in vacuumcoef. of evaporationvATOPCLGB/AB/TCAB/TC1/2vEjectaev1P10(cid:160)(cid:160)and(cid:160)(cid:160)FM10H2RTTP22r/2TL2PFe222r/2TL2PFe2(cid:160)n (33) (34) In equation 34, it is assumed that K (thermal conductivity) is independent of position, are time independent. In the full 3D time dependent solution, all of the above conditions are invoked and the equations are solved simultaneously using a 3D numeric solver (COMSOL in this case). In the 2D steady state solutions, the thermal conductivity is assumed to be small (this is shown in 3D simulations to be a valid assumption as well as from first principle calculations) and a combination of radiation and mass ejection (phase change) is used: (35) (36) Inversion is not analytically tractable, so numerical inversion is used to get T(FT), which gives Pv(FT), Γe(FT), etc. In this inversion, a function fit is found (to 10th order typically): (37) A Gaussian approximation to the laser profile is used (this is not critical) to get T(r), P v(r), e(r) where r is the distance from the center of the spot. Since radiation goes as the 4th power of T, while the mass ejection from evaporation goes roughly exponentially in T, at low flux levels the outward flow is completely dominated by radiation (the asteroid is heated slightly and it radiates). As the spot flux level increases (spot size shrinks or power increases or both) evaporation becomes increasingly dominant and eventually at about T ~2000-3000 K or fluxes of 106 - 107 W/m2 mass ejection by evaporation becomes the dominant outward power flow and (just as water boiling on a stove) the temperature stabilizes and increasing flux only increases the rate of mass ejection with only very small increases in temperature. The three methods:    1D Energetics alone. Use heat of vaporization and set spot flux to correspond T ~6000 K if the system were completely radiation dominated. No radiation or conduction included, only vaporization. 2D Analytic - Model elements and compound vapor pressure vs. T. Includes radiation emission. Ignore thermal conduction. 3D Numeric - Full 3D FEA including phase change, vapor pressure, mass ejection, radiation and thermal conduction. 7.2 1D Energetics Alone The heat of vaporization of a compound is the energy (per mole or per kg) to remove it from the bulk. Removal energy is related to an effective speed and an effective temperature, which are related to but somewhat different than the physical speed of ejection and the physical temperature of vaporization. To be more precise, the term evaporation refers to molecules or atoms escaping from the material (for example water evaporating), while boiling is the point at which the vapor pressure equals or exceeds the ambient pressure. At any non-zero temperature, there is a probability of escape from the surface: evaporation happens at all temperatures and hence vapor pressure is a quantitative measure of the rate of evaporation. The heat of vaporization is also temperature and pressure dependent to some extent. Table 2 gives thermal properties for various materials in asteroids. These materials have relatively high effective temperatures reflecting the fact that there is a probability distribution of energies and an increase in vapor pressure with respect to temperature (Lubin and Hughes, 2015). Material Hf [kJ/mol] Hv [kJ/mol] M [g/mol] Hv [106 J/kg] Cv [J/kg-K] Veff [km/s] Teff [104 K] SiO2 Al2O3 MgO ZnS 9.0 14.2 77.4 38.0 143 293 331 320 60.1 102.0 40.3 97.5 2.38 2.87 8.21 2.46 730 930 1030 472 1.54 1.69 2.87 1.57 0.573 1.15 1.32 1.28 v(KT)(c)0(cid:160)TddT2vK(cid:160)T(cid:160)c0dTdtv(cid:160)(cid:160)and(cid:160)cLradEjectaTF(cid:160)(cid:160)FFFAB/TC1/241/2TvFTM2RT10(cid:160)HNnnTn1a(logF)T Table 2. List of thermo-physical properties of common high temperature asteroid compounds. Here Hf is the heat of fusion and Hv is the heat of vaporization. veff = [J/kg ] and Teff = (M )/3R where R = k NA ~8.31 The thermal probability distribution has tail areas allowing for escape from the surface at lower temperatures than one would naively conclude from a mean analysis only. If power PT from the laser impinges on the asteroid in a small enough spot to heat to above the radiation dominated point (typically 2000-3000 K for rocky (monolithic) asteroids vs. 300-500 K for comets) it is possible to compute the evaporation flux (mass ejection rate) as: Me = PT / . This is the maximum possible rate of mass ejection. It is possible to get quite close to this maximum if the system is designed properly. 7.3 2D Analytic As mentioned above, this calculation assumes that the thermal conduction is small compared to radiation and mass ejection (a good assumption for most asteroids). Using the equations above and the numerical inversions it is possible to solve for the temperature distribution and thus the mass ejection and thrust on the asteroid among many other parameters. A summary is shown in Fig. 19 for SiO2. The parameter σ (sigma) in the Gaussian beam profile is allowed to vary to show the effects of non-ideal beam formation as well as beam and pointing jitter. As can be seen the system is quite tolerant to errors in beam formation, focus, beam jitter and pointing errors even beyond 10σ as long as the power is high enough. The requirements on a low power system at equivalent distances are more severe. These relationships also show that it is possible to nearly achieve the theoretical maximum mass ejection rate. Also, note the thrust (N) per Watt is close to 0.001 N/W for the 1000 kW case. This is comparable to the Shuttle SRB in thrust per watt. This is not really surprising, considering that conventional propellants are approximately thermal in nature with temperatures close to the maximum sustainable in the combustion chamber and exhaust nozzle (i.e., a few x103 K). More conservative numbers are assumed for system performance, typically 80 µN/Woptical though calculations show the coupling to be between 100 and 500 µN/Woptical depending on the asteroid material composition and the laser flux on target used (Riley et al., 2014). More laboratory measurements are needed for various materials and flux levels. For now, a conservative value of 80 µN/Woptical is assumed. Figure 19. Using SiO2 as the equivalent material. (a) Integrated mass ejection rates vs. sigma case for different powers between 1 kW and 1 MW. (b) Similarly, integrated thrust (N) per watt vs. sigma 7.4 3D Numeric Calculations and Simulations Thousands of 3D model simulations have been run, and a few salient results are apparent. Calculations based on the simplest assumptions, namely energetics, and the conservation of spot flux, were validated. The more sophisticated tools are needed for further analysis and optimization of the system. For the case of dynamic targeting and rotating objects, time evolution has been added to the 3D solver. Some of this is motivated by the need to understand the time evolution of the mass ejection under dynamic situations. This is partially shown in Fig. 20 and Fig. 21, where the time evolution 1/2vHvHvH of the temperature at the center of the spot is shown. It is now possible to simulate full dynamics and apply this to the case of rotating asteroids. The same techniques can be applied to pointing jitter and laser machining (deliberate interior targeting) of the asteroid or other target. The time evolution of the heated spot is shown in Fig. 21. Again, all cases refer to SiO2 as the equivalent material for an asteroid. DE-STARLITE (as a stand-on system) is modeled here with a 1 m laser array, with a Gaussian beam and a total optical power of 1 MW, and spot diameter ~30 mm (σ ~5 mm). We use SiO2 as a reference material; we have also run simulations for 92 elements and a number of compounds relevant to asteroid composition, including olivine family and other ultramafic minerals. All simulations produce essentially similar results for most of the appropriate compounds, within a factor of a few (Lubin and Hughes, 2015). Figure 20. Rotating and stationary 3D plots for SiO2: Using 1 hour rotation period for a 100 m diameter asteroid, yields equal surface temperature distribution as in the stationary steady state case. Temperatures rise to the point of being mass ejection limited, which is about 2600 K in the center of the spot. Solar illumination is modeled with an isotropic average of 350 W/m2. The 1 hour rotation period is faster than the self-gravitating case and is shown as an extreme example of a large rotating asteroid that is not a rubble pile. Figure 21. (a) Temperature, vapor pressure and mass loss distribution vs. distance from center (angle from beam axis). High frequency sub structure is due to numerical meshing. (b) Transient time solution (stationary) of temperature in the spot center (K) vs. Time (seconds) after the laser is turned on at t = 0. Initial temperature is 200 K. Mass ejection begins within 1 second. This case is for a DE-STARLITE with a 1 m optical aperture and 1 MW of optical power (this is a large DE-STARLITE) with a spot diameter ~30 mm (σ ~5 mm) on the target which is approximately 15 km away from the spacecraft. The same spacecraft could be over 100 km away from the target and still have about the same deflection. 7.5 Comparing Results Among Models While the 3D simulations give time transient solutions and include full thermal conduction, they lack the numerical flexibility of the 2D solutions. Results of the temperature distributions for a Gaussian laser illumination are compared, and found to be very close in their predictions. This builds confidence that it is possible to do both 2D and 3D simulations with high fidelity. Fig. 22 shows comparisons of Gaussian beam illuminations; results are nearly identical in the critical central region. Figure 22. Comparison of 2- and 3-D models, hence numeric + analytic values for (a) integrated surface thrust (N) vs. total laser power for sigma between 1 and 25 mm. Note that the spot diameter (~6σ) for a DE-STARLITE kW class system is typically 3 to 75 mm. (b) Central spot temperatures. (In this case: DE-STARLINE – 1 m aperture) The ultimate test will come when comparing model results with laboratory tests. As laboratory tests are refined, the results will feed back into the models for various materials. 7.6 Thermal Conduction Unfortunately it is not possible to bring asteroids into the laboratory to study their thermal properties, so it is necessary to rely on astronomical observations, primarily in the infrared, combined with assumptions about their formation and likely structure, to deduce their properties. Several references (Mueller, 2007; Mueller et al., 2007; Harris, 1998; Delbò et al., 2007; Margot et al., 2002), among many others, have done excellent work in this area and it is possible to use their results. One can derive the thermal properties by studying the time varying temperature as deduced from infrared observations. In this way the thermal inertia  (J/m2 K s1/2) and thermal conductivity K [W/m K] are derived. The relationship between them is: Where:  = [ K C]1/2 (38) = Density [kg/m3] C = heat capacity [J/kg K] Hence: The data is shown in Fig. 23 best fit to published data (Delbò et al., 2007), where D is the asteroid diameter [km] is: With d = 300 [km],  = 0.4, and (39) (40) (41) 2(C(cid:160)K /)(cid:160)(cid:160) dD0.8K3e4D/C(cid:160) Figure 23. Thermal Inertia  - [J/m2 K s1/2] and Thermal Conductivity: K [W/m K] The trend (with some significant deviations) is towards smaller asteroids having larger thermal conductivity and larger asteroids having smaller thermal conductivity as shown in Fig. 23. Some of this may be the point contacts from rubble- pile effect for larger asteroids. A similar trend between asteroid size and thermal inertia is also observed. It is the values that are of interest in the models. A relatively conservative case of K = 1 [W/m K] is assumed. To put this in perspective, some values for common materials are given in Table 3. Material Nickel Iron Granite Ice (solid) SiO2 (solid) Water (liq 0C) Snow (firm) Soil (sandy) Pumice Styrofoam Air Moon (regolith) K [W/m K]  [kg/m3] C [J/kg K]  [J/m2 K s1/2] 91 81 2.9 2.3 1.04 (at 200 C) 0.56 0.46 0.27 0.15 0.03 0.026 0.0029 8850 7860 2750 917 2200 1000 560 1650 800 50 1.2 1400 448 452 890 2000 1000 4200 2100 800 900 (varies significantly) 1500 1000 640 1.9x104 1.7x104 2600 2040 1510 1500 740 600 330 47 5.6 51 Table 3. Common material thermal properties for comparison to the asteroid thermal properties in Fig. 24 Raising laser power from 10 kW to 20 kW resulted in slightly smaller range between minimum and maximum final temperatures with a relatively small effect on the final temperature between the two laser powers. This is to be expected since the effective vapor pressure and hence mass ejection rate and hence power into mass ejection is a strong function of the temperature. For these simulations, a relatively conservative case of K= 1 W/m K is assumed. For values of thermal conductivity between 0.01 and 250 W/m K, the evaporation mass flux and thrust change only slightly, shown in Fig. 24. Figure 24. Extreme values inputs of thermal conductivity set to 0.01-250 W/m K for SiO2 – Using 1 MW laser power, spot diameter is 60 mm, with sigma 10 mm, in this case for a 2 m diameter asteroid. 8. EFFECTS OF ROTATION ON SYSTEM REQUIREMENTS From the simulations shown in Figure 20 it is clear that the mass ejection process begin rather quickly, typically within a second of laser initiation. The time scale for mass ejection is also dependent on thermal conductivity, density, heat capacity and heat of vaporization. It is possible to make an estimate of the effects of asteroid rotation by considering the effective motion of the laser spot in the worst case of the spot on the equator. This is shown in Fig. 25. The spot will then move relative to the asteroid surface at a surface speed determined by both the rotation period and the diameter of the asteroid. One simple way to think about the relative time scales is to compare to mass ejection time (after laser initiation) to the time to move the laser spot by about one spot size. If the spot moves a large amount (compared to the spot size) in the time it takes to begin mass ejection then the system will be seriously compromised in terms of effectiveness. The bottom line is that faster rotating asteroid need higher power levels and slower can use lower power. A possible solution to reduce the average power is to use the laser in a pulsed (higher peak power) mode to de-spin it and then run CW to deflect it to optimize the lowest possible average power needed. The pulsed high peak power mode allows for higher flux so spot smearing effects are not as important and allows the target to be spun down to near zero rate relative to the velocity vector. Once the asteroid is spinning slowly enough the CW laser mode can start for full deflection capability. The next section discusses using directed energy to de-spin the asteroid. Figure 25. Left: Laser spot surface speed at the equator vs. rotation period vs. diameter of the asteroid. The 2.3 hour gravitational binding limit (rubble pile limit) is shown for reference for a density = 2 g/cc asteroid. Right: Laser power needed vs. the spot sigma for a Gaussian laser beam for three different flux level requirements. Typically 10 MW/m2 is sufficient flux for most materials to be efficient at mass ejection. The effective spot diameter can be estimated as ~6 σ depending on how one views a Gaussian beam to spot diameter conversion. The way to think about this is you want the surface speed (Left figure) to be less than one spot diameter per second (since the time to mass ejection is typically around 1 s or less). Knowing the spot speed you can then determine the laser power needed to reach the flux required (Right figure). This then drives you to larger power levels for larger diameter asteroid for the same rotation period. As an example to effectively work with a 100 m diameter asteroid that is a rubble pile (2.3 hour rotation rate) you have about a 5 cm/s spot speed. This then requires a spot sigma around 1-5 cm which requires a power level of about 10-60 kW. For a 300 m diameter asteroid rotating at the rubble limit you have a 12 cm/s spot speed and need a spot sigma of 2-10 cm with a power level of 30-700 kW. Slower rotating asteroids need less power and faster ones need more. Running the laser array in an optional pulsed high power mode (short duty cycle so average power remains the same) can overcome this problem allowing the asteroid to be de-spun first and then fully deflected while running in CW mode. See Griswold et al. (2015) for details. 9. ASTEROID ROTATION MITIGATION 9.1 De-spinning a Rotating Asteroid With laser ablation technology it is possible to change the spin of an asteroid. The small spot and fine control allow the ability to do precision manipulation on a target. This could be useful in de-spinning an asteroid for capture, landing or mining missions as examples. The time it takes to de-spin an asteroid depends on thrust (torque), initial angular velocity and asteroid diameter. Simple calculations allow calculating the torque necessary to de-spin a rotating spherical solid, assuming homogeneous composition and density. The torque can be varied by changing the power level, changing the spot size or moving the spot to different locations relative to the spin axis is shown below. To be able to spin down or spin up an asteroid is one of the unique abilities of a directed energy system. If 1000 hours (about 40 days) are allotted to spin down a 150 m diameter asteroid that has an initial period at the gravitation binding limit (about 2.3 hours) it requires about 20 N of thrust to do so. This would require a fairly large system with about 200 kW of optical power. One option to reduce the required power is to allot more time to spin down (this depends on the threat time to impact) - for example allotting a year to de-spin this asteroid would only require about 2 N of thrust or about 20 kW of optical power. The optimization requires specific details of the threat parameters but de-spinning an asteroid remains an interesting option for a directed energy system. It is possible to derive the time required to de-spin a rotating asteroid by modeling the system below. Assumptions are that the laser power and thus the flux and hence mass ejection is constant over the time used to de-spin and that the illuminated spot is at a constant location relative to the spin axis. The worst case of the spin axis perpendicular to the velocity axis is also assumed. In general the spacecraft will be aligned (or anti aligned) along the velocity axis. The mass loss during this time is assumed to be minimal compared to the total asteroid mass. For these assumptions the torque from the ejection plume is constant and thus the angular acceleration α is constant. In practice a real system will be more complex for many reasons as discussed below but this gives us a first order solution. Solutions are plotted in Fig. 26 and Fig. 27. Consider a rotating asteroid with the following parameters: Ρ = 2000 kg/m3 T = initial rotational period (s) ω0 = 2π/T initial rotation speed (rad s-1) t = desired time to stop rotation (s) L = lever arm, 0< L < R (m) R = Asteroid Radius (m) The sum of torques on the asteroid is: where: To de-spin the asteroid, the final rotational speed must be zero: hence: For a solid sphere: The torque τ(N – m) vs. T becomes: The required thrust F to spin down (stop rotation) for L = R becomes: (42) (43) (44) (45) (46) (47) (48) IItFLFRsin(90I(cid:160))final0t(cid:160)00t233252424IMRMRI8R(cid:160)(cid:160)(cid:160)or(cid:160)(cid:160)(cid:160)1RR53535525016R215(T)IttTtT8R(cid:160)152450F16R(T)8115RL15tRTtT Figure 26. Thrust required as a function of rotation period (hours) to de-spin a 150 m diameter asteroid with a density of 2000 kg/m3. As an example to spin down a 150 m diameter asteroid, that is rotating at a period of about 5 hours, in 1000 hours of illumination (about 40 days) takes about 10 N of thrust. Figure 27. Time to spin down as a function of asteroid diameter and toque applied assuming the asteroid is spinning at the gravitational binding rate (~2.3 hours) with a density of 2000 kg/m3. 9.2 Some Future Work in Rotational Studies • Additional 4D simulations with beam size and flux with varying rotation rates. • Asteroids that are smaller than ~100 km in diameter are rarely close to spherical. It will be necessary to run simulations that are for non-spherical geometries. • Run heterogeneous composition models with a rubble surface and regolith like coating. • Shapes and binary systems (Margot et al., 2002): Asteroids like stars, comes in multiplicity (13% NEA's - Near Earth Asteroids/deflected from the main belt). Precession, Synchronic, and Chaotic motion • • Run full simulations including rotation with orbital dynamics. • Simulations that combine pulsed and continuous (CW) modes to look at optimization. 10. CONCLUSIONS The DE-STARLITE system provides a feasible solution to asteroids and comets that pose a threat to Earth. By utilizing a directed energy approach with a high powered phase locked laser array to vaporize the target surface the thrust generated from the mass ejection plume is able to propel the asteroid threat away from the original collision trajectory towards Earth. DE-STARLITE is a very system at a modest cost. As outlined above, DE-STARLITE employs laser ablation technologies which use the asteroid as the propellant source for its own deflection, and thus is able to mitigate much larger targets than would be possible with other proposed technologies such as IBD, gravity tractors, and kinetic impactors. With the equivalent mass of an ARM Block 1 arrangement (14 tons to LEO - full SLS block 1 is 70 tons to LEO), designed to capture a 5-10 m diameter asteroid, DE-STARLITE can mitigate an asteroid larger than Apophis (325 m diameter), even without keyhole effects. Much smaller DE-STARLITE systems could be used for testing on targets that are likely to pass through keyholes. The same technology proposed for DE-STARLITE has significant long-range implications for space missions, as outlined in other DE-STAR papers. Among other benefits, the DE-STARLITE system utilizes rapidly developing technologies to perform a task previously thought to be mere science fiction and can easily be increased or decreased in scope given its scalable and modular nature. DE-STARLITE is capable of launching on an Atlas V 551, Falcon Heavy, SLS, Ariane V or Delta IV Heavy, among others. Many of the items needed for the DE-STARLITE system currently have high TRL; however, one critical issue currently being worked on is the radiation hardening of the lasers, though it appears achievable to raise this to a TRL 6 within 3-5 years. Laser lifetime also poses an issue, though this is likewise being worked on; a path forward for continuous operation looks quite feasible, with or without redundancy options for the lasers. Given that the laser amplifier mass is small and the system is designed to take multiple fibers in each configuration, redundant amplifiers can be easily implemented if needed. DE-STARLITE is a critical step towards achieving the long-term goal of implementing a standoff system capable of full planetary defense and many other tasks including spacecraft propulsion. DE-STARLITE represents a practicable technology that can be implemented within a much shorter time frame at a much lower cost. DE-STARLITE will help to establish the viability of many of the critical technologies for future use in larger systems. Since all asteroids rotate at varying rates, this will cause the average applied thrust to decrease and this must be taken into account in the system design. A lower limiting rotation period for gravitationally bound objects greater than 150 m is observed to be 2-3 hours consistent with being rubble piles. This effect needs to be taken into account for larger asteroids and for small fast rotators. Since the plume thrust begins within 1 second after the laser is initiated it is possible to compare the time scales of the laser spot motion to the mass ejection time scale to determine the effect of the rotation. In many cases rotation is not a fundamental concern but for those cases where it is, an option is to de-spin the asteroid, since this is an option with the proposed system. Running in a high peak power pulsed mode is one option available to mitigate rotation and allow de-spin. In summary, directed energy is an extremely promising option for true planetary defense. It is modular and scalable and allows for a very cost effective approach that has wide applications beyond planetary defense. We gratefully acknowledge funding from the NASA California Space Grant NASA NNX10AT93H in support of this research. ACKNOWLEDGEMENTS REFERENCES Apgar, H., "Cost Estimating", in Wertz, J.R., Everett, D.F. and Puschell, J.J. (Eds.) (2011). Space mission engineering: the new SMAD, Microcosm Press, Hawthorn, CA, 663-700. Aaron, K. (2002). Spacecraft Thermal Control Handbook, Volume 1: Fundamental Technologies, The Aerospace Press, El Segundo, ch. 6. Belton, M.J.S., Morgan, T.H., Samarasinha, N.H. and Yeomans, D.K. (Eds.) (2004). Mitigation of hazardous comets and asteroids, Cambridge University Press. Bombardelli, C., Amato, D., and Cano, J.L. (2016) "Mission Analysis for the ion beam deflection of fictitious asteroid 2015 PDC," Acta Astronautica, Vol. 118 (2016):296-307. Bombardelli, C., Urrutxua, H., Merino, M., Peláez, J. and Ahedo, E.. (2013) "The ion beam shepherd: A new concept for asteroid deflection," Acta Astronautica, Vol. 90 (2013):98-102. Bombardelli, C., and Baù, G. (2012) "Accurate analytical approximation of asteroid deflection with constant tangential thrust," Celestial Mechanics and Dynamical Astronomy, 114(3), 279-295. Bombardelli, C., and Peláez, J. (2011). "Ion Beam Shepherd for Asteroid Deflection," Journal of Guidance, Control, and Dynamics, Vol. 34, No. 4, July–August 2011. Brashears, T., Lubin, P., Hughes, G.B., Meinhold, P., Suen, J., Batliner, P., Motta, C., Griswold, J., Kangas, M., Johansson, I., Alnawakhtha, Y., Prater, K., Lang, A., and Madajian, J. (2015) "Directed Energy Deflection Laboratory Measurements," Nanophotonics and Macrophotonics for Space Environments IX, edited by Edward W. Taylor, David A. Cardimona, Proc. of SPIE Vol. 9616. Brophy, J.R. (2015) "Advanced Solar Electric Propulsion for Planetary Defense," Presented at Joint Conference of 30th International Symposium on Space Technology and Science 34th International Electric Propulsion Conference and 6th Nano-satellite Symposium, Hyogo-Kobe, Japan July 4 – 10, 2015. Brophy, JR. and Muirhead, B. (2013). "Near-Earth Asteroid Retrieval Mission (ARM) Study," Proceedings of the 33rd International Electric Propulsion Conference. Campbell, J.W. (2000). Laser Orbital Debris Removal and Asteroid Deflection, Occasional Paper No. 20, Center for Strategy and Technology, Air University, Maxwell AFB, Alabama. Carusi, A., Valsecchi, G. B., D'Abramo, G., and Boattini, A. (2002) "Deflecting NEOs in route of collision with the Earth," Icarus, 159(2), 417-422. Chesley, S.R., and Chodas, P.W. (2002). Asteroid close approaches: analysis and potential impact detection. Asteroids III, 55. Colombo, C, Vasile, M. and Radice, G. (2009). "Semi-analytical solution for the optimal low-thrust deflection of near- earth objects," Journal of Guidance, Control, and Dynamics 32.3: 796-809. Conway, B.A. (2001) "Near-optimal deflection of earth-approaching asteroids," Journal of Guidance, Control, and Dynamics, 24(5), 1035-1037. Conway, B.A. (2004) "Optimal interception and deflection of Earth-approaching asteroids using low-thrust electric propulsion," in: Mitigation of hazardous comets and asteroids Belton, M.J.S., Morgan T.H., Samarasinha, N, and Yeomans, D.K. (Eds.): 292-312. Delbò, M., Cellino, A., Tedesco, E. F. (2007). "Albedo and Size Determination of Potentially Hazardous Asteroids: (99942) Apophis," Icarus 188, 266-270. Garshnek, V., Morrison, D., and Burkle, F.M. (2000). "The mitigation, management, and survivability of asteroid/comet impact with Earth," Space Policy, 16(3), 213-222. Gibbings, M. A., Hopkins, J. M., Burns, D., & Vasile, M. (2011). "On Testing Laser Ablation Processes for Asteroid Deflection," Proceedings of the 2011 IAA Planetary Defense Conference, Bucharest, Romania. Gibbings, A., Vasile, M., Watson, I., Hopkins, J. M., and Burns, D. (2013). Experimental analysis of laser ablated plumes for asteroid deflection and exploitation. Acta Astronautica, 90(1), 85-97. Griswold, J., Madajian, J., Johansson, I., et al. (2015) "Simulations of directed energy thrust on rotating asteroids," Nanophotonics and Macrophotonics for Space Environments IX, edited by Edward W. Taylor, David A. Cardimona, Proc. of SPIE Vol. 9616. Gritzner, C, and Kahle, R. (2004). "Mitigation technologies and their requirements," In Mitigation of Hazardous Comets and Asteroids, vol. 1, p. 167. Cambridge Univ. Press. Harris, A.W., Fahnestock, E.G., and Pravec, P. (2009). "On the shapes and spins of 'rubble pile' asteroids," Icarus, 199(2), 310-318. Harris, A.W. (1998), "A Thermal Model for Near-Earth Asteroids," Icarus 131, 291-301. Hughes, G.B., Lubin, P., Griswold, J., et al. (2014). "Optical modeling for a laser phased-array directed energy system," Nanophotonics and Macrophotonics for Space Environments VIII, edited by Edward W. Taylor, David A. Cardimona, Proc. of SPIE Vol. 9226. Hyland, D. C., H. A. Altwaijry, S. Ge, R. Margulieux, et al. (2010) "A permanently-acting NEA damage mitigation technique via the Yarkovsky effect," Cosmic Research 48, no. 5: 430-436. Kahle, R., Hahn, G., and Kührt, E. (2006) "Optimal deflection of NEOs en route of collision with the Earth," Icarus 182, no. 2: 482-488. Koenig, J.D., and Chyba, C.F. (2007) "Impact deflection of potentially hazardous asteroids using current launch vehicles," Science and Global Security 15, no. 1: 57-83. Kosmo, K., Lubin, P., Hughes, G.B., Griswold, J., Zhang, Q. and Brashears, T. (2015) "Directed Energy Planetary Defense," Aerospace Conference 2015 IEEE Proceedings, 7-14 March 2015, ISBN: 978-1-4799- 5379-0. Kosmo, K., Pryor, M., Lubin, P., et al. (2014) "DE-STARLITE - a practical planetary defense mission," Nanophotonics and Macrophotonics for Space Environments VIII, edited by Edward W. Taylor, David A. Cardimona, Proc. of SPIE Vol. 9226, pp. 922604. Lin, J.C. (2002). "Space solar-power stations, wireless power transmissions, and biological implications," Microwave Magazine, IEEE, 3(1), 36-42. Lu, E.T., and Love, S.G. (2005) "A gravitational tractor for towing asteroids." Nature, Vol. 438, No. 7065, pp. 177–178. Lubin, P., Hughes, G.B., Bible, J., et al. (2014) "Toward Directed Energy Planetary Defense," Optical Engineering, Vol. 53, No. 2, pp 025103-1 to 025103-18, doi: 10.1117/1.OE.53.2.025103. Lubin, P. and Hughes, G.B. "Directed Energy for Planetary Defense." Chapter in: Allahdadi, Firooz, and Pelton, Joseph N. (Eds.), (2015). Handbook of Cosmic Hazards and Planetary Defense, Springer Reference, 1127 p., ISBN 978-3- 319-03951-0. Maddock, C, Cuartielles, J.P.S., Vasile, M. and Radice, G. (2007) "Comparison of Single and Multi-Spacecraft Configurations for NEA Deflection by Solar Sublimation," AIP Conference Proceedings, vol. 886, p. 303. Mankins, J. C. (1997). "A fresh look at space solar power: New architectures, concepts and technologies," Acta Astronautica, 41(4), 347-359. Margot, J.L., Nolan, M.C., Benner, L.A.M., et al. (2002). Binary asteroids in the near-Earth object population. Science, 296(5572), 1445-1448. Mazanek, D.D., Reeves, D.M., Hopkins, J.B., Wade, D.W., Tantardini, M., and Shen, H. (2015) "Enhanced Gravity Tractor Technique for Planetary Defense," IAA-PDC 15-04-11, presented at the 4th IAA Planetary Defense Conference-PDC 2015 13-17 April 2015, Frascati, Roma, Italy. McInnes, C.R. (2007) "Near Earth Object Orbit Modification Using Gravitation Coupling," J. of Guidance, Control and Dynamics, Vol. 30, No. 3:870-3. McInnes, C.R. (2004) "Deflection of near-Earth asteroids by kinetic energy impacts from retrograde orbits," Planetary and Space Science 52.7: 587-590. Melosh, H. J., and E. V. Ryan (1997). "Asteroids: Shattered but not dispersed," Icarus 129, no. 2: 562-564. Morrison, D, Harris, A.W., Sommer, G., Chapman, C.R. and Carusi, A. (2002) "Dealing with the impact hazard," Asteroids III" in Bottke, W., (Ed.), Asteroids III, Univ. Ariz. Press: 739-754. Müller, M. (2012). "Surface properties of asteroids from mid-infrared observations and thermophysical modeling," arXiv preprint arXiv:1208.3993. Mueller, M, Harris, A.W. and Fitzsimmons, A. (2007). "Size, albedo, and taxonomic type of potential spacecraft target Asteroid (10302) 1989 ML." Icarus 187.2: 611-615. Murphy, D., Eskenazi, M., Spink, J., et al. (2014). "ATK MegaFlex Solar Array Development and Test", Proceedings of the Space Power Workshop, 1-46. Olds, J., Charania, A., and Schaffer, M. G. (2007). "Multiple mass drivers as an option for asteroid deflection missions," Proceedings of the 2007 Planetary Defense Conference, Washington, DC, pp. S3-7. Park, S.Y. and Mazanek, D.D. (2005). "Deflection of earth-crossing asteroids/comets using rendezvous spacecraft and laser ablation," Journal of Astronautical Sciences, 53(1), 21-37. Patel, P., Aiken, D., Chumney, D., et al. (2013). "Initial Results of the Monolithically Grown Six-Junction Inverted Metamorphic Multi-junction Solar Cell", Proceedings of the 38th Photovoltaics Specialist Conference, 2. Phipps, C.R. (2010). "Can Lasers Play a Role in Planetary Defense?," AIP Conference Proceedings Vol. 1278, No. 1. American Institute of Physics, College Park, MD 20740. Phipps, C.R. and Michaelis, M.M. (1994). "NEO-LISP: Deflecting Near-Earth Objects Using High Average Power, Repetitively Pulsed Lasers," Proceedings of the 23rd European Conference on Laser Interaction with Matter, Oxford, England, LA-UR 94-3124. Popova, O.P., Jenniskens, P., Emel'yanenko, V., et al. (2013). "Chelyabinsk Airburst, Damage Assessment, Meteorite Recovery, and Characterization", Science: Vol. 342 no. 6162 pp. 1069-1073. Riley, J., Lubin, P., Hughes, G.B., et al. (2014) "Directed energy active illumination for near-Earth object detection," Nanophotonics and Macrophotonics for Space Environments VIII, edited by Edward W. Taylor, David A. Cardimona, Proc. of SPIE Vol. 9226. Rogers, D. and Brophy, J. (2001). "Ion Propulsion Technology for Fast Missions to Pluto," Proceedings of the 27th International Electric Propulsion Conference, Pasadena, CA, IEPC-01-179. Sanchez Cuartielles, J. P., Colombo, C., Vasile, M. et al. (2007). "A Multi-criteria Assessment of Deflection Methods for Dangerous NEOs," AIP conference proceedings. Vol. 886. Scheeres, D. J., and Schweickart, R. L. (2004) "The Mechanics of Moving Asteroids," 2004 Planetary Conference: Protecting Earth from Asteroids, Orange County, CA, AIAA Paper 2004‐1446. Schweickart, R., Chapman, C., Durda, D., and Hut, P. (2006) "Threat mitigation: the gravity tractor," arXiv preprint physics/0608157. Vasile, M., and Colombo, C. (2008) "Optimal impact strategies for asteroid deflection," Journal of guidance, control, and dynamics, 31(4), 858-872. Vasile, M., and Maddock, C.A. (2010) "On the deflection of asteroids with mirrors," Celestial Mechanics and Dynamical Astronomy 107, no. 1: 265-284. Vasile, M., Vetrisano, M., Gibbings, A., et al. (2013). "Light-touch2: a laser-based solution for the deflection, manipulation and exploitation of small asteroids." Proceedings of the IAA Planetary Defense Conference. Vetrisano, M., Thiry, N., and Vasile, M. (2015). "Detumbling large space debris via laser ablation," Proceedings of the 2015 IEEE Aerospace Conference, pp. 1-10. Walker, R., Izzo, D., de Negueruela, C., Summerer, L., Ayre, M., and Vasile, M. (2005) "Concepts for Near-Earth Asteroid deflection using spacecraft with advanced nuclear and solar electric propulsion systems," Journal of the British Interplanetary Society 58, no. 7-8: 268-278. Walsh, K. J., Richardson, D. C., and Michel, P. (2012) "Spin-up of rubble-pile asteroids: Disruption, satellite formation, and equilibrium shapes," Icarus, 220(2), 514-529. Wie, B. (2008) "Dynamics and Control of Gravity Tractor Spacecraft for Asteroid Deflection." Journal of Guidance, Control, and Dynamics, Vol. 31. No.5. Sept.-Oct. 2008, pp. 1413-1423. Wie, B. (2007) "Hovering control of a solar sail gravity tractor spacecraft for asteroid deflection," Proceedings of the 17th AAS/AIAA Space Flight Mechanics Meeting, AAS, vol. 7, p. 145. Zhang, Q., Walsh, K.J., Melis, C., Hughes, G.B. and Lubin, P. (in press, 2016). "Orbital Simulations on Deflecting Near Earth Objects by Directed Energy", Publications of the Astronomical Society of the Pacific. Zhang, Q., Walsh, K.J., Melis, C., Hughes, G.B. and Lubin, P. (2015-1). "Orbital Simulations for Directed Energy Deflection of Near-Earth Asteroids", in: Schonberg, W.P., Ed., Proceedings of the 2015 Hypervelocity Impact Symposium (HVIS 2015), Procedia Engineering, Vol. 103: 671-678. Zhang, Q., Walsh, K.J., Melis, C., Hughes, G.B., and Lubin, P. (2015-2) "Orbital simulations on the deflection of Near Earth Objects by directed energy," Nanophotonics and Macrophotonics for Space Environments IX, edited by Edward W. Taylor, David A. Cardimona, Proc. of SPIE Vol. 9616. Zuiani, F., Vasile, M., a Gibbings, A. (2012) "Evidence-based robust design of deflection actions for near Earth objects," Celestial Mechanics and Dynamical Astronomy, 114(1-2), 107-136.
1710.00434
1
1710
2017-10-01T23:20:33
Origin of the RNA World: The Fate of Nucleobases in Warm Little Ponds
[ "astro-ph.EP" ]
Prior to the origin of simple cellular life, the building blocks of RNA (nucleotides) had to form and polymerize in favourable environments on the early Earth. At this time, meteorites and interplanetary dust particles delivered organics such as nucleobases (the characteristic molecules of nucleotides) to warm little ponds whose wet-dry cycles promoted rapid polymerization. We build a comprehensive numerical model for the evolution of nucleobases in warm little ponds leading to the emergence of the first nucleotides and RNA. We couple Earth's early evolution with complex prebiotic chemistry in these environments. We find that RNA polymers must have emerged very quickly after the deposition of meteorites (< a few years). Their constituent nucleobases were primarily meteoritic in origin and not from interplanetary dust particles. Ponds appeared as continents rose out of the early global ocean but this increasing availability of "targets" for meteorites was offset by declining meteorite bombardment rates. Moreover, the rapid losses of nucleobases to pond seepage during wet periods, and to UV photodissociation during dry periods means that the synthesis of nucleotides and their polymerization into RNA occurred in just one to a few wet-dry cycles. Under these conditions, RNA polymers likely appeared prior to 4.17 billion years ago. Significance: There are two competing hypotheses for the site at which an RNA world emerged: hydrothermal vents in the deep ocean and warm little ponds. Because the former lacks wet and dry cycles, which are well known to promote polymerization (in this case, of nucleotides into RNA), we construct a comprehensive model for the origin of RNA in the latter sites. Our model advances the story and timeline of the RNA world by constraining the source of biomolecules, the environmental conditions, the timescales of reaction, and the emergence of first RNA polymers.
astro-ph.EP
astro-ph
Submitted: 7 Jun 2017 Accepted: 28 Aug 2017 Published Online: 2 Oct 2017 Proc Natl Acad Sci USA (2017) vol. 114, no. xx, pg. xx Origin of the RNA World: The Fate of Nucleobases in Warm Little Ponds Ben K. D. Pearce1,4, Ralph E. Pudritz1,2,3, Dmitry A. Semenov2, and Thomas K. Henning2 Prior to the origin of simple cellular life, the building blocks of RNA (nucleotides) had to form and polymerize in favourable environments on the early Earth. At this time, meteorites and in- terplanetary dust particles delivered organics such as nucleobases (the characteristic molecules of nucleotides) to warm little ponds whose wet-dry cycles promoted rapid polymerization. We build a comprehensive numerical model for the evolution of nucleobases in warm little ponds leading to the emergence of the first nucleotides and RNA. We couple Earth's early evolution with complex prebi- otic chemistry in these environments. We find that RNA polymers must have emerged very quickly after the deposition of meteorites (< a few years). Their constituent nucleobases were primarily meteoritic in origin and not from interplanetary dust particles. Ponds appeared as continents rose out of the early global ocean but this increasing availability of "targets" for meteorites was offset by declining meteorite bombardment rates. Moreover, the rapid losses of nucleobases to pond seepage during wet periods, and to UV photodissociation during dry periods means that the synthesis of nucleotides and their polymerization into RNA occurred in just one to a few wet-dry cycles. Under these conditions, RNA polymers likely appeared prior to 4.17 billion years ago. Significance: There are currently two competing hypotheses for the site at which an RNA world emerged: hydrothermal vents in the deep ocean and warm little ponds. Because the former lacks wet and dry cycles, which are well known to promote polymerization (in this case, of nucleotides into RNA), we construct a comprehensive model for the origin of RNA in the latter sites. Our model advances the story and timeline of the RNA world by constraining the source of biomolecules, the environmental conditions, the timescales of reaction, and the emergence of first RNA polymers. Keywords: life origins astrobiology planetary science meteoritics RNA world One of the most fundamental questions in science is how life first emerged on the Earth. Given its ubiquity in living cells and its ability to both store genetic in- formation and catalyze its own replication, RNA prob- ably formed the basis of first life [1]. RNA molecules are made up of sequences of 4 different nucleotides, the latter of which can be formed through reaction of a nu- cleobase with a ribose and a reduced phosphorous (P) source [2, 3]. The evidence suggests that first life ap- peared earlier than 3.7 Gyr ago (Ga) [4, 5] and thus the RNA world would have developed on a violent early Earth undergoing meteoritic bombardment at a rate of ∼1–1000 ×1012 kg/yr [6], which is approximately 8–11 orders of magnitude greater than today [7]. At this time, the atmosphere was dominated by volcanic gases, and dry land was scarce as continents were rising out of the global ocean. What was the source of the building blocks of RNA? And what environments enabled nucleotides to polymerize and form the first functioning RNA molecules under such conditions? Although experiments have pro- duced simple RNA strands in highly idealized laboratory 1 Origins Institute and Department of Physics and Astronomy, McMaster University, ABB 241, 1280 Main St, Hamilton, ON, L8S 4M1, Canada 2 Max Planck Institute for Astronomy, Konigstuhl 17, 69117 Hei- delberg, Germany 3 Center for Astronomy Heidelberg, Institute for Theoretical As- trophysics, Albert-Ueberle-Str. 2, 69120 Heidelberg, Germany 4 To whom correspondence should be addressed. E-mail: [email protected] conditions [8, 9], the answer to these questions are largely unknown. As to the sources of nucleobases, the early Earth's at- mosphere was likely dominated by CO2, N2, SO2, and H2O [10]. In such a weakly reducing atmosphere, Miller- Urey type reactions are not very efficient at producing organics [11]. One solution is that the nucleobases were delivered by interplanetary dust particles (IDPs) and me- teorites. During these early times, these bodies delivered ∼6 ×107 kg yr−1 and ∼2 ×103 kg yr−1 of intact car- bon, respectively [11]. Although nucleobases have not been identified in IDPs yet, three of the five nucleobases (uracil, cytosine, and thymine) have been formed on the surfaces of icy IDP analogues in the lab through exposure to UV radiation [12]. Nucleobases are found in mete- orites (guanine, adenine, and uracil) with concentrations of 0.25–515 parts-per-billion (ppb) [13, 14]. The ultimate source of nitrogen within them could be molecules such as ammonia and HCN that are observed in the disks of gas and dust around all young stars [15–17]. A second pos- sible source of nucleobases is synthesis in hydrothermal vents that well up from spreading cracks on the young Earth's ocean floors [18]. A potential problem here is the lack of concentrated and reactive nitrogen sources in these environments [18]. A central question concerning the emergence of the RNA world is how the polymerization of nucleotides oc- curred. Warm little ponds (WLPs) are excellent candi- date environments for this process because their wet and dry cycles have been shown to promote the polymeriza- tion of nucleotides into chains possibly greater than 300 7 1 0 2 t c O 1 . ] P E h p - o r t s a [ 1 v 4 3 4 0 0 . 0 1 7 1 : v i X r a 2 FIG. 1. History of carbonaceous meteorite deposition in warm little ponds: (A) History of mass delivery rate to the early Earth and of effective warm little pond (WLP) surface area. Three models for mass delivered to the early Earth are compared: a Late Heavy Bombardment model, and a minimum and maximum bombardment model. All mass delivery models are based on analyses of the lunar cratering record [6, 19]. The effective WLP surface area during the Hadean eon is based on a continental crust growth model [20], the number of lakes and ponds per unit crustal area [21], and the lake and pond size distribution [21]. (B) Cumulative WLP depositions for the small fragments of carbonaceous meteoroids of diameter 40–80 m landing in any 1–10 m-radius WLP on the early Earth. A deposition is characterized as a meteoroid debris field that overlaps with a WLP. We assume the ponds per area during the Hadean eon is the same as today, and that all continental crust remains above sea-level, however we vary the ponds per area by ±1 order of magnitude to obtain error bars. 95% of WLP depositions in each model occur before the corresponding dotted vertical line. 40–80 m is the optimal range of carbonaceous meteoroid diameters to reach terminal velocity, while still landing a substantial fraction of mass within a WLP area. 1–10 m is the optimal range of WLP radii to avoid complete evaporation within a few months, while also allowing non-negligible nucleobase concentrations to be reached upon meteorite deposition. The numbers of carbonaceous meteoroid impactors in the 20–40 m-radius range are based on the mass delivery rate, the main-belt size-frequency distribution, and the fraction of impactors that are of type CM, CR or CI [6, 19, 22] (see SI Text). links [8]. Furthermore, clay minerals in the walls and bases of WLPs promote the linking of chains up to 55 nucleotides long [23]. Conversely, experiments simulating the conditions of hydrothermal vents have only succeeded in producing RNA chains a few monomers long [24]. A critical problem for polymerizing long RNA chains near hydrothermal vents is the absence of wet-dry cycles. Model: Fates of Nucleobases in Evolving WLPs We compute a well posed model for the evolution of WLPs, the fates of nucleobases delivered to them, and the emergence of RNA polymers under early Earth con- ditions. The sources of nucleobases in our model are carbonaceous meteorites and IDPs whose delivery rates are estimated using the lunar cratering record [6, 11, 19], the distribution of asteroid masses [19], and the fraction of meteorites reaching terminal velocity that are known to be nucleobase carriers [13, 22]. The WLPs are "tar- gets" in which molecular evolution of nucleobases into nucleotides and subsequent polymerization into RNA oc- curs. The evolution of ponds due to precipitation, evap- oration, and seepage constitute the immediate environ- ments in which the delivered nucleobases must survive and polymerize1. The data for these sources and sinks are gathered from historical precipitation records [25], or are measured experimentally [26, 27]. Sinks for these molecules include destruction by unattenuated UV rays (as the early Earth had no ozone), hydrolysis in the pond water, as well as seepage from the bases of ponds. The data for these sinks are measured experimentally [27–29]. The nucleobases that survive go on to form nucleotides that ultimately polymerize. To calculate the number of carbonaceous meteorite source depositions in target WLPs on the early Earth, we combine a continental crustal growth model [20], the lake and pond size distribution and ponds per unit crustal area estimated for ponds on Earth today [21], the asteroid belt mass distribution [19] and three possible mass deliv- ery models based on the lunar cratering record [6, 19]. This results in a first order linear differential equation (Equation S18), which we solve analytically. The details 1 We don't emphasize groundwater-fed ponds (hot springs) be- cause their small number on Earth today (∼thousands) com- pared to regular lakes/ponds (∼304 million [21]) suggests they didn't contribute greatly to the combined WLP target area for meteorite deposition. 3.73.83.94.04.14.24.34.44.5Ga102010211022102310241025Mass Delivery Rate (kg/Gyr)AMin. Bombard. ModelMax. Bombard. ModelLHB ModelWLP Area3.73.83.94.04.14.24.34.44.5Ga10-1100101102103104105106Cumulative WLP Depositions3840+3e4−346015+135−1410+90−9BMin. Bombard. ModelMax. Bombard. ModelLHB Model0100200300400500600Effective WLP Surface Area (km2) are in the SI Text, and the main results are discussed in the following section. Nucleobase abundances in WLPs from IDPs and me- teorites are described in our model by first order linear differential equations (Equations S34 and S38, respec- tively). The equations are solved numerically (see SI Text for details). Our fiducial model WLPs are cylindrical (because sed- imentation flattens their initial bases) and have a 1 m radius and depth. We find these ponds to be optimal be- cause they are large enough to not dry up too quickly but small enough that high nucleobase concentrations can be achieved (see SI Text). Meteorite Sources and Targets In Figure 1A, we show the history of meteoritic mass delivery rates on the early Earth, and of WLP targets. From total meteoritic mass, we extract just the nucle- obase sources, i.e. only carbonaceous meteoroids whose fragments slow to terminal velocity in the early atmo- sphere. In Figure 1B we show the history of depositions of such nucleobase sources into 1–10 m radius WLP tar- gets from 4.5–3.7 Ga. The lunar cratering record shows either a continuously decreasing rate of impacts from 4.5 Ga forwards, or a brief outburst beginning at ∼3.9 Ga known as the Late Heavy Bombardment (LHB). Due to the lack of lunar im- pact melts with ages prior to ∼3.92 Ga, the rate of mass delivered to the early Earth preceding this date is uncer- tain [30]. Therefore we compare three models for mass delivery to the early Earth based on analytical fits to the lunar cratering record [6, 19]: a LHB, a minimum and a maximum bombardment model. The latter two mod- els represent lower and upper limit fits for a sustained declining bombardment, and are subject to considerable uncertainty [31]. The total target area of WLPs increases as the rising continents provide greater surface area for pond sites. An ocean world, in this view, is an implau- sible environment for the emergence of an RNA world. Iron and rocky bodies impact the solid surface at speeds high enough to form small craters [32]; contributing to the WLP population. Our calculations show that 10–3840 terminal velocity carbon-rich meteoroids would have deposited their frag- ments into WLPs on the Earth during the Hadean eon. Given the large uncertainty of the ponds per unit area and the growth rate of continental crust, we vary the WLP growth function by ± 1 order of magnitude. From this we get a minimum of 1–384 depositions and a max- imum of 100–38,400 WLP depositions from 4.5–3.7 Ga. The optimal model for WLP depositions-the maximum bombardment model-suggests the majority of deposi- tions occurred prior to 4.17 Ga. The less optimal mod- els for WLP depositions, the LHB and minimum bom- bardment models, suggest the majority of depositions oc- curred before 3.82 Ga and 3.77 Ga, respectively. Life Cycles of WLPs 3 Figure 2 illustrates the variation of physical conditions for WLPs during the Hadean eon, approximately 4.5– 3.7 Ga. Annual rainfall varies sinusoidally [25], creating seasonal wet and dry environments. The increased heat flow from greater abundances of radiogenic sources at this time [33] causes temperatures of around 50–80 ◦C [34]. The various factors that control the water level and thus the wet-dry cycles of WLPs are precipitation, evaporation, and seepage (through pores in the ground). In Figure 3A, we present the results of these calcula- tions. We select two different temperatures (65◦C, and 20◦C) as analogues for hot and warm early Earths. For each of these, we examine three different environments: dry, intermediate, and wet (see Table 1 for details). The water levels in the wet environment WLPs range from approximately 60–100% full. WLPs experiencing dry states of approximately half (intermediate environment) and three-quarters of a year (dry environment) only fill up to 20% and 10%, respectively. These results clearly establish the existence of seasonal wet-dry cycles. Nucleobase Evolution in WLPs As shown in Figure 2, the build up of nucleobases in WLPs is offset by losses due to hydrolysis [29], seepage [27], and dissociation by UV radiation that was incident on the early Earth in the absence of ozone [28, 35]. Some protection would be afforded during WLP wet phases, as a 1 m column of pond water can absorb UV radiation up to ∼95% [36]. It is of particular interest that sediment, which collects at the base of WLPs, also attenuates UV radiation. Studies show it only takes a ∼0.6 mm layer of basaltic sediment to attenuate UV radiation by >99.99% [37]. Nucleotide formation and stability are sensitive to tem- perature. Phosphorylation of nucleosides in the lab is slower at low temperatures, taking a few weeks at 65◦C compared to a couple hours at 100◦C [38]. The stability of nucleotides on the other hand, is favoured in warm conditions over high temperatures [39]. If a WLP is too hot (>80◦C), any newly formed nucleotides within it will hydrolyze in several days to a few years [39]. At temper- atures of 5–35 ◦C that either characterize more temper- ate latitudes or a post-snowball Earth, nucleotides can survive for thousand-to-million-year timescales. How- ever at such temperatures, nucleotide formation would be very slow. Considering this temperature sensitivity, we model the evolution of nucleobases in WLPs in match- ing warm (5–35 ◦C) and hot (50–80 ◦C) early Earth en- vironments. Hotter ponds evaporate quicker, therefore we choose rainier analogue sites on a hot early Earth to maintain identical pond environments to our warm early Earth sites. In Figure 3B, we focus on the adenine concentrations in WLPs from only IDP sources. The combination of spikes, 4 FIG. 2. An illustration of the sources and sinks of pond water and nucleobases in our model of isolated warm little ponds on the early Earth. The only water source is precipitation. Water sinks include evaporation and seepage. Nucleobase sources include carbonaceous IDPs and meteorites, which carry up to ∼1 picogram and ∼3 mg of each nucleobase, respectively. Nucleobase sinks include hydrolysis, UV photodissociation, and seepage. Nucleobase hydrolysis and seepage are only activated when the pond is wet, and UV photodissociation is only activated when the pond is dry. TABLE 1. Precipitation models matching dry, intermediate, and wet environments on a warm (5–35◦C) and hot (50–80◦C) early Earth. Precipitation data from a variety of locations on Earth today [25, 40] represent 2 classes of matching early Earth analogues: warm (Thailand, Brazil, and Mexico) and hot (Columbia, Indonesia, Cameroon). For example, the conditions in Mexico on a warm early Earth match the conditions in Cameroon on a hot early Earth. P is the mean precipitation rate, δp is the seasonal precipitation amplitude, and sp is the phase shift. Model Warm early Earth (5–35◦C) Hot early Earth (50–80◦C) (cid:104) Environment Analogue Site Mexico (MEX) Brazil (BRA) Thailand (THA) Cameroon (CAM) Indonesia (IDN) Columbia (COL) Dry Intermediate Wet Dry Intermediate Wet (cid:16) 2π(t−sp) (cid:17)(cid:105) P (m yr−1) δp 1.69 0.50 0.91 0.94 1.8 3.32 3.5 4.5 6.0 0.5 0.2 0.5 sp (yr) 0.3 0.85 0.3 0.3 0.85 0.3 To obtain the rate of the decrease in pond water for a given analogue site, table values are input into this equation: dL dt = 0.83 + 0.06T − P 1 + δpsin τs (see SI Text). flat tops, and troughs in Figure 3B reflects the variations of adenine concentration in response to drying, balance of input and destruction rates, and precipitation during wet periods. In any environment and at any modeled temperature, the maximum adenine concentration from only IDP sources remains below 2 × 10−7ppb (see Fig- ure S6). This is two orders of magnitude below current detection limits [14], making subsequent reactions negli- gible. The nucleobase mass fraction curves are practically independent of pond size (1m < rp ≤ 10 m) once a stable, seasonal pattern is reached (< 35 years). This is because although larger ponds have a larger collecting area for 5 FIG. 3. Histories of pond water and adenine concentration from interplanetary dust particles: (A) The change in water level over time in our fiducial dry, intermediate, and wet environment WLPs due to evaporation, seepage, and precipitation. Precipitation rates from a variety of locations on Earth today are used in the models, and represent 2 classes of matching early Earth analogues: hot (Columbia, Indonesia, Cameroon), and warm (Thailand, Brazil, and Mexico) (for details see Table 1). All models begin with an empty pond, and stabilize within 2 years. (B) The red and black-dotted curves represent the adenine concentrations over time from carbonaceous interplanetary dust particles (IDPs) in our fiducial WLPs. The degenerate intermediate WLP environment used in this calculation is for a hot early Earth at 65 ◦C and a warm early Earth at 20 ◦C. The blue curve represents the corresponding water level in the WLP, with initial empty and full states labeled vertically. Three features are present: 1) The maximum adenine concentration at the onset of the dry phase. 2) A flat-top equilibrium between incoming adenine from IDPs and adenine destruction by UV irradiation. 3) The minimum adenine concentration just before the pond water reaches its highest level. IDPs, they have an equivalent larger area for collecting precipitation and seeping nucleobases. Dominant Source of Surviving Nucleobases In Figure 4 we assemble all of these results and com- pare carbonaceous IDPs to meteorites as sources of ade- nine to our fiducial WLPs. Small meteorite fragments (1 cm in radius) are compared with IDPs in panel A. The effects of larger meteorite fragments (5 cm and 10 cm) on adenine concentration are displayed in panel B. The maximum adenine concentration in our model WLPs from carbonaceous meteorites is 10 orders of magnitude higher than the maximum adenine concen- tration from carbonaceous IDPs. The reason for this huge disparity is simply that carbonaceous meteoroid fragments-each carrying up to a few mg of adenine-are deposited into a WLP in a single event. This allows ade- nine to reach ppb–ppm-level concentrations before seep- age and UV photodissociation efficiently remove it from the WLP in one to a few wet-dry cycles. The maximum guanine and uracil accumulated in our model WLPs from meteorites are also more than 10 orders of magnitude higher than those accumulated from IDPs (see Figures S7 and S11). A maximum adenine concentration of 2 ppm is still ∼1–2 orders of magnitude lower than the initial adenine concentrations in aqueous experiments forming adenosine and AMP [2], however these experiments only ran for an hour. Adenine within larger fragments diffuses over several wet-dry cycles, and during the dry phase no outflow oc- curs. For fragments 1, 5, and 10 cm in radius, 99% of the adenine is released into the pond water within ∼10 days, 8 months, and 32 months, respectively (see Fig- ure S3). Only the adenine that has already flowed out of the fragments gets rapidly photodestroyed, therefore, adenine within larger fragments can survive up to ∼7 years. Thus, even though the carbon delivery rates from IDPs to the early Earth vastly exceed those from meteorites, it is the meteoritic material that is the dominant source of nucleobases for RNA synthesis. Nucleotide and RNA Synthesis To form nucleotides in WLPs, ribose and a reduced P source must be available. Ribose may have formed and stabilized in borate-rich WLPs via the formose reaction (polymerization of H2CO) [41]. Additionally, phosphite has been detected in a pristine geothermal pool represen- tative of early earth, suggesting the potential availability of reduced P to WLP environments [42]. Only the AMP nucleotide has been experimentally synthesized in a sin- gle step involving ribose, reduced P and UV radiation [2]. Because of the rapid rate of seepage (∼1.0–5.1 mm 012345678Time (yr)0255075100Percent full (%)Wet Env.Intermediate Env.Dry Env.ACOL 65◦CIDN 65◦CCAM 65◦CTHA 20◦CBRA 20◦CMEX 20◦C012345678Time (yr)10-310-210-1Adenine Mass Fraction (ppq)Intermediate Env.123BIDN 65 ◦CBRA 20 ◦CWater Level050010001500200025003000Water Mass in WLP (kg)050010001500200025003000Water Mass in WLP (kg)EmptyFull 6 FIG. 4. Comparative histories of adenine concentrations from interplanetary dust particles and meteorites: (A) A comparison of the accumulation of adenine from carbonaceous IDPs and meteorites in our fiducial WLPs. The meteorite fragments are small (1 cm), and originate from a 40 m-radius carbonaceous meteoroid. Adenine concentrations for intermediate (wet-dry cycle) and wet environments (never dry) are compared and correspond to both a hot early Earth at 65 ◦C and a warm early Earth at 20 ◦C (for details see Table 1). (B) The effect of meteorite fragment sizes on adenine concentration. The degenerate intermediate WLP environment used in these calculations is for a hot early Earth at 65 ◦C and a warm early Earth at 20 ◦C. The fragments are either only small in size (1 cm in radius), only medium in size (5 cm in radius), or only large in size (10 cm in radius). Three features are present: 1) Adenine is at its highest concentration at the onset of the pond's dry phase. 2) Upon drying, adenine ceases to outflow from large fragments, and UV radiation rapidly destroys all previously released adenine. 3) Re-wetting allows the remaining adenine within large fragment pores to continue to outflow. The U-shape is due to the increase and decrease in water level. day−1 [27]), nucleotide synthesis would need to be fast, occurring within a half-year to a few years after nucle- obase deposition, depending on meteoroid fragment sizes. This is ample time given that lab experiments show hour- to-week-long timescales are sufficient to form adenosine and AMP [2, 3]. once synthesized Nucleotides, using meteorite- delivered nucleobases, are still subject to seepage, regardless of the temperature. Therefore nucleotide polymerization into RNA would also need to be fast, occurring within one to a few wet-dry cycles, in order to reduce their likelihood of seeping through the esti- mated 0.001–400 µm-sized pores at the WLP base [43]. Experiments show that nucleotides can polymerize into RNA chains possibly greater than 300 monomers by subjecting them to just 1–16 wet-dry cycles [8]. Discussion and Conclusions Seepage is one of the dominant nucleobase sinks in WLPs. It will be drastically reduced if nucleobases are encapsulated by vesicles (spheres of size 0.5-5 µm that form spontaneously upon hydration, whose walls consist of lipid bilayers derived from fatty acids within mete- orites) [44]. However, even if seepage is turned off, max- imum adenine concentrations from IDPs are still negligi- ble ((cid:46)150 parts per quadrillion) (see Figure S8). Also, we note that a cold early Earth (if it occurred [30]) with seasonal or impact-induced freeze-thaw cycles could also be suitable for RNA polymerization and evo- lution for reasons similar to those we have analyzed for WLPs. The cyclic thawing and freezing resembles wet- dry cycles [8]. We conclude that the physical and chemical conditions of WLPs place strong constraints on the emergence of an RNA world. A hot early Earth (50–80◦C) favours rapid nucleotide synthesis in WLPs [38]. Meteorite-delivered nucleobases could react with ribose and a reduced P source to quickly (< a few years) create nucleotides for polymerization. Polymerization then occurs in one to a few wet-dry cycles to reduce the likelihood that these molecules are lost to seepage. This rapid process also re- duces the likelihood of setbacks for the emergence of the RNA world due to frequent large impacts, a.k.a. impact frustration [45]. Sedimentation would be of critical im- portance as UV protection for nucleobases, nucleotides, and RNA [28, 46, 47]. The mass-delivery model providing the most WLP de- positions indicates that the majority of meteorite deposi- tions occurred prior to 4.17 Ga. The first RNA polymers would have formed in WLPs around that time, prefigur- ing the emergence of the RNA world. This implies that the RNA world could have appeared within ∼200–300 million years after the Earth became habitable. 012345678Time (yr)10-910-710-510-310-1101103Adenine Mass Fraction (ppb)AMeteorites - Intermediate Env.Meteorites - Wet Env.IDPs - Intermediate Env.IDPs - Wet Env.012345678Time (yr)10-410-310-210-1100101102103123Intermediate Env.B10cm frags5cm frags1cm frags Acknowledgements We thank the referees for their thoughtful insights. We thank M. Rheinstadter for providing valuable comments on the manuscript. B.K.D.P. thanks the Max Planck Institute for Astronomy for their hospitality during his summer abroad, supported by a Natural Sciences and En- gineering Research Council of Canada (NSERC) Michael Smith Foreign Study Supplement. R.E.P. thanks the Max Planck Institute for Astronomy and the Institute 7 of Theoretical Astrophysics for their support during his sabbatical leave. The research of B.K.D.P. was supported by an NSERC Canada Graduate Scholarship and Ontario Graduate Scholarship. R.E.P. is supported by an NSERC Discovery Grant. This research is part of a collaboration between the Origins Institute and the Heidelberg Initia- tive for the Origins of Life. Author Disclosure Statement The authors declare no conflict of interest. [1] W. Gilbert, Nature 319, 618 (1986). [2] C. Ponnamperuma, C. Sagan, and R. Mariner, Nature 199, 222 (1963). [3] M. Gull, M. A. Mojica, F. M. Fern´andez, D. A. Gaul, T. M. Orlando, C. L. Liotta, and M. A. Pasek, Sci Rep 5, 17198 (2015). [4] A. P. Nutman, V. C. Bennett, C. R. L. Friend, M. J. Van Kranendonk, and A. R. Chivas, Nature 537, 535 (2016). [5] Y. Ohtomo, T. Kakegawa, A. Ishida, T. Nagase, and M. T. Rosing, Nature Geosci 7, 25 (2014). [6] C. F. Chyba, Nature 343, 129 (1990). [7] P. A. Bland, T. B. Smith, A. J. T. Jull, F. J. Berry, A. W. R. Bevan, S. Cloudt, and C. T. Pillinger, Mon Not R Astron Soc 283, 551 (1996). [8] L. Da Silva, M.-C. Maurel, and D. Deamer, J Mol Evol 80, 86 (2015). [9] L. E. Orgel, Crit Rev Biochem Mol Biol 39, 99 (2004). [22] T. H. Burbine, T. J. McCoy, A. Meibom, B. Gladman, and K. Keil, "Meteoritic Parent Bodies: Their Number and Identification," in Asteroids III, edited by W. F. Bot- tke, Jr., A. Cellino, P. Paolicchi, and R. P. Binzel (Uni- versity of Arizona Press, Tucson, 2002) pp. 653–667. [23] J. P. Ferris, A. R. Hill, R. Liu, and L. E. Orgel, Nature 381, 59 (1996). [24] B. T. Burcar, L. M. Barge, D. Trail, E. B. Watson, M. J. Russell, and L. B. McGown, Astrobiology 15, 509 (2015). [25] W. R. Berghuijs and R. A. Woods, Int J Climatol 36, 3161 (2016). [26] C. E. Boyd, Trans Am Fish Soc 114, 299 (1985). [27] C. E. Boyd, Trans Am Fish Soc 111, 638 (1982). [28] O. Poch, M. Jaber, F. Stalport, S. Nowak, T. Georgelin, J.-F. Lambert, C. Szopa, and P. Coll, Astrobiology 15, 221 (2015). [10] D. Trail, E. B. Watson, and N. D. Tailby, Nature 480, [29] M. Levy and S. L. Miller, Proc Natl Acad Sci USA 93, 79 (2011). 7933 (1998). [11] C. Chyba and C. Sagan, Nature 355, 125 (1992). [12] M. Nuevo, C. K. Materese, and S. A. Sandford, Astro- [30] K. Zahnle, N. Arndt, C. Cockell, A. Halliday, E. Nisbet, F. Selsis, and N. H. Sleep, Space Sci Rev 129, 35 (2007). phys J 793, 125 (2014). [13] B. K. D. Pearce and R. E. Pudritz, Astrophys J 807, 85 (2015). [14] M. P. Callahan, K. E. Smith, H. J. Cleaves, J. Ruz- icka, J. C. Stern, D. P. Glavin, C. H. House, and J. P. Dworkin, Proc Natl Acad Sci USA 108, 13995 (2011). [15] B. K. D. Pearce and R. E. Pudritz, Astrobiology 16, 853 (2016). [16] V. N. Salinas, M. R. Hogerheijde, E. A. Bergin, L. I. Cleeves, C. Brinch, G. A. Blake, D. C. Lis, G. J. Melnick, O. Pani´c, J. C. Pearson, L. Kristensen, U. A. Yıldız, and E. F. van Dishoeck, Astron Astrophys 591, A122 (2016), arXiv:1604.00323 [astro-ph.SR]. [17] I. Pascucci, D. Apai, K. Luhman, T. Henning, J. Bouw- man, M. R. Meyer, F. Lahuis, and A. Natta, Astrophys J 696, 143 (2009), arXiv:0810.2552. [18] M. A. A. Schoonen and Y. Xu, Astrobiology 1, 133 (2001). [19] O. Abramov and S. J. Mojzsis, Nature 459, 419 (2009). [20] M. T. McCulloch and V. C. Bennett, Lithos 30, 237 (1993). [21] J. A. Downing, Y. T. Prairie, J. J. Cole, C. M. Duarte, L. J. Tranvik, R. G. Striegl, W. H. McDowell, P. Korte- lainen, N. F. Caraco, J. M. Melack, and J. J. Middelburg, Limnol Oceanogr 51, 2388 (2006). [31] C. F. Chyba, Icarus 92, 217 (1991). [32] J. G. Hills and M. P. Goda, Astron J 105, 1114 (1993). [33] N. H. Sleep, Cold Spring Harbor Perspect Biol 2, a002527 (2010). [34] C. S. Cockell, "Astrobiology: Understanding life in the universe," (Wiley-Blackwell, New Jersey, 2015) p. 187. [35] S. Rugheimer, A. Segura, L. Kaltenegger, and D. Sas- selov, Astrophys J 806, 137 (2015), arXiv:1506.07200 [astro-ph.EP]. [36] B. Wozniak and J. Dera, "Light Absorption by Water Molecules and Inorganic Substances Dissolved in Sea Wa- ter," in Light Absorption in Sea Water, Atmospheric and Oceanographic Sciences Library (Springer-Verlag, New York, 2007) pp. 11–81. [37] B. L. Carrier, L. W. Beegle, R. Bhartia, and W. J. Abbey, in Lunar and Planetary Science Conference, Lu- nar and Planetary Science Conference, Vol. 48 (2017) p. 2678. [38] G. J. Handschuh, R. Lohrmann, and L. E. Orgel, J Mol Evol 2, 251 (1973). [39] H. R. Hulett, Nature 225, 1248 (1970). [40] R. H. Reichle, R. D. Koster, G. J. M. De Lannoy, B. A. Forman, Q. Liu, S. P. P. Mahanama, and A. Tour´e, J Clim 24, 6322 (2011). 8 [41] Y. Furukawa, M. Horiuchi, and T. Kakegawa, Orig Life [66] J. van Brakel and P. M. Heertjes, Int J Heat Mass Trans- Evol Biosph 43, 353 (2013). fer 17, 1093 (1974). [42] H. Pech, A. Henry, C. S. Khachikian, T. M. Salmassi, G. Hanrahan, and K. L. Foster, Environ Sci Technol 43, 7671 (2009). [43] A. Tugrul, in Engineering Geology and the Environment, Proceedings International Symposium on Engineering Geology and the Environment, Vol. 1, edited by P. G. Marinos, G. C. Koukis, G. C. Tsiambaos, and G. C. Stournaras (Balkema, Rotterdam, 1997) pp. 71–90. [44] B. Damer and D. W. Deamer, Life 5, 872 (2015). [45] K. Zahnle and N. H. Sleep, in Comets and the Origin and Evolution of Life, edited by P. J. Thomas, R. D. Hicks, C. F. Chyba, and C. P. McKay (Springer-Verlag, Berlin, 2006) p. 207. [46] N. Y. Dodonova, M. N. Kiseleva, L. A. Remisova, and N. M. Tsycganenko, Photochemistry and Photobiology 35, 129 (1982). [67] S. Car´e, Cem Conc Res 33, 1021 (2003). [68] T. B. Boving and P. Grathwohl, J Contam Hydrol 53, 85 (2001). [69] P. Baaske, F. M. Weinert, S. Duhr, K. H. Lemke, M. J. Russell, and D. Braun, Proc Natl Acad Sci USA 104, 9346 (2007). [70] R. J. Macke, G. J. Consolmagno, and D. T. Britt, Me- teoritics 46, 1842 (2011). [71] D. El Abassi, A. Ibhi, B. Faiz, and I. Aboudaoud, J Geophys Eng 10, 055003 (2013). [72] P. A. Domenico and V. V. Palciauskas, Geol Soc Am Bull 84, 3803 (1973). [73] M. A. Sheremet, J Eng Thermophys 22, 298 (2013). [74] L. Otero, A. D. Molina-Garc´ıa, and P. D. Sans, Crit Rev Food Sci Nutr 42, 339 (2002). [75] A. F. G. Jacobs, B. G. Heusinkveld, A. Kraai, and K. P. [47] W. Kladwang, J. Hum, and R. Das, Sci Rep 2, 517 Paaijmans, Int J Biometeorol 52, 271 (2008). (2012), arXiv:1202.4794 [q-bio.BM]. [48] M. J. Van Kranendonk, Am J Sci 310, 1187 (2010). [49] T. Kenkmann, N. A. Artemieva, K. Wunnemann, M. H. Poelchau, D. Elbeshausen, and H. N´unez Del Prado, Meteoritics 44, 985 (2009). [50] R. S. Clarke, Jr., E. Jarosewich, B. Mason, J. Nelen, M. Gomez, and J. R. Hyde, Smithson Contrib Earth Sci 5, 1 (1971). [51] R. G. Strom, R. Malhotra, T. Ito, F. Yoshida, and D. A. Kring, Science 309, 1847 (2005), astro-ph/0510200. [52] W. F. Bottke, D. D. Durda, D. Nesvorn´y, R. Jedicke, A. Morbidelli, D. Vokrouhlick´y, and H. Levison, Icarus 175, 111 (2005). [53] P. Ehrenfreund, D. P. Glavin, O. Botta, G. Cooper, and J. L. Bada, Proc Natl Acad Sci USA 98, 2138 (2001). [54] E. W. Weisstein, CRC Concise Encyclopedia of Mathe- matics, Second Edition (CRC Press, Boca Raton, 2002). [55] B. Lang and M. Kowalski, Meteoritics 6 (1971). [56] G. J. Consolmagno, R. J. Macke, P. Rochette, D. T. Britt, and J. Gattacceca, Meteoritics 41, 331 (2006). [57] C. Bounama, S. Franck, and W. von Bloh, Hydrol Earth Syst Sci 5, 569 (2001). [58] N. L. Evans, C. J. Bennett, S. Ullrich, and R. I. Kaiser, Astrophys J 730, 69 (2011). [59] J. Kua and J. L. Bada, Orig Life Evol Biosph 41, 553 (2011). [60] C. S. Cockell, "The ultraviolet radiation environment of Earth and Mars: past and present," in Astrobiology. The Quest for the Conditions of Life, edited by G. Horneck and C. Baumstark-Khan (Springer, Berlin, 2002) pp. 219–232. [61] C. M. Corrigan, M. E. Zolensky, J. Dahl, M. Long, and P. J. Burkett, Meteoritics 32 J. Weir, C. Sapp, (1997), 10.1111/j.1945-5100.1997.tb01296.x. [62] G. Matrajt, D. Brownlee, M. Sadilek, and L. Kruse, Meteoritics 41, 903 (2006). [63] P. A. Bland, M. D. Jackson, R. F. Coker, B. A. Cohen, J. B. W. Webber, M. R. Lee, C. M. Duffy, R. J. Chater, M. G. Ardakani, D. S. McPhail, D. W. McComb, and G. K. Benedix, Earth Planet Sci Lett 287, 559 (2009). [64] J. Palguta, G. Schubert, and B. J. Travis, Earth Planet Sci Lett 296, 235 (2010). [65] K. P. Saripalli, R. J. Serne, P. D. Meyer, and B. P. McGrail, Ground Water 40, 346 (2002). SUPPORTING INFORMATION (SI) Calculating WLP Surface Area on the Early Earth The earliest fairly conclusive evidences of life are in the form of light carbon signatures in graphite globules formed from marine sediments, and stromatolite fossils, both which are dated to be 3.7 Gyr old [4, 5]. Therefore nucleobase deposition in WLPs and subsequent reactions to form nucleotides and RNA, would have occurred some- time between ∼4.5–3.7 Ga. Once Theia impacted the proto-Earth ∼50 Myr after Solar System formation, it may have only taken 10–100 Myr for the magma mantle to cool [30]. Furthermore, basic evidence (from the zircon oxygen isotope record) exists of crustal material interacting with liquid water at or near the Earth's surface since 4.3 Ga [48]. Such evidence makes it clear that the Earth had a hydrosphere at that point, and therefore could have formed WLPs on any rising continental crust. An estimate of the surface area of WLPs on the early Earth has not been previously attempted, however it is often suggested that they were typical on early Earth con- tinents [33, 44, 48]. Water-deposited sediments dated at 3.8 Ga indicate early erosion and transport of sediment, therefore at least some continental mass must have been exposed above sea level at that time, on which WLPs could have formed [33]. Continental Crust Growth Model The number of WLPs present at any given time de- pends on the fraction of continental crust above sea level. Calculations from a continental crust growth model shows a linear formation of early crust, increasing from 0 to 12.8% of today's crustal surface area from 4.5–3.7 Ga [20]. This is expressed in the equation below. 9 fcr = 0.16t (S1) Total WLP Surface Area where fcr is the fraction of today's crustal surface area, and t is the time in Gyr (t = 0 starts at 4.5 Ga). We assume the number of bodies of water per unit area of continental crust is constant over time, thus by mul- tiplying Equation S1 by the number of lakes and ponds on Earth today, we get the number of bodies of water on Earth at any date from 4.5–3.7 Ga. The number of lakes and ponds on Earth today (down to 0.001 km2) is estimated to be 304 million [21], therefore the equation becomes: From Equations S2 and S4, the total surface area of cylindrical, 1–10 m-radii WLPs on the early Earth at time 0 Gyr ≤ t ≤ 0.8 Gyr is Atot(t) = 651t [km2]. (S5) For this calculation, we choose c = 1.06 and k = 3.14×10−6 km2. Note t = 0 starts at 4.5 Ga. Nt = 4.864 × 107t (S2) Calculating Carbonaceous Meteorite Depositions in WLPs where Nt is the total number of bodies of water on Earth for times 0 Gyr ≤ t ≤ 0.8 Gyr (t = 0 starts at 4.5 Ga). Lake and Pond Size Distribution The size distribution of lakes and ponds on Earth today follows a Pareto distribution function [21]. dNA = NtckcA−(c+1)dA, (S3) where dNA is the number of bodies of water in the km2 area range A to A + dA, Nt is the total number of bodies of water on Earth, c is the dimensionless shape parame- ter, and k is the smallest lake/pond area in the distribu- tion. The shape parameter was calculated for lakes and ponds down to 0.001 km2 to be c = 1.06 [21]. The total area of ponds on Earth in a given size range can then be calculated by multiplying Equation S3 by A and integrating from Amin to Amax, which gives Ntkc(cid:104) Atot = c −c + 1 (cid:105) max − A(−c+1) A(−c+1) min [km2]. (S4) There is an upper size limit on WLPs in which substan- tial concentrations of nucleobases from meteorites can be deposited. If the surface area of a WLP is comparable to the area of a meteoroid's strewnfield, then partial over- lapping of the strewnfield and WLP is probable. This would lead to fewer fragment depositions per unit pond area, and lower nucleobase concentrations. For this rea- son, we choose the upper limit on WLP radii to be 10 m. This equates to pond areas that are <0.04 % of the strewnfield area of typical carbonaceous meteoroids (the latter of which are ∼0.785 km2). There is also a lower size limit as ponds that evaporate too quickly spend the majority of their time in the dry state. This would prevent nucleobase outflow from the pores of meteorites. Moreover, the probability of mete- orite deposition in WLPs decreases for decreasing pond radii. A cylindrical WLP at 65 ◦C with a radius and depth of < 1 metre will likely evaporate in less than 3 months without replenishment from rain or geysers [26]. Therefore we set the lower WLP radius and depth limit for our interests at 1 m. Aerodynamic forces fragment meteoroids that enter the atmosphere, which increases their total meteoroid cross-section, and thus their aerodynamic braking [32]. Numerical simulations show that the fragments of car- bonaceous meteoroids with initial diameters up to 80 metres and atmospheric entry velocities near the me- dian value (15 km/s [31]) reach terminal velocity [32]. These fragments do not produce craters upon impact [49] and can be intactly deposited into WLPs. Larger meteoroids produce fragments with impact speeds too fast to avoid cratering and partial or complete melting or vapourization of the impactor. The same numerical simulations calculate the largest fragments from a 80- metre-diameter carbonaceous meteoroid of initial veloc- ity 15 km/s to be ∼20 cm in diameter [32]. (This roughly makes sense as the biggest carbonaceous meteorite recov- ered, the Allende, is only 110 kg-corresponding to a ∼ 40 cm-diameter sphere [50]). The optimal diameter range for carbonaceous mete- oroids to deposit a substantial fraction of mass into WLPs at terminal velocity is therefore 40–80 m. We base our calculation of carbonaceous meteoroid fragment depositions on this range. Mass Delivery Models The lunar cratering record analyzed by the Apollo pro- gram has revealed a period of intense lunar bombardment from ∼3.9–3.8 Ga [51]. Whether a single lunar cataclysm (lasting ∼10–150 Myr) or a sustained declining bombard- ment preceded 3.9 Ga is still debated [30]. We choose three models for the rate of mass delivered to the early Earth: a LHB model, and a minimum and maximum bombardment model. All models are based on analyses of the lunar cratering record [6, 19]. Analyses from both dynamic modeling and the lunar cratering record estimate the total mass delivered to the early Earth during the LHB to be ∼ 2× 1020 kg [19]. We assume a Gaussian curve for the rate of impacts during the LHB [30], which centers on 3.85 Ga, integrates to 2 × 1020 kg, and drops to nearly zero at 3.9 and 3.8 Ga. Thus, dmLHB(t) = 5.33 × 1021e −(t−0.65)2 2(0.015)2 dt [kg], (S6) where dmLHB is the total mass from t to t + dt (Gyr) (t = 0 starts at 4.5 Ga). Equations for the minimum and maximum mass de- livered to the early Earth, given a sustained declining bombardment preceded 3.9 Ga, are displayed below [6]. mminB(t) = 1 × 1021 − 0.76 4.5 − t + 4.57 × 10−7(cid:16) e(4.5−t)/τa (cid:104) − 1 m1−b 2 4πR2 mmaxB(t) = 7×1023−0.4 (cid:104) 4.5−t+5.6×10−23(cid:16) moonΣ [kg]. (S7) e(4.5−t)/τc (cid:17)(cid:105) (cid:17)(cid:105) − 1 m1−b 2 4πR2 moonΣ [kg]. (S8) where mminB and mmaxB are the total mass delivered from t = 0 to t = t (t = 0 starts at 4.5 Ga), τa and τb are decay constants (τa = 0.22 Gyr, τb = 0.07 Gyr), m2 is the maximum impactor mass (m2 = 1.5 × 1018 kg), b = 0.47, Rmoon is the mean radius of the moon (Rmoon = 1737.1 km), and Σ is the ratio of the gravitational cross sections of the Earth and Moon (Σ ∼23). Taking the derivatives of Equations S7 and S8 gives us the corresponding rates of mass delivered to the early Earth. dmminB(t) (cid:18) = 0.76 1 4.57 × 10−7 τa + e(4.5−t)/τa dmmaxB(t) (cid:18) = 0.4 + 1 5.6 × 10−23 τc e(4.5−t)/τc (cid:19) (cid:19) m1−b 2 4πR2 moonΣdt [kg]. (S9) m1−b 2 4πR2 moonΣdt [kg]. (S10) See Figure 1 in the main text for a plot of the three mass delivery models. 10 to that of the main-belt asteroids today [52]. Though there is no conclusive evidence to constrain the fraction of early Earth impactors that were of cometary origin, some suggest approximately 10% of the total accreted mass was from comets [6]. The early-Earth impactor mass distribution for im- pactors with radii 20–40 m, adjusted for the total mass delivered during the LHB, follows the linear relation dmLHB(r) =(cid:2)7.5 × 1013r + 3 × 1015(cid:3) dr [kg], (S11) where dm is the mass of impactors with radii r to r + dr (m) [19]. To get the impactor mass distributions for the mass delivered between t and t+dt in each of our three models, we multiply Equation S11 by Equation S6, S9, or S10, and divide by 2 × 1020 kg. dmi(t, r) =(cid:2)7.5 × 1013r + 3 × 1015(cid:3) dmi(t) 2 × 1020 dr [kg], (S12) where i is the model (LHB, minB, or maxB). Impactor Number Distribution Equation S12 can be turned into a number distribu- tion (from t to t + dt) for asteroids of a specific size and type (e.g. carbonaceous chondrites, ordinary chondrites, irons) by multiplying by the fraction of impactors that are asteroids (fa) and the fraction of asteroid impacts that are of a specific meteorite parent body-type (ft), and then dividing by the volume and average density of such asteroids. After simplification, this is 3 × 1015 (cid:20) 7.5 × 1013 (cid:21) dmi(t) dNi(t, r) = 3faft 4πρ r2 + r3 2 × 1020 dr, (S13) where i is the model (LHB, minB, or maxB). Total Number of Carbonaceous Impactors Although some carbonaceous chondrites may have originated from comets [53], for this calculation we con- servatively assume that all carbonaceous meteorites orig- inated from asteroids. Integrating Equation S13 from rmin to rmax gives the total number of meteoroids in that radii range to have impacted the early Earth, from t to t+dt, for each model. (cid:20) (cid:18) 1 (cid:19) (cid:19)(cid:21) dmi(t) − 1 rmax 2 × 1020 , 3faft 4πρ 3 × 1015 7.5 × 1013 (cid:18) 1 + 2 r2 min rmin − 1 r2 max (S14) where i is the model (LHB, minB, or maxB). Impactor Mass Distribution dNi(t) = Chemical analyses of lunar impact samples, and crater size distributions suggest that the impactors of the Earth and Moon before ∼3.85 Ga were dominated by main- belt asteroids [51]. It also likely that the size-frequency distribution of impactors on the early Earth is similar Probability of WLP Deposition The probability of an infinitesimally small object hit- ting within a target area, given the probability of hitting anywhere within the total area is equal, is the geometric probability, Pg = Atarg Atot . (S15) This equation can be used to estimate the probability of blindly hitting the bull's eye on a dart board from rel- atively close in, where Atarg is the area of the bull's eye and Atot is the area of the dart board. In this case, since the tip of a dart is relatively small with respect to the bull's eye, the approximation is relatively accurate. In the case of estimating the probability of meteoroid frag- ments falling into a WLP, the tip of the "dart" (i.e. the debris area of meteoroid fragments) is not always going to be larger than the "bull's eye" (i.e. the total surface area of WLPs at any time). For example, the probabil- ity of fragments from a single meteoroid falling into an WLP at ∼4.5–4.45 Ga, when there was a small fraction of continental crust and few WLPs, is more analogous to the probability of blindly hitting the bull's eye with a small ball; which would be slightly more likely than with a dart. Any large enough part of the ball overlap- ping with the bull's eye counts as a hit, as does any large enough part of the debris field overlapping with the com- bined WLP surface area. "Large enough" in our case corresponds to the area of the largest WLP for which the meteoroid fragment deposition probability is being cal- culated (or equivalently 100 of the smallest WLP in our study). (This minimum area is necessary in order to as- sume a homogeneous surface deposition when calculating the total mass of meteoroid fragments to have entered a WLP.) In order to use Equation S15 to calculate the probabil- ity for the fragments of a single meteoroid landing in a primordial WLP, Atarg must be the effective target area. The effective target area is illustrated in Figure S1 below. Any meteoroid that enters the atmosphere within the ef- fective target area (of radius d), disperses its fragments over at least one WLP's entire surface. The area of the asymmetric "lens" in which any two (cid:112)(−d + r + R)(d + r − R)(d − r + R)(d + r + R), 2dR + R2cos−1 − 1 2 (S16) for circles of radii r and R, and distance between their centers d [54]. In Figure S1, the asymmetric lens created by the in- tersection of the combined WLP surface area at a given circles intersect is A = r2cos−1 (cid:18) d2 + r2 − R2 (cid:19) (cid:19) (cid:18) d2 + R2 − r2 2dr 11 FIG. S1. The dotted circle represents the target area for landing the dispersed fragments from a single meteoroid into any WLP on the early Earth. The smaller, light grey circle in the center is the total combined WLP surface area on the early Earth at a given time (WLPs would have been individu- ally scattered, however for a geometric probability calculation they can be visualized as a single pond cross-section). The larger, surrounding circles represent the area of debris when the fragments of the meteoroid hit the ground. Shortly af- ter 4.5 Ga, the area of debris from a single meteoroid would be large compared to the the combined WLP surface area. At any time, the target area for a meteoroid to deposit frag- ments homogeneously into at least 1 WLP is slightly larger than the combined area of WLPs. The effective target area grows linearly with total WLP surface area. The diagonally striped intersections between the circles represent the largest individual WLP for which the meteoroid fragment deposition probability is being calculated. The logic is that if a mete- oroid enters the atmosphere at distance d from the center of the combined pond cross-section, at least 1 pond of any size in the WLP distribution will completely overlap with the area of debris. time and the meteoroid fragment debris area corresponds to the area of the largest individual WLP in our distri- bution. Because the effective target area radius, d, grows linearly with total WLP surface area radius, R, we can plot Equation S16 at t = 1 Gyr to solve for d, and input the linear time dependence afterwards. For this calcu- lation, A = 3.14 × 10−4 km2, r = rg = 0.5 km, and R = 14.4 km. This gives us the effective tar- = get radius, d = 14.9 km, and corresponds to an effective target area: (cid:113) 651km2 π Aef f (t) = 697t [km2]. (S17) See Figure 1 in the main text for a plot of the effective WLP surface area over time. Finally, the probability of the fragments from all CM- , CI-, or CR-type meteoroids with radii rmin = 20 m to rmax = 40 m landing in WLPs on the early Earth of radii 1–10 m from time t to t + dt is the product of Equations S14 and S15. dPi(t) = 697t 4πR2⊕ dNi(t), (S18) where i is the model (LHB, minB, or maxB). In Figure S2, we plot the normalized probability dis- tributions (dPi/dt) for the deposition of carbonaceous meteorites into WLPs from 4.5–3.7 Ga. The LHB, the minimum and the maximum bombardment models are compared. For the LHB model, there are 10 WLP de- positions from 3.9–3.8 Ga. 95% of these depositions oc- cur between 3.88–3.82 Ga. For the minimum bombard- ment model, there are 15 WLP depositions during the entire Hadean eon, however 95% of these depositions oc- cur between 4.47–3.77 Ga. Finally, for the maximum bombardment model, there are 3840 depositions during the Hadean eon, with 95% of these depositions occurring between 4.50–4.17 Ga. See Figure 1 in the main text for cumulative deposition distributions. FIG. S2. Normalized probability distributions of fragments from CM-, CI-, and CR-type meteoroids with radii 20–40 m landing in WLPs on the early Earth of radii 1–10 m. Three models for mass delivery are compared: the Late Heavy Bom- bardment model, and minimum and maximum mass models for a sustained, declining bombardment preceding 3.9 Ga. All models are based on analyses of the lunar cratering record [6, 19]. See Figure 1 in the main text for display of mass delivery rates. The 95% confidence intervals are shaded, and correspond to the most likely deposition intervals for each model. One assumption is made in our deposition probability calculation, which is, given that the largest WLP consid- ered is completely within the carbonaceous meteoroid's 12 strewnfield, at least one fragment will land in the WLP. (An equivalent assumption would be, given that 100 of the smallest WLPs considered are completely within the meteoroid's strewnfield, at least one fragment will de- posit into one of the ponds.) To determine whether this is the case, we need a good estimate for the number of fragments that spread over a debris field from a single car- bonaceous meteoroid. For the Pu(cid:32)ltusk stony meteoroid, which entered Earth's atmosphere above Poland in 1868 [55], the number of fragments is estimated to be 180,000 [55]. Although the Pu(cid:32)ltusk meteorites may not repre- sent the typical fragmentation of stony meteoroids, since this meteoroid is about 1.6 times denser than the aver- age carbonaceous meteoroids in our study [56], we would expect a carbonaceous meteoroid to fragment more than the typical stony meteorite. If meteoroid fragments are randomly spaced within a strewnfield, then Equation S15 multiplied by the the number of meteoroid fragments will give us roughly the number of fragments that will enter a WLP. (In this case Atarg is the area of the cylindri- cal pond, and Atot is the area of the strewnfield.) From this calculation, given a 10 m-radius WLP is within a 180,000-fragment strewnfield of radius 500 m, roughly 72 carbonaceous meteoroid fragments will land in the pond. In fact, as long as 40–80 m carbonaceous meteoroids frag- ment into at least 2500 pieces, fragment deposition into a 10 m-radius WLP is probable. We therefore consider this assumption reasonable. Sensitivity Analysis In our estimate of the number of carbonaceous mete- oroids that led to WLP depositions on the early Earth, we assume that from 4.5–3.7 Ga, the continental crust grows at 16%/Gyr [20], and the ponds per unit area is the same as today. However, the uncertainty in the growth rate of continental crust and the WLPs per unit area on the early Earth is high. For example, the number of WLPs per unit area was probably higher on the above- sea-level crust of the Hadean Earth than it is on Earth's continents now. The greater rate of asteroid bombard- ment during the Hadean [30] would have created many small craters, which over a few rain cycles could fill to become ponds. On the other hand, geophysical models suggest the surface ocean was increasing in volume from ∼4.5–4.0 Ga [57]. This would slow the growth rate of above sea-level crust. Because of these uncertainties, we adjust the growth rate of WLPs on the early Earth by ± 1 order of magnitude to obtain error bars. Because the probability of deposition (Equation S18) is directly proportional to the number of WLPs on the early Earth, varying the growth rate of WLPs by ± 1 order of mag- nitude equates to a ± 1 order of magnitude uncertainty in the WLP deposition expectation values (see Figure 1 in the main text). 3.73.83.94.04.14.24.34.44.5Ga0123456Probability Distribution (depositions/Gyr)LHB ModelMin. Bombard. ModelMax. Bombard. Model Sources and Sinks Model Overview Calculating the water and nucleobase content in WLPs over time is a problem of sources and sinks. These sources and sinks are illustrated in Figure 2 in the main text, and displayed in Table S1 below. TABLE S1. Sources and sinks of pond water and nucleobases in our model of early Earth WLPs. Ingredient Sources Sinks Pond water Precipitation Evaporation Nucleobases IDPs Meteorites∗ Seepage Hydrolysis∗ Seepage∗ Photodissociation† Forming nucleotides‡ *Only when pond is wet † Only when pond is dry ‡ To be added in a future model Because nucleobase diffusion from carbonaceous mete- orite fragments is slow (see section on nucleobase outflow and mixing below), this source is only turned on in the wet phase. Although it may be possible for nucleobases to occasionally enter WLPs from runoff, we do not con- sider this as a source. IDPs that fall on dry land are the most likely nucleobase source to be carried by runoff into WLPs (because of their low mass). However, these IDPs are exposed to photodissociating UV light until they are picked up by a runoff stream, by which time few, if any enclosed nucleobases likely remain. Hydrolysis of nucle- obases only occurs in the presence of liquid water, there- fore we only turn on this sink when our ponds are wet. Similarly, seepage of nucleobases through the pores in the bases of WLPs only occurs during the wet phase. A 1 m column of pond water can absorb UV radiation up to ∼95% [36], therefore as a first order approximation, we only turn on UV photodissociation when our WLPs are in the dry phase. We do not consider cosmic rays as a sink for nucleobases, as a study has shown that adenine has a half-life of millions of years from cosmic ray disso- ciation at 1 AU [58]. This is several orders of magnitude greater than the timescales of nucleobase decay from the other sinks. Finally, since we are interested in accumulat- ing nucleobases in WLPs so that they can react to form nucleotides, a future model will include such reactions as a sink. Pond Water Sources and Sinks Evaporation There are many variables that could go into a pond evaporation calculation. However, a simple relation was obtained by measuring the depth and temperature of a 13 ∼1 m-radius lined pond (to prevent seepage) and a class- A cylindrical evaporation pan over time: dE dt = −9.94 + 5.04T [mm month−1], (S19) where dE is the drop in pond depth, and T is the pond temperature in ◦C [26]. This relation is converted to m yr−1 below. = −0.12 + 0.06T [m yr−1]. (S20) dE dt Seepage Unless the material (e.g. basalt, soil, clay) at the base of a WLP is saturated in water, gravity will cause pond solution to seep through the pores in this material. The average seepage rate of 55 small ponds in Auburn, Al- abama was measured to be 5.1 mm day−1 [27]. This value is high compared to the average seepage rates from small ponds in North Dakota and Minnesota (1.0 mm day−1), and the Black Prairie region of Alabama (1.6 mm day−1) [27]. These seepage rates are comparable in magnitude to the pond evaporation rates above, and therefore must be considered as a sink for water and nucleobases in our WLP model. We take an average of the above three val- ues and apply a constant seepage rate of 2.6 mm day−1 to our WLP water evolution calculations. dS dt = 0.95 [m yr−1]. (S21) We assume that the majority of seepage occurs from the base of the pond, where the water pressure is the highest. The effect of seepage on nucleobase mass loss is handled in the nucleobase sinks section below. Precipitation (cid:20) (cid:18) 2π(t − sp) (cid:19)(cid:21) τs If we ignore the possibility of local water geysers, the main source of pond water is likely to be precipitation. It has been shown that the vast majority of monthly pre- cipitation climates around the world can be adequately described by a sinusoidal function with a 1 year period [25]. That is, = P dP dt 1 + δpsin [m yr−1], (S22) where dP is the amount of precipitation, P is the mean precipitation rate (m yr−1), δp is the dimensionless sea- sonal precipitation amplitude, sp is the phase shift of pre- cipitation (yrs), t is the time (yrs), and τs is the duration of the seasonal cycle (i.e. 1 yr) [25]. From 1980–2009, the mean precipitation on the Earth ranged from ∼0.004–10 m yr−1 (with a global mean of ∼0.7 m yr−1), the seasonal precipitation amplitudes ranged from ∼0–4.7, and the latitude-dependent phase shift ranged from 0–1 year [25]. Summary UV photodissociation 14 (cid:20) (cid:18) 2π(t − sp) (cid:19)(cid:21) τs The evaporation and seepage rates minus the rate of water rise due to precipitation gives us the overall rate of water decrease in a WLP. dL dt 1 + δpsin = 0.83+0.06T−P [m yr−1]. (S23) Equation S23 is too complex to be solved by standard analytical techniques or mathematical software, there- fore we solve it numerically using a forwards time finite difference approximation with the boundary conditions 0 ≤ L(T, t) ≤ rp. We then convert the drop in pond depth as a function of time, L(t), to the mass of water in the WLP using the water density and the volume of a cylindrical portion. This equates to: m(t) = πρwr2 p(rp − L) [kg]. (S24) For our models we assume pond water of density ρw = 1000 kgm−3. Nucleobase Sinks Hydrolysis The first-order hydrolysis rate constants for adenine (A), guanine (G), uracil (U), and cytosine (C) have been measured from decomposition experiments at pH 7 [29]. These rate constants are expressed in the Arrhenius equa- tions below. kA = 10 −5902 T +8.15 kG = 10 −6330 T +9.40 kU = 10 −7649 T +11.76 (S25) (S26) (S27) kC = 10 −5620 T +8.69 [s−1]. (S28) The nucleobase decomposition rate due to being dis- solved in water can then be obtained by plugging Equa- tions S25–S28 into the first-order reaction rate law below. = −miγk where γ = 3600 · 24 · 365.25 s yr−1. dmi dt [kg yr−1], (S29) The hydrolysis rates of adenine, guanine, and cytosine remain relatively stable in solutions with pH values from 4.5–9 [29, 59]. WLPs may have been slightly acidic (from pH 4.8–6.5) due to the higher partial pressure of CO2 in the early Earth atmosphere [34]. The photodestruction of adenine has been studied by irradiating dried samples under Martian surface UV conditions [28]. This quantum efficiency of photode- composition (from 200–250 nm)-which is independent of the thickness of the sample-was measured to be 1.0 ± 0.9 × 10−4 molecule photon−1 [28]. For the calculation of the quantum efficiency of ade- nine, a beam of UV radiation was focused on a thin com- pact adenine sample formed through sublimation and re- condensation [28]. In this case, all the photons in the beam of UV radiation were incident on the nucleobases. In the WLP scenario, not all the incoming UV photons will be incident on nucleobases. Instead, large gaps can exist between nucleobases that collect at the base of the pond. Therefore the number of photons incident on the pond area is not the same as those incident on the scat- tered nucleobases unless there are at least enough nu- cleobases present to cover the entire pond area. Since the nucleobases are mixed well into the pond water be- fore complete evaporation, we assume nucleobases will spread out evenly as they collect on the base of the pond. This means we assume that all locations on the base of the pond are covered in nucleobases before nucleobases stack on top of one another. For low abundances of to- tal nucleobases in a WLP, this approach will lead to a slightly higher estimate for the rate of photodissociation than expected. We deem this acceptable for a first-order estimate of nucleobase photodissociation. The mass of nucleobases photodestroyed per year, per area covered by nucleobases (i.e. the photodestruction flux), is constant over time and is dependent on the ex- perimentally measured quantum efficiency of photode- composition. Mi = ΦF λγµi hcNA [kg yr−1 m−2], (S30) where Φ is the quantum efficiency of photodecomposi- tion of the molecules (molecules photon−1), F is the UV flux incident on the entire pond area (W m−2), λ is the average wavelength of UV radiation incident on the sample (m), γ is 3600 · 24 · 365.25 s yr−1, µi is the molecular weight of the irradiated molecules (kg mol−1), h is Planck's constant (m2 kg s−1), c is the speed of light (ms−1), and NA is Avogadro's number (molecules mol−1). Our estimation of the photodestruction rate of a nucle- obase depends on the total mass of the nucleobase within the WLP. If there is enough of the nucleobase present for the entire base of the pond to be covered, we can mul- tiply the photodestruction flux by the entire pond area. Otherwise, we must multiply the photodestruction flux by the combined cross-sectional area of the nucleobase present in the WLP to get the photodestruction rate. − Mi mi , ρid − MiAp, dmi dt = if mi ρid < Ap [kg yr−1], (S31) otherwise where mi is the mass of the sample (kg), ρi is the mass density of the nucleobase (kg m−3), d is the distance between two stacked nucleobases in the solid phase (m), and Ap is the area of the WLP (m2). Assuming cloudless skies, an upper limit on the early Earth integrated UV flux from 200–250 nm is ∼0.4 W m−2 [35, 60]. UV wavelengths < 200 nm would be com- pleted attenuated by CO2 and H2O in the early atmo- sphere [35]. The mass density of solid adenine is 1470 kg m−3, making the distance between two stacked adenine molecules in the solid phase ∼6.6 A. For guanine, uracil, and cytosine, we take the mass densities to be 2200 kg m−3, 1320 kg m−3, and 1550 kg m−3, respectively. Seepage The constant pond water seepage S = 0.95 m yr−1 was determined in the pond water source and sinks sec- tion above. This can be used to calculate the nucleobase seepage rate via the equation: dmi dt = wiρwAp S [kg yr−1], (S32) where wi is the nucleobase mass fraction, ρw is the den- sity of water (kg yr−1), and Ap is the area of the WLP (m2). Nucleobase Outflow and Mixing Chondritic IDPs and meteorites are porous [61–63]. With the exception of the nucleobases potentially formed due to surface photochemistry [12], any soluble nucle- obases delivered to the prebiotic Earth by carbonaceous IDPs and meteorites would have layed frozen in the pores of these sources upon them entering the atmo- sphere. Pulse heating experiments show approximately ∼1–6% of the organics within IDPs could have survived atmospheric entry [62]. And, based on their presence in carbonaceous chondrites today, nucleobases in carbona- ceous meteorites evidently would have also survived at- mospheric entry heating. Therefore both sources, upon entering WLPs on the prebiotic Earth, would have slowly released their remaining soluble nucleobases into the sur- rounding pond water. These nucleobases would then slowly homogenize into the pond solution. We model the outflow of nucleobases from carbona- ceous IDPs and meteorites and the mixing of a local con- centration of nucleobases into a WLP using finite differ- ence approximations of the one-dimensional advection- diffusion equation (see Appendix A for complete details). Our model of nucleobase outflow is run for average- sized IDPs (r = 100µm), and for small (r = 1 cm), 15 medium (r = 5 cm) and large (r = 10 cm) carbonaceous meteorites. The fraction of nucleobases remaining in each of these sources as a function of time is plotted in Fig- ure S3 below. Our models show that the duration of nucleobase diffu- sion from carbonaceous IDPs and meteorites into WLPs is mostly determined by the radius of the source. For typical, 100 µm-radius carbonaceous IDPs laying at the bottom of a WLP, it takes < 2 minutes for > 99% of the soluble nucleobases to diffuse into the surrounding pond water. For 1 cm-radius carbonaceous meteorites, this time increases to 10 days. For the largest carbona- ceous meteoroid fragments, with radii of 5 cm and 10 cm, this duration increases to approximately 8 and 32 months, respectively. For nucleobase mixing, we model a base-to-surface con- vection cell within cylindrical ponds with equal radii and depths of 1 m, 5 m, and 10 m. This model gives us the timescale of mixing a local concentration of nucleobases within WLPs. The maximum percent local nucleobase concentration difference from the average is plotted as a function of time in Figure S4. This metric allows us to characterize the nucleobase homogeneity in a convection cell of the WLP. Our simulations suggest that the mixing of local de- posits of nucleobases in WLPs, resulting from their dif- fusion out of carbonaceous IDPs and meteorites, is a very efficient process. For a cylindrical WLP 1 m in depth, it will take about 35 minutes for a local deposition to ho- mogenize in a convection cell within the pond. For larger WLPs, with 5 m and 10 m depths, the nucleobase mix- ing time increases to 104 and 150 minutes, respectively. These short mixing times make it clear that, for carbona- ceous meteorites ≥ 1 cm in radius, nucleobase homoge- nization in WLPs is dominated by the nucleobase outflow time from these bodies. In contrast, being that carbona- ceous IDPs are much smaller than meteoroid fragments, the nucleobase outflow time from IDPs is negligible com- pared to the nucleobase homogenization time in WLPs. Nucleobase Evolution Equation From IDPs It is estimated that at 4 Ga, approximately 6 × 108 kgyr−1 of carbonaceous IDPs were being accreted onto the Earth [11]. Since IDPs are tiny (typically ∼100 µm in radius), they are circulated by the atmosphere upon ac- cretion and thus likely reached almost everywhere on the prebiotic Earth. Approximately 1–6% of the organic con- tent within IDPs could have survived the pulse heating of atmospheric entry [62]. Assuming IDPs accreted uni- formly, the surface nucleobase mass accretion per square area would be dmi dtdA = wi mI fs 4πR2⊕ [kg yr−1m−2], (S33) where wi is the nucleobase mass fraction within IDPs for nucleobase i, mI is the mass accretion rate of IDPs 16 FIG. S3. Fraction of the total initial nucleobases remaining in (A) a 100µm-radius IDP and (B) 1 cm-, 5 cm-, and 10 cm-radii meteorites over time as a result of diffusion across a rock-pond boundary. The IDP and meteorites are considered to be laying on the bottom of a WLP, and are diffusing nucleobases symmetrically in the radial direction. The times at which 99% of the initial contained nucleobases have diffused into the WLP are labeled on the plots. IDPs are thought to correspond to origins of asteroids or comets [11], therefore at best, the average nucleobase abundances in IDPs could match the average nucleobase content within the nucleobase-rich CM, CR, and CI me- teorites. The average abundances of adenine, guanine, and uracil in CM, CR, and CI meteorites are listed in Table S2 along with the weighted averages based on rel- ative fall frequencies. These abundances might be an upper limit for the guanine, adenine, and uracil content of IDPs because unlike the interior of large meteorites, molecule-dissociating UV radiation can penetrate every- where within µm–mm-sized pieces of dust. TABLE S2. Average guanine, adenine and uracil abundances (in ppb) in the CM, CR, and CI carbonaceous chondrites. Abundances obtained from [13]. Some CM and CR meteorite analyses found no adenine or uracil, these samples were ex- cluded from the average. Weighted nucleobase averages are also displayed based on relative fall frequencies [22]. Uracil has not been measured in CR meteorites. Cytosine has not been measured in any meteorites. Meteorite Type Guanine Adenine Uracil CM CR CI 183.5 1.8 81.5 Weighted Avg 141.3 69.8 9.3 60.5 60.7 50.0 - 73.0 48.6 The amount of cytosine that could have formed on the surfaces of primordial IDPs is not well constrained. Cy- tosine has been detected in experiments exposing IDP analogs to UV radiation, but hasn't been quantified [12]. Furthermore, these analog experiments formed cytosine via photoreactions involving pyrimidine: a molecule that FIG. S4. The nucleobase mixing time in a base-to-surface convection cell (length = 2rp) within 1 m-, 5 m-, and 10 m-deep WLPs, beginning from a local concentration at the base of the pond. Nucleobase mixing is measured using the maximum percent local nucleobase concentration difference from the average. The time at which the maximum local nucleobase concentration difference from the average drops to 10% is labeled on the plot for each pond size. At this time, we consider the nucleobases in the WLP to be well mixed. For WLPs with radii 1 m, 5 m, and 10 m, the convection cell nucleobase mixing times are 35 minutes, 104 minutes, and 150 minutes, respectively. on the prebiotic Earth (kg yr−1), and fs is the average fraction of nucleobases that survive pulse heating from atmospheric entry. 020406080100Time (s)0.00.20.40.60.81.0Fraction of nucleobases remaining84 secondsA100µm IDP1% Line100101102103Time (d)0.00.20.40.60.81.010 days240 days960 daysB1cm frags5cm frags10cm frags101102Time (minutes)101102103Max local concentration difference from average (%)35 minutes104 minutes150 minutesrp = 1mrp = 5mrp = 10m has no measured abundance in IDPs or meteorites. For this analysis we explore a best-case scenario, and set the maximum cytosine IDP abundance to 141.3 ppb (the pos- sible upper limit of guanine in carbonaceous IDPs). In the nucleobase outflow and mixing section above we learned that nucleobase diffusion from IDPs is quick (lasting < 2 minutes), and nucleobase homogenization in WLPs takes one to a few hours, depending on the pond size. If we spread out the 6 × 108 kgyr−1 accretion rate of IDPs uniformly across the surface of the Earth- assuming IDPs are all 100 µm-radius spheres of CI, CM, and CR type (ρ = ∼2185 kgm−3)-then IDPs would drop into 1–10 m-radius primordial WLPs at approximately 0.05–5 per hour. Since each carbonaceous IDP can only carry a tiny mass in nucleobases ( ∼1 picogram), nucle- obase inhomogeneities within WLPs would be a maxi- mum of ∼15 picograms. This abundance of nucleobases is negligible, therefore for our calculations of nucleobase accumulation in WLPs from IDP sources, we can assume that nucleobase deposition and pond homogenization is instantaneous. Thus, the differential equation for the mass of nucle- obase i accumulated in a WLP from IDP sources over time is the sum of the nucleobase mass accretion rate from IDPs, the nucleobase mass decomposition rate, the nucleobase mass seepage rate, and the nucleobase pho- todissociation rate: dmi,IDP (t) dt = wi mI fsAp 4πR2⊕  Mi mi , ρid MiAp, − − γmiki − wiρwAp S if mi ρid < Ap [kg yr−1]. otherwise (S34) The second and third terms after the equal sign are only activated when the pond is wet, and the fourth term is only activated when the pond is dry. The first term is always activated as recently accreted nucleobases are susceptible to UV dissociation while still laying in the pores of IDPs, and they effectively instantaneously out- flow from IDPs upon wetting. Using Equation S34, we compute the nucleobase mass as a function of time numerically using a forwards time finite difference approximation. We then divide the nu- cleobase mass by the water mass at each time step to obtain the nucleobase mass concentration over time. Be- cause some ponds are seasonally dry, we freeze the water level at 1 mm during the dry phase in order to calculate a nucleobase concentration during this phase. Nucleobase Evolution Equation From Meteorites Simulations show the fragments from carbonaceous meteoroids with diameters from 40–80 m and initial ve- locities of 15 km/s will expand over a radius of ∼500 m, and ∼ 32% of the original meteoroids will survive 17 ablation [32]. Since a single meteoroid impacting the at- mosphere may break up into 180,000 fragments before spreading across its strewnfield, it is probable that a meteorite deposition event involves multiple meteorites landing in a single WLP. The best estimate of the to- tal nucleobase mass deposited into a WLP is thus cal- culated assuming the ablated meteoroid mass is spread uniformly over its strewnfield. Considering this, the total nucleobase mass to deposit into a WLP from a meteoroid would be mi0 = 4 3 wifsr3ρAp r2 g [kg], (S35) where wi is the nucleobase mass fraction within the mete- oroid for nucleobase i, fs is the fraction of the meteoroid to survive ablation, r is the meteoroid radius as it enters Earth's atmosphere (m), ρ is the density of the meteoroid (kg m−3), rg is the radius of the debris when the mete- oroid fragments hit the ground (m), and Ap is the area of the WLP (m2). After the deposition of meteoroid fragments into a WLP, the frozen meteorite interiors will thaw to pond temperature, allowing hydrolysis to begin inside the frag- ments' pores. This means the total mass of nucleobases that diffuse from the fragments' pores into the pond will be less than the total initial nucleobase mass within the deposited fragments. By integrating the nucleobase hy- drolysis rate (Equation S29), we obtain the mass of nu- cleobase i remaining after a given time of hydration th, mi = mi0e−γkith [kg]. (S36) Note that th (yr) may be different than t, as the hydration clock is paused when WLPs are dry. Unlike carbonaceous IDPs, which unload their nucle- obases into a WLP in seconds, nucleobases may not com- pletely outflow from all deposited carbonaceous mete- oroid fragments before the WLP evaporates. (However, the nucleobases that do outflow from the meteorites will mix homogeneously into the WLP within a single day- night cycle.) Thus, we calculate the nucleobase outflow time constants for 1 cm-, 5 cm-, and 10 cm-radius car- bonaceous meteorites by performing least-squares regres- sions of our nucleobase diffusion simulation results to the function below. f (th) = α 1 − e − th τd . (S37) (cid:16) (cid:17) This equation shows that as time increases, the mass of nucleobases that have flowed out of a single meteorite, into the WLP, approaches the coefficient α-which rep- resents the total initial nucleobase mass contained in the meteorite. Since nucleobase outflow is mass-independent, we can use an arbitrary initial total nucleobase mass for α in our simulations to obtain the nucleobase outflow time constants. The results of the fits give diffusion time constants for 1 cm-, 5 cm-, and 10 cm-radius fragments of τd = 4.9 ×10−3 yr, 0.12 yr and 0.48 yr, respectively. 18 Adding up the sources and sinks gives us the nucle- obase mass within the WLP from meteorite sources as a function of time and hydration time. (cid:16) (cid:17) dmi,M et(t, th) dt = −th γki+ 1 τd e mi0 τd −  Mi mi , ρid MiAp, − γmiki − wiρwAp S mi ρid [kg yr−1]. < Ap if otherwise (S38) The first three terms after the equal sign are only acti- vated when the pond is wet, and the fourth term is only activated when the pond is dry. Since there are many possibilities for the sizes of me- teoroid fragments that will enter a WLP, we consider 3 simplified models: all fragments that enter a WLP from a meteoroid of radius 20–40 m are either 1 cm in radius, 5 cm in radius, or 10 cm in radius. These three models represent a local part of the strewnfield that deposited either many small fragments, mostly medium-sized frag- ments, or just a couple to a few large fragments. Cytosine is unlikely to have sustained within meteorite parent bodies long enough to be delivered to the early Earth by meteorites [15], therefore we only model the accumulation of adenine, guanine, and uracil in WLPs from meteorite sources. We solve Equation S38 numerically using a forwards time finite difference approximation. Nucleobase concen- tration is then obtained by dividing the nucleobase mass by the water mass in the WLP at each time step. Additional results Pond Water In Figure S5 we explore the effects of changing tem- perature on wet environment WLPs of 1 m radius and depth (see Table 1 in main text for wet environment model details). To do this, we vary our fiducial model temperatures (65◦C for a hot early Earth and 20◦C for a warm early Earth) by ±15◦C. As temperature increases, evaporation becomes more efficient, however the wet en- vironment WLPs never dry completely. Nucleobase Accumulation from IDPs In Figure S6 we explore the evolution of adenine con- centration from only IDP sources in WLPs of 1 m radius and depth. We model the adenine accumulation in three environments (dry, intermediate, and wet) on a hot and warm early Earth (see Table 1 in the main text for pre- cipitation model details). The left panel is for 65◦C on a hot early Earth and 20◦C on a warm early Earth, and the left panel is for 50◦C on a hot early Earth and 5◦C on a warm early Earth. FIG. S5. The effect of temperature on the change in water mass over time in wet environment cylindrical WLPs with radii and depths of 1 m. Temperatures are varied for a hot early Earth model (50 ◦C, 65 ◦C, and 80 ◦C) and a warm early Earth model (5 ◦C, 20 ◦C, and 35 ◦C). Precipitation rates from Columbia and Thailand on Earth today are used to represent the hot, and warm early Earth analogues, respec- tively (for details see Table 1 in the main text). All adenine concentration curves reach a stable sea- sonal pattern within ∼5 years. The highest adenine concentrations occur for models with a dry phase (i.e. the dry and intermediate models at 65◦C, and the dry model at 50◦C.) The lower maximum concentrations in the models without a dry phase are due to the sustained high water levels. The maximum adenine to accumulate in any model from IDP sources is ∼0.2 ppq. This occurs just before the pond dries. Upon drying, UV photodis- sociation immediately drops the adenine concentration to an amount which balances the incoming adenine from IDP accretion. The curves in Figure S6 do not change drastically with increasing pond radius and depth once a stable seasonal pattern is reached. This is because al- though ponds of increasing surface area collect more nu- cleobases, these ponds have an equivalent increase in area to nucleobase seepage. In Figure S7 we explore guanine, uracil, and cytosine accumulation in degenerate dry environment WLPs for a hot early Earth at 65 ◦C and a warm early Earth at 20 ◦C (See Table 1 in main text for details). The differences in each nucleobase mass fraction over time is caused by the different initial abundances of each nucleobase in IDPs (see Table S2). Although hydrolysis rates differ between nucleobases, decay due to hydrolysis is negligible over < 10 year periods at temperatures lower than ∼70 ◦C. In Figure S8 we turn off seepage, e.g. resembling a scenario where a lipid biofilm has covered the WLP base, and explore the evolution of adenine concentrations from IDP sources. This model is displayed for a hot early Earth at 65 ◦C and a warm early Earth at 20 ◦C (See Table 1 in main text for details). UV photodissociation is always the dominant nucleobase sink for the dry and 0246810Time (yr)0255075100Percent full (%)Wet Env.COL 50◦CTHA 5◦CCOL 65◦CTHA 20◦CCOL 80◦CTHA 35◦C050010001500200025003000Water Mass in WLP (kg) 19 FIG. S6. The accumulation of adenine from only carbonaceous IDP sources in cylindrical WLPs with radii and depths of 1 m. The three curves (dry, intermediate, and wet environments) differ by their precipitation rates, which are from a variety of locations on Earth today, and represent 2 classes of matching early Earth analogues: hot (Columbia, Indonesia, Cameroon), and warm (Thailand, Brazil, and Mexico) (for details see Table 1 in the main text). (A) The degenerate WLP models used for these calculations correspond to a hot early Earth at 65 ◦C and a warm early Earth at 20 ◦C. (B) The degenerate WLP models used for these calculations correspond to a hot early Earth at 50 ◦C and a warm early Earth at 5 ◦C. ronment, where hydrolysis takes over as the dominant nucleobase sink. FIG. S7. The accumulation of guanine, adenine, uracil, and cytosine from only carbonaceous IDP sources in cylindrical WLPs with radii and depths of 1 m. The degenerate dry WLP models used for these calculations correspond to a hot early Earth at 65 ◦C and a warm early Earth at 20 ◦C. Pre- cipitation rates from Cameroon and Mexico on Earth today are used to represent the hot, and warm early Earth ana- logues, respectively (for details see Table 1 in the main text). The cytosine abundance in IDPs used to calculate the max cytosine curve, matches the average abundance of guanine in IDPs (see Table S2). The curves are obtained by numerically solving Equation S34. intermediate environments, therefore for this model we only display the evolution of adenine in the wet envi- FIG. S8. The no seepage limit: The accumulation of ade- nine from only carbonaceous IDP sources in cylindrical WLPs with radii and depths of 1 m. The two curves represent the adenine concentrations in a wet environment pond on a hot (65 ◦C), and warm (20 ◦C) early Earth (for details see Ta- ble 1 in the main text). The thickness of the lines is due to the seasonal oscillations in adenine concentrations. In the absence of seepage, adenine concentrations can build up in wet environment WLPs until the rate of in- coming adenine from IDPs matches the decay rate due to hydrolysis. Hydrolysis rates are faster at hotter temper- 012345678Time (yr)10-310-210-1Adenine Mass Fraction (ppq)ADry Env.Intermediate Env.Wet Env.012345678Time (yr)10-310-210-1BDry Env.Intermediate Env.Wet Env.012345678Time (yr)10-310-210-1100Nucleobase Mass Fraction (ppq)Guanine/max CytosineAdenineUracil100101102103104105Time (yr)10-210-1100101102Adenine Mass Fraction (ppq)Wet Env.65◦C20◦C atures, therefore maximum adenine concentrations are higher and take longer to converge in the 20 ◦C pond compared to the 65 ◦C pond. However, these maximum adenine concentrations are 145 and 0.3 ppq, respectively, which are negligible in comparison to the ppb–ppm-level adenine concentrations reached in WLPs from carbona- ceous meteorite sources. Nucleobase Accumulation from Meteorites In Figure S9 we explore the evolution of adenine con- centration in WLPs with radii and depths of 1 m, from 1-cm fragments of an initially 40 m-radius carbonaceous meteoroid. The models correspond to degenerate envi- ronments on a hot (65 ◦C) and warm (20 ◦C) early Earth. The maximum adenine concentration in the intermedi- ate environment is ∼1.4 ppm, and occurs 16 hours after the fragments deposit into the nearly empty pond. The dry and wet environments allow for a maximum adenine concentration of ∼2 ppm. Because adenine outflow from 1 cm-sized meteorites occurs in just 10 days, only ade- nine sinks exist from 10 days onwards. As the ponds wet, the adenine concentrations decrease exponentially, and as the ponds dry again, the curves flatten out and ramp up slightly before the ponds completely dry up. When the ponds dry, UV radiation quickly wipes out the adenine at the base of the ponds. Since the wet environ- ment doesn't have a dry phase, the adenine concentration slowly diminishes in this model due mainly to seepage. The adenine mass fraction curves in Figure S9, over a 2 year period, do not change as pond radii and depths in- crease equally. This is because, although the mass of wa- ter in a WLP increases for larger collecting areas, larger pond areas also collect more meteorite fragments, which counterbalances the water mass and keeps the nucleobase concentration the same. In Figure S10 we explore how initial meteoroid radius affects the maximum concentration of adenine accumu- lated (from its fragments) in WLPs with radii and depths of 1 m. The maximum adenine concentration only dif- fers by at most a factor of 8 when varying the initial meteoroid radius from 20–40 m. This is because the nu- cleobase mass to enter a WLP scales with the meteoroid mass, i.e. ∝ r3. Finally, in Figure S11 we explore guanine and uracil ac- cumulation in intermediate and wet environment WLPs with radii and depths of 1 m. These models correspond to a hot early Earth at 65 ◦C and a warm early Earth at 20 ◦C. The small differences between each nucleobase mass fraction over time is due to the different initial nucle- obase abundances in the deposited meteorite fragments (see Table S2). Although each nucleobase has a different hydrolysis rate (see Equation S25), the decay of guanine, adenine, and uracil due to hydrolysis is negligible in < 10 years for temperatures < 65◦C. 20 Appendix A: Advection and Diffusion Model Advection and diffusion are the two main considera- tions of solute transport in water. Because nucleobases will diffuse out of the pores of carbonaceous IDPs and meteorites at a different rate than they will mix homoge- neously in the WLP, we separate our nucleobase trans- port model into two distinct parts. In part one, we model the outflow of nucleobases from carbonaceous IDPs and meteorites. In part two, we model the mixing of a local concentration of nucleobases into a WLP. Both parts of our simulation can be modeled with the advection-diffusion equation below, with either one or both RHS terms "turned on." φ ∂Ci ∂t = ∇ · [Def f∇Ci] − ∇ · [uCi] , (S39) where φ is the porosity of the medium, Ci is the mass concentration of the species, Def f is the effective diffu- sion coefficient (often in m2s−1), and u is the convective fluid velocity [64]. For a 1D case, where the diffusion coefficient and fluid velocity are constant along the simulated path, the advection-diffusion equation can be written as, φ ∂Ci ∂t = Def f ∂2Ci ∂r2 − u ∂Ci ∂r . (S40) For part one of our nucleobase transport model (the nucleobase outflow portion), we set u = 0. This is be- cause carbonaceous IDPs and meteoroid fragments are likely too small to attain noticeable interior pressure dif- ferences (thus the convective velocity within these bodies is probably negligible). We do not consider hydrolysis in our nucleobase trans- port model, as we only intend on estimating nucleobase outflow and mixing timescales from these models (rather than the nucleobases remaining after these processes). Since nucleobase decay is uniform within the carbona- ceous sources and WLPs, and is also very slow at WLP temperatures (t1/2 ∼ tens to hundreds of years [15]), it is not likely to affect the timescales of complete nucle- obase outflow from the source or homogenization within the WLP. A non-zero amount of nucleobases will decom- pose during diffusion from the source, and during mixing within the WLP. However, this is considered in the fi- nal calculations of nucleobase accumulation within WLPs from each source (see Sections and ). The advection-diffusion equation also does not include adsorption or formation reactions. However, for the dif- fusion of soluble nucleobases from small porous environ- ments which previously reached chemical equilibrium, the effects of these extra sources and sinks will proba- bly be minimal. Also, to adjust the diffusion equation for a free-water medium, one simply needs to set φ = 1, and Def f = Df w. The effective diffusion coefficient of a species is pro- portional to, but smaller than its free water diffusion 21 FIG. S9. The accumulation of adenine from 1 cm fragments of an initially 40 m-radius carbonaceous meteoroid in a cylindrical WLP with a radius and depth of 1 m. The degenerate WLP models used for these calculations correspond to a hot early Earth at 65 ◦C and a warm early Earth at 20 ◦C. (A) The accumulation of adenine from carbonaceous meteorites in an intermediate environment (for details see Table 1 in the main text). The wet-dry cycles of the pond are also shown to illustrate the effect of water level on adenine concentration. (B) The three curves (dry, intermediate, and wet environments) differ by their precipitation rates, which are from a variety of locations on Earth today, and represent 2 classes of matching early Earth analogues: hot (Columbia, Indonesia, Cameroon), and warm (Thailand, Brazil, and Mexico) (for details see Table 1 in the main text). The curves are obtained by numerically solving Equation S38. coefficient. Many equations exist for modeling the effec- tive diffusion coefficient in porous media [65–68]. These equations depend on variables such as the porosity, tortu- osity, and constrictivity of the medium, which represent the void space fraction, the curves in the pores, and the bottleneck effect, respectively. These equations are listed in Table S3. Carbonaceous meteorites of type CM, CR, and CI have average porosities of 24.7%, 9.5%, and 35.0% [70]. Based on the relative fall frequency of these meteorites on Earth [22], the weighted average porosity of these meteorite types is ∼25%. Chondritic IDPs have similar porosities to carbonaceous chondrites [61], therefore a 25% porosity may also well represent nucleobase-containing IDPs. The constrictivity of a porous medium is only impor- tant when the size of the species is comparable to the diameter of the pores [68]. Therefore given that nucle- obases are < 1 nm in diameter and the bulk of pore diameters in, for example, the Acfer 094 carbonaceous chondrite, range from 20–200 nm [63], we can neglect δ from the listed models. Tortuosities of chondritic meteorites have an aver- age value of 1.45 [71], and the empirical exponent m for carbonaceous meteorites might be similar to that of nearshore sediments with a value of 2 [68]. Finally, the free water diffusion coefficient of a sin- gle nucleobase has not been measured, however the free water diffusion coefficient of a single nucleotide is 400 µm2s−1 [69]. Since nucleotides are heavier than nucle- obases by a ribose and phosphate molecule, they will likely diffuse slower than nucleobases. Therefore 400 FIG. S10. The maximum concentration of adenine accumu- lated from 1 cm fragments of a carbonaceous meteoroid 20–40 m in radius in cylindrical WLPs with radii and depths of 1 m. The degenerate WLP models used for these calculations cor- respond to a hot early Earth at 65 ◦C and a warm early Earth at 20 ◦C. The three curves (dry, intermediate, and wet envi- ronments) differ by their precipitation rates, which are from a variety of locations on Earth today, and represent 2 classes of matching early Earth analogues: hot (Columbia, Indone- sia, Cameroon), and warm (Thailand, Brazil, and Mexico) (for details see Table 1 in the main text). The curves are obtained by numerically solving Equation S38. 0.00.51.01.52.0Time (yr)10-410-310-210-1100101102103Adenine Mass Fraction (ppb)AIDN 65 ◦CBRA 20 ◦CWater Level0.00.51.01.52.0Time (yr)10-410-310-210-1100101102103Adenine Mass Fraction (ppb)BWet Env.Intermediate Env.Dry Env.0500100015002000250030003500Water Mass in WLP (kg)Dry2025303540Initial Meteoroid Radius (m)05001000150020002500Max Adenine Mass Fraction (ppb)Wet Env.Intermediate Env.Dry Env. TABLE S3. Different equations for modeling the effective diffusion coefficient of species in porous media. Estimates of the effective diffusion coefficients of single nucleobases through the pores of carbonaceous IDPs and meteorites are also calculated for each model. The free water diffusion coefficient, D0, represents the unobstructed diffusion of a species, the porosity factor, φ, represents the void fraction of the medium, the constrictivity factor, δ, represents the bottleneck effect due to small pore diameters, and the tortuosity factor, τ , represents the restiction in diffusive flow due to curves in the pores. Estimates of these factors, and the empirical exponent, m, for carbonaceous IDPs and meteorites are, D0 = 4×10−10m2s−1, φ = 0.25, δ = 1, τ = 1.45, and m = 2 [68–71]. 22 Source Estimate for this work (×10−11 m2s−1) 6.90 4.76 7.27 2.50 Using the above estimates, we calculate the effective diffusion coefficients for nucleobases in carbonaceous me- teorites and IDPs using each of the four models and dis- play them in their respective columns in Table S3. The average value of the effective diffusion coefficient across all four models is 5.36 ×10−11 m2s−1. As previously stated, convective velocity within the pores of carbonaceous IDPs and meteorites is consid- ered negligible. However, this is not the case within 1–10 m-radius WLPs. Due to the day-night cycles of the Earth, WLPs likely experienced a temperature gra- dient from the atmospherically exposed top of the pond to the constant geothermally heated base. Since con- vection is likely the dominant form of heat transport within hydrothermal ponds [72], convection cells would have formed, with warm (higher pressure) parcels of wa- ter flowing upwards and recently cooled (lower pressure) parcels flowing downwards. The convective fluid velocity can be estimated with the equation u =(cid:112)gβ∆T L [m s−1], (S41) where g is the gravitational acceleration experienced by the fluid (m s−2), β is the fluid's volumetric thermal ex- pansion coefficient (K−1), and ∆T is the temperature difference (K) over a scale length L (m) [73]. The volu- metric thermal expansion coefficients for water at 50, 65, and 80 ◦C, are 4.7, 5.6, and 6.5 × 10−4 K−1 respectively [74]. Small ponds and even lakes can experience a temper- ature difference of 1–5 ◦C over the course of a day-night cycle [75]. However, convection begins cycling water well before temperature differences of this magnitude are reached. To estimate the lower bound of a constant tem- Def f φδ τ Df w Saripalli et al. [65] φδ τ 2 Df w van Brakel and Heertjes [66] 2φ 3 − φ φmDf w Boving and Grathwohl [68] Car´e [67] Df w Df w = free water diffusion coefficient φ = porosity δ = constrictivity τ = tortuosity m = empirical exponent FIG. S11. The accumulation of guanine, adenine, and uracil from 1 cm fragments of an initially 40 m-radius carbonaceous meteoroid in cylindrical WLPs with radii and depths of 1 m. The degenerate WLP models used for these calculations correspond to a hot early Earth at 65 ◦C and a warm early Earth at 20 ◦C. The two curves for each nucleobase differ by their precipitation rates, which create intermediate (solid lines), and wet (dotted lines) environments, and are from a variety of locations on Earth today representing 2 classes of matching early Earth analogues: hot (Columbia, Indonesia), and warm (Thailand, Brazil) (for details see Table 1 in the main text). The curves are obtained by numerically solving Equation S38. µm2s−1 is a good estimate of the lower limit of the free water diffusion coefficient of a single nucleobase. 0.00.51.01.52.0Time (yr)10-310-210-1100101102103104Nucleobase Mass Fraction (ppb)Guanine Wet Env.Guanine Inter. Env.Adenine Wet Env.Adenine Inter. Env.Uracil Wet Env.Uracil Inter. Env. perature difference between the surface and the base of a WLP during a day-to-night period, we assume that each convection cycle cools a surface parcel of water by ∆T . Then, given ∆T , we match the corresponding cy- cle length and number of cycles with the approximate 12 hour period required for a 1 ◦C change in pond temper- ature. The equation for this calculation is summarized below. (cid:19) 2 3 (cid:18) 2LTc √ Pdn gβL ∆T = [K], (S42) where Tc is the change in pond temperature over a day- to-night period (K) and Pdn is the duration of that period (s). Using Equation S42, the minimum constant temper- ature difference between the surface and the base of a cylindrical WLP with a radius and depth of 1 m, at 65 ◦C is ∼0.007 K. For a WLP with a radius and depth of 10 m, this minimum constant temperature difference increases to ∼0.016 K. However, given that smaller ponds experi- ence greater temperature changes than larger ponds due to faster heat transfer, a ∆T of ∼0.01 K may be a rea- sonable lower-bound estimate for all WLPs in the 1–10 m-radius/depth range. Given a constant temperature difference of 0.01 K be- tween the base and surface of 1 m-, 5 m-, and 10 m-deep WLPs at around 65 ◦C, the convective flow velocities would be approximately 0.7, 1.7, and 2.4 cms−1, respec- tively. Due to the 1D nature of our simulations, we are as- suming a radially symmetric outflow of nucleobases from spherical carbonaceous IDPs and meteorites. We also assume that local concentrations of nucleobases recently flowed out of these sources will remain within a single convection cell. Lastly, we assume that the nucleobase homogenization timescale within a 1D convection cell of a WLP is mostly representative of the nucleobase ho- mogenization timescale within the entire WLP. Though the 1D handling of this part of our model is a simplifi- cation of advection and diffusion within WLPs, since we are only attempting to estimate nucleobase homogeniza- tion timescales to within a few factors, a 1D model is probably sufficient. For both model parts we use a backward time, cen- tered space (BTCS) finite difference method for the dif- 23 fusion term in the advection-diffusion equation. For part two of the model, we use the upwind method to approxi- mate the additional advection term. The BTCS method was selected over the more accurate Crank-Nicolson (CN) method based on the former's stability for sharply edged initial conditions and convergence for increasing levels of refinement. However, differences in diffusion timescales are found to be within rounding error upon comparison of these methods for a 100 µm carbonaceous IDP. The upwind method was selected over higher-order advec- tion approximation methods (e.g. Beam-Warming, Lax- Wendroff) due to its lack of spurious oscillations, lowest error accumulation in mass conservation tests and con- vergence for increasing levels of refinement. These two models are summarized in Table S4 below. For part one of our nucleobase transport model, the simulation frame starts at the center of the IDP or me- teorite, and ends at the rock-pond interface. The left ∂Ci ∂t (r = 0) boundary is Neumann (i.e. = 0 across the ∂Ci ∂t ary is open (i.e. before the boundary equals boundary), simulating zero inflow, and the right bound- ∂Ci ∂t after the boundary), simulating outflow into the WLP. The nucleobase content of the modeled sources are ini- tially homogeneous, but drop sharply to zero at the open boundary (i.e. the initial condition represents an expo- nentially smoothed step function). This is made to rep- resent the searing of the outermost layer of carbonaceous IDPs and meteorites from atmospheric entry heating. For part two, the simulation frame is an eccentric 1D convection cell, which loops between the bottom and the top of the WLP (length = 2rp). The convection cell is modeled with cyclic boundaries, i.e., continuous nucle- obase flow, and no nucleobases exit the convection cell. The initial concentration of nucleobases is a sharp Gaus- sian beginning at the base of the pond. The width of the spike is ∼1% of the WLP's radius. Example simulations, including initial conditions, for parts one and two of our nucleobase transport model are plotted in Figure S12. Because the time it takes for com- plete nucleobase outflow from a source, or complete nu- cleobase homogenization in a WLP, is independent of ini- tial nucleobase abundance, the solution to the diffusion equation at each grid point is displayed as a fraction of the total initial nucleobase concentration. For ease of viewing, part two of the model is displayed in the con- vection cell's moving frame, with coordinate r(cid:48) = r - u∆t. 24 TABLE S4. Summary of parts one and two of our 1D nucleobase transport model. Part one is a model of nucleobase outflow from carbonaceous IDPs and meteorites while they lay at the base of a WLP. Part two is the mixing (i.e. homogenization) of a local concentration of nucleobases throughout the WLP. Part Model Description Num. Method Boundaries Initial condition φ Def f (m2s−1) Nucleobase outflow from carbonaceous IDPs and meteorites BTCS Neumann and open See Figure S12a 0.25 5.36 ×10−11 Nucleobase mixing BTCS and in WLPs upwind Cyclic See Figure S12b 1 4.0 ×10−10 1 2 FIG. S12. Example simulations, including the initial conditions, for our two part nucleobase transport model. Each line represents a different snapshot in time. (A) Initially homogeneous nucleobase diffusion from a 100 µm-radius carbonaceous IDP. (B) Initially locally concentrated nucleobase mixing in a convection cell within a 1 m-deep WLP. The convection cell is a L = 2rp eccentric loop flowing between the bottom to the top of the WLP. This loop is sliced at r(cid:48) = 1 m in the convection cell's moving frame and unraveled for display in the 1D plot in B. 020406080100r (µm)0.0000.0010.0020.0030.0040.005Fraction of total initial nucleobase concentrationA) Outflow From IDP0 sec10 sec20 sec80 sec1.01.50.00.51.0r′ = r − u∆t (m)0.000.020.040.060.080.10u = 0.7 cms−1B) Mixing in WLP0 sec30 sec3 min30 min
1902.08022
1
1902
2019-02-21T13:13:01
Hydrogen Cyanide in Nitrogen-Rich Atmospheres of Rocky Exoplanets
[ "astro-ph.EP" ]
Hydrogen cyanide (HCN) is a key feedstock molecule for the production of life's building blocks. The formation of HCN in an N$_2$-rich atmospheres requires first that the triple bond between N$\equiv$N be severed, and then that the atomic nitrogen find a carbon atom. These two tasks can be accomplished via photochemistry, lightning, impacts, or volcanism. The key requirements for producing appreciable amounts of HCN are the free availability of N$_2$ and a local carbon to oxygen ratio of C/O $\geq 1$. We discuss the chemical mechanisms by which HCN can be formed and destroyed on rocky exoplanets with Earth-like N$_2$ content and surface water inventories, varying the oxidation state of the dominant carbon-containing atmospheric species. HCN is most readily produced in an atmosphere rich in methane (CH$_4$) or acetylene (C$_2$H$_2$), but can also be produced in significant amounts ($> 1$ ppm) within CO-dominated atmospheres. Methane is not necessary for the production of HCN. We show how destruction of HCN in a CO$_2$-rich atmosphere depends critically on the poorly-constrained energetic barrier for the reaction of HCN with atomic oxygen. We discuss the implications of our results for detecting photochemically produced HCN, for concentrating HCN on the planet's surface, and its importance for prebiotic chemistry.
astro-ph.EP
astro-ph
Hydrogen Cyanide in Nitrogen-Rich Atmospheres of Rocky Exoplanets P. B. Rimmera, S. Rugheimerb,c aUniversity of Cambridge Department of Earth Sciences, Downing St, Cambridge CB2 3EQ and University of Cambridge, Cavendish Astrophysics, JJ Thomson Ave, Cambridge CB3 0HE and MRC Laboratory of bCentre for Exoplanet Science, University of St. Andrews, School of Earth and Environmental Sciences, Molecular Biology, Francis Crick Ave, Cambridge CB2 OQH cUniversity Oxford, Atmospheric, Oceanic, and Planetary Physics Department, Clarendon Laboratory, Irvine Building, North Street, St. Andrews, KY16 9AL, UK Sherrington Road, Oxford, OX1 3PU Abstract Hydrogen cyanide (HCN) is a key feedstock molecule for the production of life's building blocks. The formation of HCN in an N2-rich atmospheres requires first that the triple bond between N≡N be severed, and then that the atomic nitrogen find a carbon atom. These two tasks can be accomplished via photochemistry, lightning, impacts, or volcanism. The key requirements for producing appreciable amounts of HCN are the free availability of N2 and a local carbon to oxygen ratio of C/O ≥ 1. We discuss the chemical mechanisms by which HCN can be formed and destroyed on rocky exoplanets with Earth-like N2 content and surface water inventories, varying the oxidation state of the dominant carbon-containing atmospheric species. HCN is most readily produced in an atmosphere rich in methane (CH4) or acetylene (C2H2), but can also be produced in significant amounts (> 1 ppm) within CO-dominated atmospheres. Methane is not necessary for the production of HCN. We show how destruction of HCN in a CO2-rich atmosphere depends critically on the poorly-constrained energetic barrier for the reaction of HCN with atomic oxygen. We discuss the implications of our results for detecting photochemically produced HCN, for concentrating HCN on the planet's surface, and its importance for prebiotic chemistry. 1. Introduction Hydrogen cyanide (HCN) is a key molecule for prebiotic photochemistry. HCN is a precursor molecule to amino acids via Strecker synthesis, where HCN is reacted with ammonia and an aldehyde to form various amino acids (Strecker, 1854), and has for this reason been invoked as the mechanism for generating the amino acid glycine in the Miller-Urey experiment (Miller, 1957). The structure of the nucleobase adenine (C5H5N5) is effectively five connected HCN molecules. Adenine can be formed by Email addresses: [email protected] (P. B. Rimmer), [email protected] (S. Rugheimer) Preprint submitted to Elsevier February 22, 2019 HCN reacting with its conjugate base and polymerizing, followed by photochemical rearrangement of the resulting HCN tetramer and further HCN polymerization (Ferris and Orgel, 1966; Sanchez et al., 1966, 1967; Sanchez and Orgel, 1970). More recently, HCN has been discovered to be present at each of the initial photo- chemical reactions that produce lipids, amino acids and nucleosides, the three building blocks of life (Ritson and Sutherland, 2012; Patel et al., 2015; Xu et al., 2018). In this scenario, UV light is absorbed by hydrogen cyanide and hydrogen sulfide or bisulfite in water. The UV photon initiates photodetachment of an electron from the hydrogen sulfide or bisulfite, and the solvated electron reduces the hydrogen cyanide, forming an imine complex that, when hydrolised, forms formaldehyde. This photochemical homologation can continue, with the dissociated products of cyanide reacting with the product sugars. Other plausible feedstock molecules generated from reactions involving HCN, such as cyanamide and cyanoacetylene, react with the sugars to form the pyrimidine nu- cleotides. In addition, these feedstock molecules, along with methylamine and nitro- gen oxides, act as activating agents to facilitate the ligation of these building blocks into polypeptides, a key next step in prebiotic chemistry (Mariani et al., 2018). HCN can be produced photochemically (Zahnle, 1986; Tian et al., 2011), by in- teraction of the atmospheric gas with energetic particles (Airapetian et al., 2016), by lightning (Chameides and Walker, 1981; Ardaseva et al., 2017), and by meteor impacts (Ferus et al., 2017). In each of these scenarios, the authors have recognized that re- duced carbon-bearing species greatly facilitate the production of hydrogen cyanide. In some circumstances, especially with impacts, up to 1% of the affected gas can be con- verted into HCN, but again only with a relatively reduced gas, and in the presence of significant concentrations of molecular nitrogen or ammonia. The same events occur- ring in a gas of different composition, however, can result in trace amounts of HCN, less than 1 ppb. Exoplanets can be seen as diverse laboratories for atmospheric and prebiotic chem- istry, and so we need not be bound by constraints from early Earth atmosphere. Bulk atmospheric composition is given for rocky planet atmospheres under a wide variety of elemental abundance ratios (Hu and Seager, 2014). In addition, changes in the water inventory (Gao et al., 2015), the amount of hydrogen retained in the atmosphere (Pierre- humbert and Gaidos, 2011), and the concentration of molecular nitrogen (Wordsworth and Pierrehumbert, 2014), can have profound effects on the evolution of the planet's atmosphere. Here we will focus on more Earth-like planets, in terms of their nitrogen content and water inventories. Given the presence of molecular nitrogen as a dominant atmospheric constituent, the bulk atmospheric composition will here be represented largely in terms of the C/O ratio (see Fig. 1). We present our methods for calculating the climate and photochemistry in Section 2, and then discuss the connection between the C/O ratio, the relative UV flux and HCN in Section 3. In Section 4 present the concentrations of HCN as a function of the C/O ratio, UV flux, and the barrier to destruction by atomic oxygen. We conclude with a discussion of these results (Section 5). 2 Figure 1: A representation of how the bulk chemistry changes with the bulk C/O ratio, and how more-or-less local chemical processes: lightning, impacts, volcanism, ultraviolet photons, are affected by the C/O ratio, from C/O < 1 to C/O > 1. 2. Methods To calculate the temperature and pressure profiles, we use a 1D climate model de- veloped for high-CO2/high-CH4 terrestrial atmospheres (Pavlov et al., 2000; Kharecha et al., 2005; Haqq-Misra et al., 2008). This model simulates an anoxic 1 bar atmo- sphere composed of 89% N2, 10% CO2 and 1% CH4, similar to some of the models of Zahnle (1986).Gases apart from H2O are assumed to be well mixed in the atmosphere. The atmosphere was divided into 101 layers. The code uses a δ two-stream scatter- ing approximation for the absorption of solar radiation (Toon et al., 1989) and 4-term, correlated k-coefficients to parameterize the absorption by the main greenhouse gases, namely: H2O, CH4, and CO2. We use the solar evolution model for the incoming stel- lar radiation at 3.9 Ga (Claire et al., 2012), both for the climate modelling and the UV spectrum for the photochemistry discussed below. The temperature profile from the 1D climate model is in Fig. 2. We take the prescribed stellar spectra and surface mixing ratios, and the temper- ature profile determined by the climate model, and apply them to the ARGO photo- chemistry/diffusion model (Rimmer and Helling, 2016), which solves the set of photo- chemical kinetics equations for a moving parcel of atmospheric gas: ∂ni ∂t = Pi − Li − ∂Φi ∂z , (1) Here, ni [cm−3] is the number density of species i, where i = 1, ...,Is, Is being the total number of species. The quantity Φi is the flux of the species i into and out of the parcel when it is at height z. The parcel is used to give a chemical profile of the atmosphere, and UV photo-transport is calculated in order to determine the efficiency of the pho- tochemistry at each parcel height (Rimmer and Helling, 2016). Pi [cm−3 s−1] and Li 3 Figure 2: The temperature [K] as a function of altitude [km] for an anoxic 1 bar surface pressure atmosphere composed of 89% N2, 10% CO2 and 1% CH4 (black line), and with updated mixing ratios taken from the ARGOS photochemical model using the STAND2019 chemical network (red line). 4 150200250300350Temp (K)020406080Alt (km)Initial TP profileUpdated mixing from ARGOS [cm−3 s−1] represent the production and loss rates, which are determined by using the a modified version of the STAND2016 chemical network (Rimmer and Helling, 2016), which we hereafter will refer to as STAND2019. The network is described in detail below. This model will determine a different chemical profile from the climate model, but does not update the temperature profile. The deviation between the temperature profile for the initial and updated mixing ratios is small, and is shown in Fig. 2. The STAND2019 chemical network is a modification of the STAND2016 network, which includes a robust H/C/N/O chemistry with species of up to 3 carbons, 2 nitro- gens, 3 oxygens, and 8 hydrogens, valid for temperatures up to 30000 K, and including some relevant gas-phase reactions involving Na, K, Mg, Fe, Ti, Si, and Cl. Altogether, the STAND2019 chemical network comprises over 5000 reactions involving over 350 chemical species. The full network is included in the Supplementary Information. We now describe the most significant differences between STAND2016 and STAND2019. The reverse of a reaction relevant for formaldehyde in STAND2016 has been sup- pressed in the STAND2019 network in order to properly reproduce the formaldehyde (HCHO) abundances observed in Earth's atmosphere today (Granville-Willett et al., 2018). The relevant reaction is: H + CO + M → HCO + M. (2) and its reverse is: HCO + M → H + CO + M. (3) The reverse rate is calculated using the Gibbs free energy, as described by Rimmer and Helling (2016). This is so that the chemical steady state converges to thermochemical equilibrium at sufficiently long timescales in the absence of irreversible reactions (e.g. photochemistry). This is important for reproducing hot exoplanet atmospheric compo- sitions and the gas-phase component of magma chemistries, both applications for the STAND2019 network. The consequence of this approach is that sometimes the barrier calculated by reversing the reaction does not reproduce the actual chemical barrier to the reaction, and in the case of R3, the barrier is greatly underestimated. In order to compensate for this, the rate constant for R3 is set to: k = 6.0× 10−11 e−5370 K/T cm3 s−1, and using this rate constant, with this sizable barrier, reproduces the the formaldehyde (HCHO) abundances (Granville-Willett et al., 2018). The forward and reverse barriers for this reaction will need to be revisited in the future to figure out why the barrier for R3 is considerably higher than determined when reversing R2. In addition, the relevant reaction from STAND2016 has been replaced with R7: HCN + O → NCO + H, with the rate constant of: k = 1.43× 10−12 (cid:18) T (cid:19)1.47 300K 5 e−3770 K/T cm3 s−1, (4) from Perry and Melius (1985). This choice of rate constant will be justified in Section 4.2. Finally, we have added the reaction (Becker and Hong, 1983): HCN + C2H → HC3N + H, for which we assign a rate constant of: k = 2.2× 10−12 cm3 s−1. (5) (6) This reaction was added to see if cyanoacetylene might be generated in trace but sig- nificant quantities in our models, which would be relevant for prebiotic chemistry, as acyanoacetylene is one of the key feedstock molecules for forming some of the canon- ical amino acids and the pyrimidine nucleotides. It turns out that no significant quanti- ties of cyanoacetylene are generated in any of the atmospheric models we consider. Changes to the methanol chemistry (Tsai et al., 2017), are included, and not ex- pected to make much difference here. The STAND2019 network, which appears first in this paper with new reaction numbers, is included in full in the Supplementary In- formation. The number of reactions is large (over 5000), but only a relatively small number of reactions are relevant for the HCN chemistry. The reactions relevant for atmospheric HCN production, by UV photons as well as by impacts, lighting and ener- getic particle interactions, are discussed in detail in Section 3. These reactions depend on the abundances of other species, which are determined by yet other reactions, and so we use the entire network when modeling the atmospheric photochemistry. Finally, for the alterations in the stellar spectrum in Section 4.1, we take a model spectrum (Claire et al., 2012), and artifically enhance or reduce the spetrum within the regions of 500 Å- 1000 Å and 1000 Å- 2000 Å, as shown in Fig. 3, to test the impact of the UV spectrum on HCN formation. 3. Theory In an exoplanet atmosphere composed of molecules bearing hydrogen, carbon, oxy- gen and nitrogen, there are two necessary conditions for producing HCN. First, the nitrogen bonds need to be cleaved, and then second, carbon needs to win over oxygen in the competition for the free nitrogen. The first condition can be relatively easy if the nitrogen is primarily in the form of ammonia (NH3), which has a large absorption cross section over a wider wavelength range than N2 ((cid:46) 300 nm for ammonia compared to (cid:46) 160 nm for N2). Hot temperatures can also be sufficient to cleave a hydrogen atom from ammonia. Nitrogen in the form of N2, with a strong triple bond, is far more difficult to break apart. Impacts and lightning are sufficient to dissociate N2 into N + N, as are UV photons of sufficiently short wavelength ((cid:46) 110 nm), although these are largely attenuated below the thermosphere by N2 itself, assuming that N2 is a prevalent atmospheric constituent. After the nitrogen bonds are broken, the nitrogen atoms react readily with nearby molecules, including some otherwise stable molecules. The simplest (and often the most common) case is if a lone nitrogen atom finds another lone nitrogen atom, in which case it will react to form N2. If there is a great deal of hydrogen available, 6 Figure 3: The spectrum of the young Sun (solid line) (Claire et al., 2012), artifically enhanced between 500 Å- 1000 Å and suppressed between 1000 Å- 2000 Å(dashed line), used to generate the results in Section 4.1. 7 600800100012001400160018002000 [Å]10610710810910101011101210131014F [cm2 s1 Å1] whether in the form of H or H2, the nitrogen will rapidly react to form NH or NH2, and then will continue to react with the abundant hydrogen to end in ammonia. On the other hand, if the nitrogen atom collides with an oxygen atom, the resulting eventual product will generally be NO or NO2. If the nitrogen atom finds carbon, the eventual product will generally be HCN. Thus the balance of carbon to oxygen in a given geological environment is a key metric to understanding abiotic nitrile chemistry. Fig's 4 and 5 provide the reaction networks for the formation and destruction of HCN under a variety of circumstances. O, OH, or high T hν H2 CH4 hν CH CH2 hν+H2 N2O NO N CO2 hν high T CO N (high T) CN H,H2 H2O HCN high T hν NO H H C NCO N2 CNN H2 CH4 N H2O CH3 CH3 C2H6 C2N Figure 4: Collection of known pathways to form HCN on rocky planets. HCN can be produced photochem- ically (this paper, as well as Zahnle 1986 and Tian et al. 2011), by impacts (Ferus et al., 2017), and by lightning (Ardaseva et al., 2017). The rate constants for these reactions can be found in the STAND2019 Network (Supplementary Information). ACN A+ CN− CH3O− HCN hν CN N,NO N2 OH O O CO HNCOH NCO N O NO OH hν HNCO Figure 5: Collection of known pathways to destroy HCN on rocky planets. Destruction is primarily via atomic oxygen, the hydroxyl radical, anion chemistry, or photodissociation of HCN followed by reaction with atomic nitrogen or NO. The rate constants for these reactions can be found in the STAND2019 Network (Supplementary Information). There are two ways to look at this competition: in terms of equilibrium and kinet- ics. Chameides & Walker were first to examine the thermochemical equilibrium of NO and HCN. They discuss the competition between carbon and oxygen for the nitrogen (Chameides and Walker, 1981), and express this competition in terms of the oxida- tion state of the carbon, where if the available carbon is mostly in the form of carbon 8 monoxide (CO) and methane (CH4), more HCN will result, and if it is mostly in the form of CO2, more NO will result. Chameides and Walker (1981) show this relation explicitly in terms of the C/O ratio in their Fig. 3. We examine the thermochemical formation of HCN as a function of the C/O ratio, and find that we reproduce the results of Chameides and Walker (1981). Where C/O < 1 and the nitrogen bonds are broken, NO is formed in large quantities and HCN in trace quantities. If the C/O > 1 and the nitrogen bonds are cleaved, the opposite holds, and a great deal of HCN is formed with very little NO. If C/O ≈ 1, then both HCN and NO are formed in reasonable quantities. These results match those of Chameides and Walker (1981, their Fig. 3). Methane is not necessary. All of the carbon could be in the form of CO or C2H2. What is necessary is an elemental C/O ratio ≥ 1 and a reasonable amount of hydrogen, whether bound as methane, acetylene, molecular hydrogen, or even water vapor. 3.1. Mechanisms for Generating and Destroying HCN on Rocky Exoplanets The mechanisms by which HCN is formed and destroyed vary with atmospheric composition, temperature, stellar activity, impact rate, surface processes, and lighting rates. The chemical mechanisms themselves do not change for each process, so it is helpful to look at the chemistry as a whole. Fig. 4 illustrates the pathways that can form HCN from CO2, CO, C2N (the most abundant CN-containing species produced by lightning on the Early Earth (Ardaseva et al., 2017)), as well as the formation from CH4 (Zahnle, 1986; Tian et al., 2011; Airapetian et al., 2016). Some of the reactions are photochemical, denoted by hν, and require photons of a given range of wavelengths, and these may involve reactive intermediates, e.g. H2CN, in the case of CH2 + N. Other reactions are driven by energetic particles, and some proceed at high temperatures. These driving factors are rarely exclusive. On some exoplanets, especially those with high C/O ratios, there may be a sig- nificant amount of carbon monoxide present in the atmosphere to begin with. For volcanically degassed atmospheres, the composition will largely depend on the out- gassing pressure and redox state of the crust and upper mantle (Gaillard and Scaillet, 2014). Otherwise, CO is produced in reasonable quantities by the photodissociation or thermal dissociation of CO2 (< 195 nm, Ea ∼ 415 kJ/mol). At very high temperatures (Ea = 300 kJ/mol), CO + N → CN + O, and once temperatures drop, the CN can pro- ceed to HCN. CO itself can be dissociated (< 110 nm for CO(X1Σ+), < 210 nm for CO(A3Π) , Ea = 1000 kJ/mol). The destruction of HCN (see Fig. 5) proceeds on both the present and Early Earth primarily via oxidation by atomic oxygen and hydroxyl radicals (e.g. R7). In environ- ments like Titan, with very high C/O ratios, where there is virtually no available oxy- gen, HCN is destroyed by photodissociation into CN followed by electron exchange from anions to form CN−. HCN can also react with the hydroxyl radical and a third body in order to form HNCOH, a radical precursor to formamide, potentially relevant to prebiotic chemistry (Saladino et al., 2001; Ferus et al., 2015). Most often, though, HNCOH will encounter a second OH radical, forming isocyanic acid (HNCO). Isocyanic acid is quite susceptible to photodissociation, resulting in H + NCO. 9 Finally, HCN can become absorbed into water droplets, and there has the oppor- tunity to react with a solvated electron, leading to the formation of formaldehyde, or its conjugate base can react with formaldehyde to form glycolonitrile. These are the initial steps toward RNA precursors (Ritson and Sutherland, 2012). The Henry's Law constant for HCN is about 10 M/bar, and so resulting concentration of HCN in rain- drops from the model atmosphere we present blow (C/O ratio of 0.55) is at most 10 nM, which is too small a concentration to be useful for prebiotic chemistry. If the atmospheric C/O ratio approaches 1, the surface mixing ratios could become as high as 10 ppm, and the resulting concentrations at 0.1 mM, on the cusp of prebiotically useful concentrations. The concentrations could be enhanced by overcoming Henry's Law, e.g. by the absorption of formaldehyde (HCHO) by raindrops, and the reaction of formaldehyde with HCN in the droplet (Granville-Willett et al., 2018). 4. Results We present our results for the HCN photochemistry as a function of the C/O ratio and UV flux (Section 4.1), and as a function of the barrier to destruction by atomic oxygen (Section 4.2). 4.1. The C/O Ratio, UV Flux and HCN Photochemistry Regarding the C/O ratio, the photochemistry follows thermochemistry (Fig. 6). Given an atmosphere dominated by N2: • C/O ≤ 0.5: The production of HCN is low, with mixing ratios of < 1 ppm, regardless of the particular atmospheric composition. • 0.5 < C/O ≤ 1.5: The production of HCN varies widely, with mixing ratios ranging from 1 ppm to 1000 ppm. In this regime, the particular atmospheric composition is important. If the atmosphere is 90% N2 and 10% CO, HCN mix- ing ratios only come to 10 ppm, but if the atmosphere is 90% N2, 5% CO2 and 5% CH4, the HCN mixing ratio reaches 500 ppm. • C/O > 1.5: The production of HCN is high, with mixing ratios of > 0.1 % of the atmosphere, regardless of the particular atmospheric composition. Photochemical production of HCN depends on the UV stellar spectrum. The EUV photons (which here we restrict to photons of wavelength 500 Å -- 1000 Å) dissociate the N2 bond, and artificially enhancing the flux in this wavelength range (see Methods) leads to more HCN. The FUV (here between 1000 Å -- 2000 Å) flux tends to dissociate H2O, producing free OH and O that readily destroy HCN (see Fig. 5). Artificially enhancing the flux in this wavelength range leads to less HCN. Fig. 7 shows the effect of artificially enhancing or reducing the spectrum in these wavelength ranges. This highlights the importance of the UV environment for the HCN abundance. Stars tend to have roughly 1:1 ratios of EUV and FUV light, and so deviations that would enhance or reduce HCN will mostly be due to atmospheric attenuation at specific wavelengths of light, or spectral variability due to stellar activity (e.g. flaring). 10 Figure 6: HCN mixing ratio as a function of the C/O ratio, applied to a 0.1 mbar gas composed of various ratios of CH4, CO and CO2, mixed with a 0.9 mbar gas composed of 99% N2, 1% H2O vapor and 0.1% H2, exposed to the ultraviolet flux of the young Sun. The region spans several orders of magnitude, and the range can be explained by the variety of ways one can combine carbonaceous gases to achieve the same C/O ratio. For example, C/O=1 can be achieved by having a pure CO gas, or by having a gas of 50% CO2 and 50% CH4. The composition will depend on outgassing pressure and redox state (Gaillard and Scaillet, 2014). For a given C/O ratio, the presence of methane leads to considerably more HCN. 4.2. Atomic Oxygen as an HCN Sink and Atmospheric Methylamine As with any molecule, the mixing ratio of HCN is determined by a balance of formation and loss. Hydroxyl radicals and atomic oxygen are the prime sinks of HCN. The rate constants for HCN destruction by O and OH are not all well-constrained. In particular, for the reaction: HCN + O → NCO + H. (7) There are several choices for the rate constant from the literature. The energy barriers, including error bars, range from ∼ 20 kJ/mol (Louge and Hanson, 1985) to over 100 kJ/mol (Tsang and Herron, 1991), and values in between (Perry and Melius, 1985). We vary the rate constant of Reaction 7 for a model atmosphere of bulk composition 89% N2, 10% CO2 and 1% CH4, one of the suite of atmospheres from Rugheimer & Rimmer (Rugheimer and Rimmer, 2018), following Rugheimer et al (Rugheimer and Kaltenegger, 2018). The temperature profile is determined using a climate model, and the photochemistry is calculated with this temperature profile and a young solar evolution model (see Methods and Fig. 2). The results for HCN and methylamine (CH5N) are shown in Fig. 8. 11 108107106105104103102HCN mixing ratio0.51.02.05.010.0C/O Figure 7: Ratio of a star's UV spectral irradiance to solar spectral irradiance, Fλ/Fλ,sun at EUV (500-1000 Å) and FUV (1000-2000 Å) wavelengths. 500-1500 Å light dissociates N2, the first step toward forming HCN photochemically. FUV photons dissociate H2O, leading to hydroxyl radicals and free oxygen atoms, which rapidly destroy HCN. The dashed red line indicates where all main-sequence stellar quiescent emission lies. Stellar flares or atmospheric absorption may cause the spectrum to deviate from this line. Atmospheric absorption in particular will preferentially block the 500-1000 Å photons, pushing the photoexposed atmo- sphere into the bottom right region of the plot. The amount of HCN in the atmosphere is clearly sensitive to the energetic barrier for R7. Surprisingly, the amount of methylamine in the upper atmosphere is enhanced as the energetic barrier for HCN destruction by atomic oxygen is lowered. This is because after HCN is destroyed, its products react as follows: NCO + H → CO + NH, NH + CH4 + M → CH5N + M. (8) (9) Near the surface (p > 0.5 bar), the amount of methylamine is only weakly dependent on the energy barrier for R7, because there is less atomic oxygen available to produce NCO. Methylamine is instead produced near the surface by: HCHO + NO → HNCO + OH, HNCO + hν → NH + CO, NH + CH4 + M → CH5N + M. (10) (11) (12) 12 102101100101102FFUV/FFUV,sun102101100101102FEUV/FEUV,sunEUV-enhanced(flares)EUV-shielded(atmospheric absorption)1010109108107106105104HCN mixing ratio To better assess the production of both HCN and methylamine on rocky exoplanets, the rate constants for HCN + O in particular needs to be better constrained. Figure 8: Mixing ratios of HCN (thin black) and methylamine (CH5N, thick gray) as a function of pressure [bar], for a variety of energetic barriers for R7, ranging from 5.2 kJ/mol to 52.4 kJ/mol, with the most probable barrier lying between 16.6 and 52.4 kJ/mol, according to theory and experiment. 5. Discussion and Conclusions We have shown that HCN abundances in the atmospheres of Earth-like rocky plan- ets (N2 dominated) depend critically on the atmospheric C/O ratio, with significantly greater amounts of HCN generated photochemically when C/O (cid:38) 1, regardless of the carbon-bearing species, whether it be CH4, CO, or C2H2. In addition, the tropospheric HCN concentration is determined by the rate of HCN destruction by atomic oxygen. We explore the effect of adjusting the barrier for the reaction HCN + O → NCO + H. The more efficient this reaction, the lower the tropospheric HCN concentrations, but the higher the concentrations of methylamine (CH5N), a species that helps facilitate the prebiotic activation of nucleotides (Mariani et al., 2018). When comparing our photochemistry results to alternative scenarios for forming HCN, such as impact delivery and production and production via lightning, these sce- narios only require that the local C/O ratio be (cid:38) 1, and tend to produce HCN in greater quantities, 0.1-1% concentrations in the case of impacts and lightning shocks, versus 1-100 ppm tropospheric concentrations for the photochemical production of HCN. The 13 10131011109107105103mixing ratio106105104103102101100p [bar]52.4 kJ/mol16.6 kJ/mol5.2 kJ/molCH5NHCN amount of HCN generated in impacts and lightning shocks is therefore more likely to achieve concentrations of HCN used in the lab. If the global C/O ratio is (cid:38) 1 within the atmosphere of a rocky exoplanet, this would suggest that prebiotic chemistry starting with HCN could be initiated in any surface environment for that planet, rather than in certain more exotic scenarios. Such environmental universality may be advantageous for originating life on the surfaces of planets with long-lived reducing atmospheres. As shown above, the formation of HCN depends on the availability of nitrogen. All of our models have started with ∼ 1 bar of nitrogen in the form of N2. The con- centration of N2 will be very difficult to detect remotely on an exoplanet, owing to the molecule's lack of a dipole moment. Airapetian et al. (2017) propose a novel observa- tional strategy to constrain the N2 content of exoplanet atmospheres by searching for photochemical products, especially NO. In more reduced atmospheres, NH3 and HCN itself could be the major photochemical products of N2 destruction. Either of these species would absorb at 1.5 µm, appearing as a wing on the 1.4 µm water feature, as has observed in the atmosphere of HD209458b (MacDonald and Madhusudhan, 2017). The shape and strength of these features will not be so strong in a temporate atmosphere with a significantly greater molecular weight. HCN will also have promi- nent broad features in the 7 and 14 µm regions, although at low resolution these will be difficult to disentangle from other species with similar spectra, such as C2H2 (Tessenyi et al., 2013). HCN could be distinguished from these other species at high resolution. Indeed, HCN has been detected in HD209458b in emission at high resolution (Hawker et al., 2018). The search for tracers of atmospheric conditions favorable to prebiotic chemistry is beyond the scope of this paper. The C/O ratio has been seen as an important dimension for characterizing exoplanet atmospheres(Madhusudhan, 2012), and is clearly an important dimension when con- sidering the production of hydrogen cyanide. It will likely be easier to constrain the C/O ratio for exoplanet systems in the near future than the nitrogen content, although both are essential for the photochemical production of HCN. The C/O ratio can be con- sidered at a number of different scales, in order of decreasing size: cosmic, galactic, cluster, stellar, disk, sub-disk, planetary, surface and atmospheric C/O, and the C/O ratio within particular local environments. We will discuss each of these briefly in this order. The cosmic C/O ratio increases over time, and within our galaxy, is expected to increase towards the galactic center, and to be enhanced in clusters with higher metal- licity (Esteban et al., 2005). As low mass stars evolve and reach the asymptotic giant branch (AGB), carbon and other s-process elements are preferentially dredged up to the surface where they are released to the interstellar medium by the AGB winds (Wi- escher et al., 2010). Therefore as the Universe ages, and increasingly more low mass stars evolve and die, the ratio of C/O will increase. Eventually, perhaps 30-50 Gyr from now, carbon will be more abundant than oxygen on average. This will drastically effect the geochemistry on planets produced from such an environment. When the C/O ratio of a star exceeds 0.8, formation models suggest that planets can be produced in the disk with overall C/O > 1 (Bond et al., 2010). A small frac- tion of stars meet this criterion (Fortney, 2012), and so, even today, more carbon rich areas of the disk could form these so-called "Carbon Planets" (Kuchner and Seager, 2005). When the C/O ratio is < 0.8, carbon-enhanced planets can still be formed near 14 the midplane of the disk, sufficiently far from the host star (Öberg et al., 2011), the distance being determined by ice lines and also by the formation mechanism (Booth et al., 2017). Once a planet's carbon is greater in abundance than oxygen by number, a few inter- esting chemical consequences will take place. First any oxygen will go into producing CO (Lodders and Fegley Jr, 1997). There will still be carbon left after the oxygen is depleted, and this will then bond with silicon, forming silicon carbide, SiC, titanium carbide, TiC, or will bind with itself to form graphite (Lodders and Fegley Jr, 1997). On Earth, the extra oxygen bonds with silicon forming silicates, the most abundant chemical species in Earth's crust. SiC, however, is known to be an extremely durable mineral. Since carbon bonds are stronger than oxygen, it is more difficult to melt and break these bonds. Thus plate tectonics and weathering would be more difficult with silicon carbide rocks rather than silicates. Both of these effects may cause problems in maintaining a carbon cycle on such a planet and thus a long-term habitable climate. For stars with a lower C/O ratio, like our Sun (C/O ≈ 0.55 (Asplund et al., 2009)), formation in different parts of the disk will still affect the C/O ratio, but will probably not form "carbon planets". But the story doesn't end there. During and after planet formation, carbon is differentiated either into the core (carbon at high pressure becomes a siderophile (Dasgupta, 2013)), or into the crust and atmosphere, in the form of SiC (Dasgupta, 2013; Bond et al., 2010), graphite (Dasgupta, 2013; Frost and McCammon, 2008), or outgassed as CO2, CO, CH4 (Gaillard and Scaillet, 2014), or C2H2 (Hu and Seager, 2014). At least one object in our Solar System, Titan, has an atmosphere and crust with a C/O ratio (cid:29) 1 (Lunine and Atreya, 2008). There is evidence that more than 3.8 Ga, the Earth's surface may also have had a C/O ratio (cid:38) 1 (Yang et al., 2014). Finally, it is worth noting that local C/O ratios vary by orders of magnitude even on the present-day Earth with its oxidized crust and atmosphere (Fischer, 2008), with mud volcanos releasing gas with ratios of C/O ≈ 100 (Hedberg, 1974; Etiope and Milkov, 2004). The debate about the C/O ratio continues (Sing et al., 2016; MacDonald and Mad- husudhan, 2017), and hopefully the advent of JWST will be able to provide some clar- ity on this question. The persistent connetion between the production of HCN and the C/O ratio implies that there is a connection between a rocky planet's C/O ratio and the chances for life to originate on its surface via any chemical pathway involving HCN. This C/O ratio, along with a better understanding of stellar UV spectra (Ran- jan and Sasselov, 2016, 2017; Rimmer et al., 2018) and nitrogen content of expolanets (Airapetian et al., 2016; Airapetian et al., 2017) may be key for determining whether a given rocky planet lies within the abiogenesis zone. Acknowledgements P. B. R. acknowledges support by the Simons and Kavli Foundations for this work, including Simons Foundation SCOL awards 599634. S. R. acknowledges support by a grant from the Simons Foundation (SCOL awards 339489). This work has made use of the MUSCLES M dwarf UV radiation field database. P. B. R. thanks Sukrit Ranjan and Dimitar Sasselov for helpful comments. 15 References Airapetian, V. S., Glocer, A., Gronoff, G., Hebrard, E., Danchi, W., 06 2016. Prebiotic chemistry and atmospheric warming of early earth by an active young sun. Nature Geosci 9 (6), 452 -- 455. Airapetian, V. S., Jackman, C. H., Mlynczak, M., Danchi, W., Hunt, L., Nov. 2017. Atmospheric Beacons of Life from Exoplanets Around G and K Stars. Scientific Reports 7, 14141. Ardaseva, A., Rimmer, P. B., Waldmann, I., Rocchetto, M., Yurchenko, S. N., Helling, C., Tennyson, J., Apr. 2017. Lightning Chemistry on Earth-like Exoplanets. ArXiv e-prints. Asplund, M., Grevesse, N., Sauval, A. J., Scott, P., Sep. 2009. The Chemical Compo- sition of the Sun. ARAA47, 481 -- 522. Becker, R. S., Hong, J., 1983. Photochemistry of acetylene, hydrogen cyanide, and mixtures. The Journal of Physical Chemistry 87 (1), 163 -- 166. Bond, J. C., O'Brien, D. P., Lauretta, D. S., Jun. 2010. The Compositional Diversity of Extrasolar Terrestrial Planets. I. In Situ Simulations. ApJ715, 1050 -- 1070. Booth, R. A., Clarke, C. J., Madhusudhan, N., Ilee, J. D., Aug. 2017. Chemical enrich- ment of giant planets and discs due to pebble drift. MNRAS469, 3994 -- 4011. Chameides, W. L., Walker, J. C. G., Dec. 1981. Rates of Fixation by Lightning of Carbon and Nitrogen in Possible Primitive Atmospheres. Origins of Life 11, 291 -- 302. Claire, M. W., Sheets, J., Cohen, M., Ribas, I., Meadows, V. S., Catling, D. C., Sep. 2012. The Evolution of Solar Flux from 0.1 nm to 160 µm: Quantitative Estimates for Planetary Studies. ApJ757, 95. Dasgupta, R., Jan. 2013. Ingassing, Storage, and Outgassing of Terrestrial Carbon through Geologic Time. Reviews in Mineralogy and Geochemistry 75, 183 -- 229. Esteban, C., García-Rojas, J., Peimbert, M., Peimbert, A., Ruiz, M. T., Rodríguez, M., Carigi, L., Jan. 2005. Carbon and Oxygen Galactic Gradients: Observational Values from H II Region Recombination Lines. ApJL618, L95 -- L98. Etiope, G., Milkov, A. V., 2004. A new estimate of global methane flux from onshore and shallow submarine mud volcanoes to the atmosphere. Environmental Geology 46 (8), 997 -- 1002. Ferris, J. P., Orgel, L. E., 1966. An unusual photochemical rearrangement in the syn- thesis of adenine from hydrogen cyanide. Journal of the American Chemical Society 88 (5), 1074 -- 1074. 16 Ferus, M., Kubelík, P., Knížek, A., Pastorek, A., Sutherland, J., Civiš, S., 2017. High energy radical chemistry formation of hcn-rich atmospheres on early earth. Scientific reports 7 (1), 6275. Ferus, M., Nesvorný, D., Šponer, J., Kubelík, P., Michalcíková, R., Shestivská, V., Šponer, J. E., Civiš, S., Jan. 2015. High-energy chemistry of formamide: A uni- fied mechanism of nucleobase formation. Proceedings of the National Academy of Science 112, 657 -- 662. Fischer, T. P., 2008. Fluxes of volatiles (h2o, co2, n2, cl, f) from arc volcanoes. Geo- chemical Journal 42 (1), 21 -- 38. Fortney, J. J., Mar. 2012. On the Carbon-to-oxygen Ratio Measurement in nearby Sun- like Stars: Implications for Planet Formation and the Determination of Stellar Abun- dances. ApJL747, L27. Frost, D. J., McCammon, C. A., May 2008. The Redox State of Earth's Mantle. Annual Review of Earth and Planetary Sciences 36, 389 -- 420. Gaillard, F., Scaillet, B., Oct. 2014. A theoretical framework for volcanic degassing chemistry in a comparative planetology perspective and implications for planetary atmospheres. Earth and Planetary Science Letters 403, 307 -- 316. Gao, P., Hu, R., Robinson, T. D., Li, C., Yung, Y. L., Jun. 2015. Stability of CO2 Atmospheres on Desiccated M Dwarf Exoplanets. ApJ806, 249. Granville-Willett, A., Archibald, A. T., Rimmer, P. B., Griffiths, P., Sutherland, J. D., 2018. Overcoming Henry's Law. in prep. Haqq-Misra, J. D., Domagal-Goldman, S. D., Kasting, P. J., Kasting, J. F., 2008. A revised, hazy methane greenhouse for the archean earth. Astrobiology 8 (6), 1127 -- 1137. Hawker, G. A., Madhusudhan, N., Cabot, S. H. C., Gandhi, S., Aug. 2018. Evidence for Multiple Molecular Species in the Hot Jupiter HD 209458b. ApJL863, L11. Hedberg, H. D., 1974. Relation of methane generation to undercompacted shales, shale diapirs, and mud volcanoes. AAPG Bulletin 58 (4), 661 -- 673. Hu, R., Seager, S., Mar. 2014. Photochemistry in Terrestrial Exoplanet Atmospheres. III. Photochemistry and Thermochemistry in Thick Atmospheres on Super Earths and Mini Neptunes. ApJ784, 63. Kharecha, P., Kasting, J. F., Siefert, J., 2005. A coupled atmosphere -- ecosystem model of the early archean earth. Geobiology 3, 53 -- 73. Kuchner, M. J., Seager, S., Apr. 2005. Extrasolar Carbon Planets. ArXiv Astrophysics e-prints. 17 Lodders, K., Fegley Jr, B., 1997. In astrophysical implications of the laboratory study of presolar materials; bernatowicz, tj; zinner, e., eds. In: AIP Conference Proceed- ings. Vol. 402. p. 391. Louge, M. Y., Hanson, R. K., 1985. Shock tube study of nco kinetics. In: Symposium (International) on Combustion. Vol. 20. Elsevier, pp. 665 -- 672. Lunine, J. I., Atreya, S. K., Mar. 2008. The methane cycle on Titan. Nature Geoscience 1, 159 -- 164. MacDonald, R. J., Madhusudhan, N., Aug. 2017. HD 209458b in new light: evidence of nitrogen chemistry, patchy clouds and sub-solar water. MNRAS469, 1979 -- 1996. Madhusudhan, N., Oct. 2012. C/O Ratio as a Dimension for Characterizing Exoplane- tary Atmospheres. ApJ758, 36. Mariani, A., Russell, D. A., Javelle, T., Sutherland, J. D., 2018. A light-releasable potentially prebiotic nucleotide activating agent. Journal of the American Chemical Society 140 (28), 8657 -- 8661. Miller, S. L., 1957. The mechanism of synthesis of amino acids by electric discharges. Biochimica et Biophysica Acta 23, 480 -- 489. Öberg, K. I., Murray-Clay, R., Bergin, E. A., Dec. 2011. The Effects of Snowlines on C/O in Planetary Atmospheres. ApJL743, L16. Patel, B. H., Percivalle, C., Ritson, D. J., Duffy, C. D., Sutherland, J. D., Apr. 2015. Common origins of RNA, protein and lipid precursors in a cyanosulfidic pro- tometabolism. Nature Chemistry 7, 301 -- 307. Pavlov, A. A., Kasting, J. F., Brown, L. L., Rages, K. A., Freedman, R., 2000. Green- house warming by ch4 in the atmosphere of early earth. Journal of Geophysical Research: Planets 105 (E5), 11981 -- 11990. Perry, R. A., Melius, C. F., 1985. The rate and mechanism of the reaction of hcn with oxygen atoms over the temperature range 540 -- 900 k. In: Symposium (International) on Combustion. Vol. 20. Elsevier, pp. 639 -- 646. Pierrehumbert, R., Gaidos, E., Jun. 2011. Hydrogen Greenhouse Planets Beyond the Habitable Zone. ApJL734, L13. Ranjan, S., Sasselov, D. D., Jan. 2016. Influence of the UV Environment on the Syn- thesis of Prebiotic Molecules. Astrobiology 16, 68 -- 88. Ranjan, S., Sasselov, D. D., Mar. 2017. Constraints on the Early Terrestrial Surface UV Environment Relevant to Prebiotic Chemistry. Astrobiology 17, 169 -- 204. Rimmer, P. B., Helling, C., May 2016. A Chemical Kinetics Network for Lightning and Life in Planetary Atmospheres. The Astrophysical Journal Supplement Series 224, 9. 18 Rimmer, P. B., Xu, J., Thompson, S. J., Gillen, E., Sutherland, J. D., Queloz, D., 2018. The origin of rna precursors on exoplanets. Science Advances 4 (8), eaar3302. Ritson, D., Sutherland, J. D., 2012. Prebiotic synthesis of simple sugars by photoredox systems chemistry. Nature chemistry 4 (11), 895. Rugheimer, S., Kaltenegger, L., Feb. 2018. Spectra of Earth-like Planets through Geo- logical Evolution around FGKM Stars. ApJ854, 19. Rugheimer, S., Rimmer, P. B., 2018. Exoplanet Prebiosignatures. in prep. Saladino, R., Crestini, C., Costanzo, G., Negri, R., Di Mauro, E., 2001. A possible prebiotic synthesis of purine, adenine, cytosine, and 4 (3h)-pyrimidinone from for- mamide: implications for the origin of life. Bioorganic & medicinal chemistry 9 (5), 1249 -- 1253. Sanchez, R. A., Ferbis, J. P., Orgel, L. E., 1967. Studies in prebiodc synthesis: Ii. synthesis of purine precursors and amino acids from aqueous hydrogen cyanide. Journal of molecular biology 30 (2), 223 -- 253. Sanchez, R. A., Ferris, J. P., Orgel, L. E., 1966. Cyanoacetylene in prebiotic synthesis. Science 154 (3750), 784 -- 785. Sanchez, R. A., Orgel, L. E., 1970. Studies in prebiotic synthesis: V. synthesis and photoanomerization of pyrimidine nucleosides. Journal of molecular biology 47 (3), 531 -- 543. Sing, D. K., Fortney, J. J., Nikolov, N., Wakeford, H. R., Kataria, T., Evans, T. M., Aigrain, S., Ballester, G. E., Burrows, A. S., Deming, D., Désert, J.-M., Gibson, N. P., Henry, G. W., Huitson, C. M., Knutson, H. A., Lecavelier Des Etangs, A., Pont, F., Showman, A. P., Vidal-Madjar, A., Williamson, M. H., Wilson, P. A., Jan. 2016. A continuum from clear to cloudy hot-Jupiter exoplanets without primordial water depletion. Nature529, 59 -- 62. Strecker, A., 1854. Ueber einen neuen aus aldehyd-ammoniak und blausäure entste- henden körper. Justus Liebigs Annalen der Chemie 91 (3), 349 -- 351. Tessenyi, M., Tinetti, G., Savini, G., Pascale, E., Sep. 2013. Molecular Detectabil- ity in Exoplanetary Emission Spectra. European Planetary Science Congress 8, EPSC2013 -- 817. Tian, F., Kasting, J. F., Zahnle, K., Aug. 2011. Revisiting HCN formation in Earth's early atmosphere. Earth and Planetary Science Letters 308, 417 -- 423. Toon, O. B., McKay, C. P., Ackerman, T. P., Santhanam, K., Nov. 1989. Rapid calcula- tion of radiative heating rates and photodissociation rates in inhomogeneous multiple scattering atmospheres. Journal of Geophysical Research: Atmospheres 94, 16287 -- 16301. 19 Tsai, S.-M., Lyons, J. R., Grosheintz, L., Rimmer, P. B., Kitzmann, D., Heng, K., Feb. 2017. VULCAN: An Open-source, Validated Chemical Kinetics Python Code for Exoplanetary Atmospheres. ApJS228, 20. Tsang, W., Herron, J. T., 1991. Chemical kinetic data base for propellant combustion i. reactions involving no, no2, hno, hno2, hcn and n2o. Journal of physical and chemi- cal reference data 20 (4), 609 -- 663. Wiescher, M., Görres, J., Uberseder, E., Imbriani, G., Pignatari, M., Nov. 2010. The Cold and Hot CNO Cycles. Annual Review of Nuclear and Particle Science 60, 381 -- 404. Wordsworth, R., Pierrehumbert, R., Apr. 2014. Abiotic Oxygen-dominated Atmo- spheres on Terrestrial Habitable Zone Planets. ApJL785, L20. Xu, J., Ritson, D. J., Ranjan, S., Todd, Z. R., Sasselov, D. D., Sutherland, J. D., 2018. Photochemical reductive homologation of hydrogen cyanide using sulfite and ferro- cyanide. Chemical Communications 54 (44), 5566 -- 5569. Yang, X., Gaillard, F., Scaillet, B., May 2014. A relatively reduced Hadean continen- tal crust and implications for the early atmosphere and crustal rheology. Earth and Planetary Science Letters 393, 210 -- 219. Zahnle, K. J., Feb. 1986. Photochemistry of methane and the formation of hydrocyanic acid (HCN) in the Earth's early atmosphere. JGR91, 2819 -- 2834. 20
1312.5146
1
1312
2013-12-18T14:10:41
Can eccentric debris disks be long-lived? A first numerical investigation and application to $\zeta^2$ Reticuli
[ "astro-ph.EP" ]
Imaging of debris disks has found evidence for both eccentric and offset disks. One hypothesis is that these provide evidence for massive perturbers that sculpt the observed structures. One such disk was recently observed in the far-IR by the Herschel Space Observatory around $\zeta^2$ Ret. In contrast with previously reported systems, the disk is significantly eccentric, and the system is Gyr-old. We aim to investigate the long-term evolution of eccentric structures in debris disks caused by a perturber on an eccentric orbit. Both analytical predictions and numerical N-body simulations are used to investigate the observable structures that could be produced by eccentric perturbers. The long-term evolution of the disk geometry is examined, with particular application to the $\zeta^2$ Ret system. In addition, synthetic images of the disk are produced for comparison with Herschel observations. We show that an eccentric companion can produce both the observed offsets and eccentric disks. Such effects are not immediate and we characterise the timescale required for the disk to develop to an eccentric state. For the case of $\zeta^2$ Ret, we place limits on the mass and orbit of the companion required to produce the observations. Synthetic images show that the pattern observed around $\zeta^2$ Ret can be produced by an eccentric disk seen close to edge-on, and allow us to bring additional constraints on the disk parameters of our model (disk flux, extent). We determine that eccentric planets or stellar companions can induce long-lived eccentric structures in debris disks. Observations of such eccentric structures provide potential evidence of the presence of such a companion in a planetary system. We consider the example of $\zeta^2$ Ret, whose observed eccentric disk can be explained by a distant companion at tens of AU, on an eccentric orbit ($e_p\gtrsim 0.3$).
astro-ph.EP
astro-ph
Astronomy & Astrophysics manuscript no. zetaRet September 18, 2018 c(cid:13) ESO 2018 Can eccentric debris disks be long-lived? A first numerical investigation & application to ζ2 Reticuli V. Faramaz1, H. Beust1, P. Th´ebault2, J.-C. Augereau1, A. Bonsor1, C. del Burgo3, S. Ertel1, J.P. Marshall4, J. Milli 1,5, B. Montesinos6, A. Mora7, G. Bryden8, W. Danchi9, C. Eiroa4, G.J. White10,11, and S. Wolf12 3 1 0 2 c e D 8 1 . ] P E h p - o r t s a [ 1 v 6 4 1 5 . 2 1 3 1 : v i X r a (Affiliations can be found after the references) Received 12 August 2013; Accepted 13 December 2013 ABSTRACT Context. Imaging of debris disks has found evidence for both eccentric and offset disks. One hypothesis is that these provide evidence for massive perturbers, for example planets or binary companions, that sculpt the observed structures. One such disk was recently observed in the far-IR by the Herschel(cid:63) Space Observatory around ζ2 Reticuli. In contrast with previously reported systems, the disk is significantly eccentric, and the system is Gyr-old. Aims. We aim to investigate the long-term evolution of eccentric structures in debris disks caused by a perturber on an eccentric orbit around the star. We hypothesise that the observed eccentric disk around ζ2 Reticuli might be evidence of such a scenario. If so we are able to constrain the mass and orbit of a potential perturber, either a giant planet or binary companion. Methods. Analytical techniques are used to predict the effects of a perturber on a debris disk. Numerical N-body simulations are used to verify these results and further investigate the observable structures that could be produced by eccentric perturbers. The long-term evolution of the disk geometry is examined, with particular application to the ζ2 Reticuli system. In addition, synthetic images of the disk are produced for direct comparison with Herschel observations. Results. We show that an eccentric companion can produce both the observed offsets and eccentric disks. Such effects are not immediate and we characterise the timescale required for the disk to develop to an eccentric state (and any spirals to vanish). For the case of ζ2 Reticuli, we place limits on the mass and orbit of the companion required to produce the observations. Synthetic images show that the pattern observed around ζ2 Reticuli can be produced by an eccentric disk seen close to edge-on, and allow us to bring additional constraints on the disk parameters of our model (disk flux, extent). Conclusions. We determine that eccentric planets or stellar companions can induce long-lived eccentric structures in debris disks. Observations of such eccentric structures, thus, provide potential evidence of the presence of such a companion in a planetary system. We consider the specific example of ζ2 Reticuli, whose observed eccentric disk can be explained by a distant companion (at tens of AU), on an eccentric orbit (ep (cid:38) 0.3). Key words. Circumstellar matter -- Methods: N-body Simulations -- Celestial mechanics -- Stars: ζ2 Reticuli, planetary systems 1. Introduction The first debris disk was discovered in 1984, when the Infrared Astronomical Satellite (IRAS) found a strong IR excess around Vega, revealing the presence of micron-sized dust grains (Aumann et al. 1984). For most debris disks, these grains have a limited lifetime, which, due to Poynting Robertson drag and collisions, is shorter than the system's age. Therefore, this dust is assumed to be replenished by collisional grinding of much larger parent bodies, which are at least kilometre-sized in order for this collisional cascade to be sustained over the system's age (Backman & Paresce 1993; Lohne et al. 2008). Consequently, debris disks provide evidence for the existence of solid bodies having reached km-size, and potentially the planetary-size level. Spatially resolved structures in debris disks can provide clues to the invisible planetary component of those systems. Such planets may be responsible for sculpting these disks, and may leave their signature through various asymmetries such as Send offprint requests to: V. Faramaz (cid:63) Herschel Space Observatory is an ESA space observatory with sci- ence instruments provided by European-led Principal Investigator con- sortia and with important participation from NASA. Correspondence to: [email protected] wing asymmetries, resonant clumpy structures, warps, spirals, gaps or eccentric ring structures (see, e.g., Wyatt 1999). This diversity is to be compared with the variety of exoplanetary sys- tems1 discovered around main sequence stars since 1995 (51 Peg b, Mayor & Queloz 1995). Dynamical modelling of such asym- metries is the only method to place constraints on the masses and orbital parameters of planets in systems where direct observa- tions are not possible (see, e.g., Mouillet et al. 1997; Wyatt et al. 1999; Augereau et al. 2001; Moro-Mart´ın & Malhotra 2002; Wyatt 2004; Kalas et al. 2005; Quillen 2006; Stark & Kuchner 2008; Chiang et al. 2009; Ertel et al. 2011; Boley et al. 2012; Ertel et al. 2012; Thebault et al. 2012). We focus here on cases of eccentric patterns in debris disks. The modelling of this type of asymmetry and its possible link with the dynamical influence of eccentric companions has been investigated in several earlier studies: authoritative work was carried out by Wyatt et al. (1999, 2000) for the case of HR 4796. Another case of interest is the debris disk of Fomalhaut (Stapelfeldt et al. 2004; Kalas et al. 2005; Quillen 2006; Chiang et al. 2009; Boley et al. 2012; Kalas et al. 2013, Beust et al, in revision). 1 see www.exoplanets.org 1 Faramaz et al.: Can eccentric debris disks be long-lived? This pioneering work showed that these large-scale struc- tures arise in systems where debris disks are perturbed by an eccentric companion, on a low inclination orbit relative to the disk (Wyatt et al. 1999). The disk centre of symmetry is offset from the star which may be measured explicitly in high reso- lution images (e.g., HST scattered light). Furthermore, its peri- astron is closer to the star and thus hotter and brighter, which results in a two-sided brightness asymmetry. However, it is important to emphasize that previous studies considered low eccentricity rings (e (cid:38) 0.02 for HR4796 and e = 0.11 ± 0.01 for Fomalhaut), and were limited to timescales smaller than the typical ages of mature disks (≤ 10 Myr for HR4796 simulations and ≤ 100 Myr for Fomalhaut). The issue of whether highly eccentric ring structures could be sustained over very long timescales has not been addressed thus far in the literature. This issue has become very topical because of the dis- covery of at least two Gyr-old and significantly eccentric debris disks: one around ζ2 Reticuli (e (cid:38) 0.3 Eiroa et al. 2010), which is used as a reference case in this paper, and another one around HD 202628 (e ∼ 0.18 Stapelfeldt et al. 2012; Krist et al. 2012). These systems are both older than Fomalhaut or HR 4796, with disks that are also much more eccentric. In the present work, we investigate the long-term evolution of highly eccentric structures in debris disks, and their relation to planetary or stellar perturbers, by investigating their evolu- tion over Gyr timescales. One of the questions we address is whether these structures are really Gyr-old, or might have orig- inated from a more recent event, be it a flyby or the late excite- ment of a sherpherding planet's eccentricity. We also summarize a general modelling method, based on complementary analyti- cal and numerical tools, which we apply to the specific case of ζ2 Reticuli. This paper is structured as follows: Sect. 2 presents how a perturber can generate an eccentric ring structure. Useful ana- lytical expressions are derived, to study under which conditions such a pattern can be created. We also show that these predic- tions can be complemented by numerical studies. Sect. 3 de- scribes the debris disk of ζ2 Reticuli, along with newly reduced Herschel/PACS images. This debris disk is used as a proxy to determine a numerical set-up. Then, in Sect. 4, the numerical investigation is carried out. From N-body simulations, we exam- ine both the onset and suvival of an eccentric pattern and explore their dependencies to the perturber's characteristics. This mod- elling approach allows one to put constraints on a perturber at work in shaping a debris disk into an eccentric ring over Gyr timescales, and it is applied to the case of the debris disk of ζ2 Reticuli. Sect. 5 shows synthetic images to perform a full comparison with observations of ζ2 Reticuli, and retrieve ad- ditional constraints on this disk. Finally, Sect. 6 is devoted to conclusions, discussions and proposing future work. 2. Footprints of eccentric companions on debris disks We have developed a dynamical model to investigate the shaping of a debris disk into an eccentric ring, and the timescales asso- ciated with its onset and survival. More specifically, we seek to determine if any perturbers are able to shape and maintain a disk into a significantly eccentric ring structure on Gyr timescales, and whether the asymmetry relaxes or not. Before presenting our model and our results in detail for this as yet unexplored case, we present the current understanding on how eccentric ring structures arise as a result of the dynamical effect of an eccentric perturber. 2 2.1. Basic principle For a disk to be shaped into an eccentric ring, it must be per- turbed in such a way that its components have both their eccen- tricities forced to higher values, and their orbits more or less ori- ented in a common direction. These conditions are both fulfilled if the disk is under the gravitational influence of a perturber, namely a planetary or a stellar companion, (nearly) coplanar to the disk, and on an eccentric orbit. The eccentricity of the ring causes the disk center of symmetry to be offset from the star, and the disk pericenter to be brighter than the apocenter, since it is closer to the star and thus hotter. This feature was studied by Wyatt et al. (1999) in the case of a planetary companion, and called the pericenter glow phenomenon. Spatially resolved imaging is required to determine the struc- ture of debris disks and therefore renewed efforts have been made to image as many debris disks as possible2. Most images of resolved debris disks have been obtained so far in the visi- ble or near-IR. At these wavelengths, the emission is dominated by small grains close to the blow-out limit imposed by stellar radiation (sub-micron to micron depending on stellar luminos- ity). Such grains exhibit complex evolution due to the coupled effects of collisions and radiation pressure (see, e.g., Th´ebault & Augereau 2007). This may strongly alter, or even mask the dynamical structures imparted by a massive perturber (Th´ebault et al. 2012). Observations at longer wavelengths detect bigger grains less affected by stellar radiation (few tens to few hun- dreds of micron in size, depending on observing wavelength). These should directly trace the distribution of larger parent bod- ies and thus more directly reflect the dynamical effect of a per- turber (see, e.g., Krivov 2010; Moro-Mart´ın 2012, for exhaustive reviews). Since the origin of an eccentric pattern is gravitational, we can reasonably assume that large scale asymmetries among an observed dust population already exist amongst the parent plan- etesimal population that produces it and result from pure grav- itational perturbations. This assumption allows us to study the influence of different eccentric perturbers in a simplified way, neglecting the effect of radiation pressure and considering par- ent planetesimals as massless and collisionless particles in orbit around their host star and perturbed by a companion, be it stellar or planetary. We assume that at the end of the protoplanetary phase, the planetesimals start from almost circular orbits because of orbital eccentricity damping by primordial gas. We further assume that any perturbing planet among the system is fully formed by the time the gas disappears, and evolves on a significantly eccentric orbit, due to a major perturbing event such as planet-planet scat- tering. Thus, we can consider the disappearance of the gas as time zero for the onset of planetesimal perturbations by an ec- centric companion. From this moment the planetesimal eccen- tricities start to increase and their lines of apsides tend to align with the planet's. Under these assumptions, the forced elliptic ring structure takes some time to settle in, and it is preceded by the appearance and disappearance of transient spiral features. These are due to differential precession within the disk: all the planetesimals in the disk have different precession rates (due to their different orbital distances), such that these spiral structures are expected to wind up and finally generate an eccentric ring, as shown by Augereau & Papaloizou (2004) and Wyatt (2005). The characteristic time for reaching this state is of the order of a few precession timescales at the considered distance (Wyatt 2005). 2 see, e.g., www.circumstellardisks.org Faramaz et al.: Can eccentric debris disks be long-lived? Consequently, the onset of an eccentric ring structure is a matter of timescales, while the value of the disk global eccentricity is to be linked with the planetesimals' forced eccentricity, and thus to the companion's eccentricity. 2.2. Analytical approach We show here how the onset of these structures can be under- stood from analytical considerations. Planetesimals are consid- ered to be massless particles. We focus on the secular response of a debris disk to a coplanar perturbing body, either a planet or star. More specifically, both the forced secular eccentricity ef and apsidal precession rate d/dt of test planetesimals are ex- amined, where  is the longitude of periastron with respect to the direction of the perturber's periastron, i.e., the planet and planetesimal have their periastra aligned when  = 0 and anti- aligned when  = π. When secularly perturbed, the eccentricity of a planetesimal evolves cyclically, the period of which is related to the rate of orbital precession. In particular, if we consider a dynamically cold disk of planetesimals as an initial condition, which, con- sidering the damping effect of the gas during the protoplanetary phase, is a reasonable and classical assumption. In that case, the secular behaviour of a planetesimal can be understood consid- ering the analytical solution for its eccentricity. In this case, in the Laplace-Lagrange theory, the complex eccentricity of a plan- etesimal, z(t), can be written: (cid:8)1 − exp(IAt)(cid:9) z(t) = ef where I2 = −1, and A = d dt , (1) is the secular precession rate. One can see from this expression that the maximum induced eccentricity for a planetesimal is twice the forced eccentricity, i.e., e f,max = 2ef. This occurs when At = π[2π], i.e., when the longitudes of periastra of the planetesimal and the perturber are equal ( = 0, see Fig. 1 and, for further details, see e.g., Wyatt 2005; Beust et al. 2013). There are several ways to analytically derive ef and d/dt. The most classical one is to apply linear Laplace-Lagrange the- ory, i.e. an expansion of the interaction Hamiltonian to second order in ascending powers of the eccentricities of both bodies and an averaging over both orbits (see, e.g., Eq. 6 of Mustill & Wyatt 2009). However, this approach is valid only for small ec- centricities, whereas the perturber's orbital eccentricity ep is not necessarily small. Therefore, restricting our analytical study to small ep may not be appropriate. Another way to proceed is to expand the interaction Hamiltonian in spherical harmonics, to truncate at some order in α, where α is the ratio3 between a and ap, the planetesimal and the perturber's semi-major axis respectively, and to average after over both orbits. This permits us to perform an analysis without any restriction on the eccentricities. The resulting Hamiltonian is given by Krymolowski & Mazeh (1999); Ford et al. (2000) or Beust & Dutrey (2006). To the lowest order in α (2nd order, quadrupolar), it yields a forced eccentricity ef: ef (cid:39) 5 4 αep 1 − e2 . p (2) This expression is given by Augereau & Papaloizou (2004) and Mustill & Wyatt (Eq. 8 of 2009). Note that this approach is only valid for small enough values of α to ensure a fast conver- gence of the expansion, i.e. for orbits with significantly differ- ent sizes. It is also valid only far from mean-motion resonances. However, these resonances' spatial extension in semi-major axis (of the order of ∼ 0.1 AU) is typically two orders of magnitude smaller than the extent of the observed structures (of the order of ∼ 10 AU), although in the case where particles are on ec- centric orbits, these resonances may span much larger ranges in terms of radial distance to the star than their span in semi-major axis would have let suppose. But in any case, the amount of material trapped in resonance can reasonably be assumed to be much smaller than the amount of non-resonant material, all the more since we do not suppose here that the planet has migrated, and thus exclude resonant capture during migration. Therefore, these structures are assumed to result from non-resonant mate- rial, and our approach is appropriate. Moreover, resonances will be treated in our N-body integrations, and will be confirmed im- portant only for limited parameter combinations (see Sect. 4). To derive the precession rate in the spherical harmonic ex- pansion case, we follow the method given by Mardling & Lin (2002). The variation rate for the Runge-Lenz vector of the orbit is computed, expanded to any given order, and integrated over one orbital period. Since after numerical tests, one notices that there is less than two orders of magnitude between terms of the 2nd and the 4th order (the third order terms cancel out), we retain terms up to 4th order in the spherical harmonic expansion, and average the resulting precession rate over the longitude of periastron. The precession rate is: Fig. 1. Co-evolution of a planetesimal eccentricity and orbital precession, when acted upon by an eccentric perturber. As the planetesimal orbit precesses, when the longitudes of periastra of the planetesimal and the perturber are equal ( = 0), i.e., when the planetesimal and the perturber have their pericentres aligned on the same side of the massive central body, the planetesimal eccentricity is maximum. d dt = 3n 4 mp M(cid:63) α3 + 45n 256 mp M(cid:63) 1 − e2 p)3/2 √ (1 − e2 α5 (4 + 3e2)(2 + 3e2 p) (1 − e2 p)7/2 √ 1 − e2 . (3) 3 α is such that α < 1, always, and thus α = ap/a if ap < a, and inversely, α = a/ap if ap > a. 3 Faramaz et al.: Can eccentric debris disks be long-lived? minimum and maximum eccentricities induced across the disk.4 The minimum and maximum precession timescales, tprec,min and tprec,max, are defined in the same manner. Then, using Eq. 5 and Eq. 6, one obtains: ef,max ef,min aout ain (7) = , and tprec,max tprec,min = (cid:32) aout (cid:33)3 ain . (8) It is easy to see from these equations that the secular pre- cession timescale spans a large range of values across the disk. This means that making analytical predictions by setting the wanted values for the forced eccentricity and the secular preces- sion timescale for a particle with semi-major axis at the center of the distribution suffers limitations when applied to an extended disk, especially concerning the timescale. Eq. 7 and Eq. 8 can be rewritten using ∆a, the half width of the disk extent, along with ef,c and tprec,c, the forced eccentricity and secular precession timescale at ac, respectively: (cid:32) ac ± ∆a (cid:33) (cid:32) ac ± ∆a ac ac ef,c , (cid:33)3 tprec,c . ef,max/min = and tprec,max/min = (9) (10) As an example, we set ac = 100 AU, ∆a = 25 AU, 2ec = 0.3 and tprec,c = 1 Gyr. These values are close to those derived for the disk of ζ2 Reticuli (e (cid:38) 0.3 and extent 70-120 AU: Eiroa et al. 2010, and Sect. 3 of the present work). One obtains: 2ef,min/max = 0.225 − 0.375 tprec,min/max = 0.42 − 1.95 Gyr (cid:40) (11) . In these conditions, the extent of the disk is not expected to affect too much the global eccentricity of the disk, i.e., we should recover in average a global eccentricity corresponding to the forced eccentricity at ac, once the steady state is reached. But the problem is that the extent of the disk affects a lot the timescale to reach this steady state. This is a limitation of the analytic approach that can be overcome by the use of numerical simulations. 3. Numerical Investigation: A typical set-up, the highly eccentric, old disk of ζ2 Reticuli To go beyond the simplified analytical approach and explore the high eccentricity case on Gyr timescales, we resort to numerical tools. We place ourselves in the frame of the restricted 3-body problem, i.e. one central star, a planet and a massless planetes- imal. The symplectic N-body code SWIFT-RMVS of Levison & Duncan (1994) is used to integrate the evolution of a ring of planetesimals around a solar mass star, over 1 Gyr. We use a typical timestep of ∼ 1/20 of the smallest orbital period, and en- sure a conservation of energy with a typical error of ∼ 10−6 on 4 ef,min = ef,in and ef,max = ef,out in the case of an inner perturber, and conversely in the case of an outer one. Fig. 2. Example color map of the maximum induced eccentricity 2ef imposed by a planetary perturber on a particle with semi- major axis 100 AU and eccentricity e = 0, as a function of its periastron and eccentricity, as estimated from Eq. (2). The black line corresponds to a 2ef = 0.3 condition, which is set to mimic the condition for the disk of ζ2 Reticuli. Note that it does not depend on the mass of the planet. The white lines show the pa- rameters for which the typical timescale to reach a steady state at 100 AU is tprec = 1 Gyr, using Eq. (4). This timescales depends on the mass: mp = 0.1 MJup (solid line), 1 MJup (dashed line) and 2 MJup (dotted line). For example, a perturber of mass 0.1 MJup, periastron qp = 150 AU and eccentricity ep = 0.4 should produce a significantly eccentric ring in less than 1 Gyr, although spiral patterns may remain since, as was shown by Wyatt (2005), it can take several precession timescales for them to vanish. We now follow Wyatt et al. (1999) and assume that the pre- cession timescale, tprec, corresponds to the lower limit of the typ- ical dynamical timescale for setting a dynamical steady state: tprec = 2π (d/dt)ac , (4) where ac is the typical semi-major axis of a particle in the ring. We can now make analytical predictions of the effect of a perturber on a debris disk, i.e. for any given set of values of ep and of the planet periastron qp, one can derive the preces- sion rate (d/dt)ac corresponding to planetesimals orbiting at this distance. Conversely, one can set this dynamical timescale and the forced eccentricity for a particle with semi-major axis ac to correspond to those derived from observations of an eccentric debris disk, and thus make an initial estimate of the perturber's characteristics (see Fig. 2). However, the problem is more complex for real disks which have a finite spatial extension, since these estimates depend on radial locations. To first order, it can be seen from Eq. 2, Eq. 3, and Eq. 4 that the forced eccentricity and the secular timescale scale as: ef ∝ α and tprec ∝ 1 mpα3 (6) (5) , . We now define ain and aout as the inner and outer limits of the disk in semi-major axis, and define ef,min and ef,max as the 4 Faramaz et al.: Can eccentric debris disks be long-lived? relative energy. This code has a crucial advantage over an analyt- ical approach: it is able to handle close encounters and scattering processes, along with the short-term variations of the planetes- imals orbital elements, whereas these effects are ignored in the analytical approach, for which orbits are averaged, short-term variations are lost and the approach is valid only for α << 1, i.e. far from close encounters. As will be shown in Sect. 4, the scattering events play a crucial role in the system's evolution. For the sake of clarity, the ζ2 Reticuli system is considered as a proxy for a typical mature and significantly eccentric debris ring. We shall explore different planet-disk configurations, and produce synthetic images for comparison with Herschel/PACS observations. As will be shown, the hypothesis of an eccentric debris disk around ζ2 Reticuli is fully consistent with the obser- vations. 3.1. The ζ2 Reticuli system ζ2 Reticuli (HR 1010, HIP 15371) is a G1V solar-type star (Eiroa et al. 2013) located at 12 pc (van Leeuwen 2007), of luminosity L(cid:63) = 0.97L(cid:12), log g = 4.51, and age ∼ 2 − 3 Gyrs (Eiroa et al. 2013). It has a binary companion ζ1Reticuli, a G2-4V (Gray et al. 2006; Torres et al. 2006) star located at a projected dis- tance of 3713 AU from ζ2 Reticuli (Mason et al. 2001). Bayesian analysis by Shaya & Olling (2011) of the proper motions of these stars indicates a very high (near 100%) probability that the pair are physically connected. The presence of dust around ζ2 Reticuli has already been probed with Spitzer (Trilling et al. 2008), which suggests a ∼ 150 K emission at ∼ 4.3 AU. However, the angular resolu- tion of Spitzer is limited, and the dust spatial distribution re- mained unconstrained. New observations with Herschel/PACS completed the SED, providing the suggestion of an optically thin, ∼ 40 K, emission at ∼ 100 AU, with fractional luminosity Ldust/L(cid:63) ≈ 10−5 (Eiroa et al. 2010). Moreover, Herschel/PACS provided spatially resolved images of the dust thermal emission surrounding ζ2 Reticuli at 70 µm and 100 µm (Eiroa et al. 2010). We present here newly reduced Herschel/PACS images (see Fig. 3). The images show a double-lobe feature, asymmetric both in position and brightness. Note that at 70 µm, the probability for alignment with a background source within 10(cid:48)(cid:48) is extremely low, namely 10−3 (Eiroa et al. 2010). The disk is not resolved at Herschel/SPIRE wavelengths: newly reduced images and star- disk fluxes measurements are presented in Appendix A. As suggested by Eiroa et al. (2010), the asymmetry revealed by Herschel/PACS in the disk of ζ2 Reticuli can be interpreted as a ring-like elliptical structure with e (cid:38) 0.3 seen close to edge-on, and extending from ∼ 70 to ∼ 120 AU, which is fully consistent with the information derived from the SED (Eiroa et al. 2010). Alternatively, it could also be interpreted as two clumps from an over-density of dust and planetesimals. In Appendix B, we investigate the system inclination on the line of sight, a crucial parameter required to interpret correctly the observed structures. More precisely, we determine the star inclination and assume that the disk and the star are coplanar. The 50% probability value is i = 65.5◦, i.e. the system is very inclined on the line of sight, which tends to support the eccentric ring scenario. Without a doubt, this asymmetric structure provides evi- dence that "something" is dynamically sculpting the disk. It could be the stellar companion ζ1Reticuli or a (yet undetected) planet. The latter hypothesis is fully compatible with radial ve- locity measurements of ζ2 Reticuli, which suggest there is no Jupiter-mass (or larger) planet interior to ∼ 5 − 10 AU (Mayor et al. 2003) but put no contraints on small planet, or Jupiter-like Fig. 4. Detection limits set by direct imaging on the presence of brown dwarf / close binary between 1(cid:48)(cid:48) and 2.5(cid:48)(cid:48) in projected separation. planet at larger radii. It is also compatible with growing obser- vational evidence for planets at large orbital separation, i.e., a several tens to few hundreds of AU from their host star (see, e.g., Luhman et al. 2007; Kalas et al. 2008; Marois et al. 2008, 2010). Constraints from direct imaging are presented in Fig. 4 using the two evolutionary models COND 2003 (Baraffe et al. 2003) or BT-settl 2011 (Allard et al. 2011). Details of the reduction procedure is presented in Appendix C. These constraints were obtained from VLT/NaCo archival data taken in November 2010 in the Ks band. These data do not provide constraints on com- panions beyond a projected distance of ∼ 30 AU. The presence of a brown dwarf within ∼ 20 AU is still compatible with obser- vations. 3.2. Numerical set-up We consider a ring of 150,000 massless planetesimals uniformly distributed between 70 and 140 AU (except otherwise specified) around a solar mass host star, with initial eccentricities randomly distributed between 0 and 0.05, and initial inclinations between ±3◦, while the remaining angles (longitudes of nodes and peri- astra) are randomly distributed between 0 and 2π. These values are summarized in Table 1. This reasonably well mimics the low eccentricities and inclinations expected at the end of the proto- planetary phase. The radial extent of the model disk has been configured to match closely the observed properties of the disk around ζ2 Reticuli. Using massless test particles removes both any self-gravity in the disk, and any back-reaction of the disk on the planet. In general, both of these phenomena are significant when the planet mass is comparable to the disk mass. We have no mass estimate for the debris disk of ζ2 Reticuli. However, a well-studied case is the debris disk of Fomalhaut, which mass was estimated to be ∼ 3 − 20M⊕ (Wyatt & Dent 2002; Chiang et al. 2009). Since a debris disk loses material over time due to the combined effects of collisional evolution, Poynting-Robertson drag and radiation pressure, and since ζ2 Reticuli is much older than Fomalhaut (440 Myr; Mamajek 2012), it is reasonable to assume that the debris disk surrounding ζ2 Reticuli should not contain more than a few Earth masses. In this case, it is also reasonable to assume that the disk self-gravity and back-reaction are negligible, and 5 Faramaz et al.: Can eccentric debris disks be long-lived? Fig. 3. Herschel/PACS images of ζ2 Reticuli at 70, 100 and 160 microns from left to right. North is up and East is left. The inset in the bottom-left corner shows the PSF. The North-West lobe is noted N-W, while the South-East lobe is noted S-E. the planet will still be able to put its imprint on the disk structure, if its mass is at least 0.1MJup ∼ 32M⊕. Parameters Number of particles Semi-major Axis (AU) Eccentricity Inclination (◦) Values 150,000 amin = 70 ; amax = 140 emin = 0 ; emax = 0.05 imin = −3 ; imax = 3 Table 1. Initial parameters for the planetesimals used in our N- body simulations We consider two planet-disk configurations, both inside and outside the planetesimal belt, performing parametric explo- rations of their influent orbital elements. 3.2.1. Inner perturber The first case is that of a planet interior to the initial ring. In this case, we consider that the inner edge of the disk is located at 70 AU, and make the classical assumption that it is truncated by the chaotic zone generated by coplanar planet (see, e.g., the approach used by Chiang et al. 2009). The chaotic zone is de- fined as where mean-motion resonances overlap, the width of this zone depending on the mass ratio between the central star and the perturber: aedge − aplanet = aplanet = 1.5µ2/7 ∆a aplanet where µ = mplanet/m∗ (Wisdom 1980; Duncan et al. 1989). Consequently, one can deduce the semi-major axis of a planet of a given mass which generates a disk inner edge at 70 AU: , (12) aplanet = aedge 1 + 1.5µ2/7 . (13) We choose to perform a parametric exploration of the mass and eccentricity of the perturber, fixing its semi-major axis to the value deduced from the formula above. The values explored are 0.1 MJup, 0.5MJup and 1 MJup for the mass and 0.2, 0.4 and 0.6 for the eccentricity (see Table 2). Here the disk initial inner edge is fixed halfway between 70 AU and the planet semi-major axis. In addition, we also considered the case of more massive perturber such as a brown dwarf, located further inside the sys- tem. This was motivated by the observation of a suspicious point 6 source on observations in Ks-band data from ESO archive (ID 086.C-0732(A); PI: Lohne,71574), which disappeared after re- reduction of the data, and consequently, since its existence is controversial, we chose not to show these images. However, if it were true, this point source was showing a 42 MJup brown dwarf located at a projected distance of 17.5 AU from the star. Therefore, we investigated the possibility for the presence of a brown dwarf in the inner parts of the system. For an inner edge to be produced at 70 AU, the planet's semi-major axis should be 43.8 AU. Since this constraint is less relevant for very eccentric inner perturbers (see Sect. 4), we choose here to set this value to the perturber's apastron rather than its semi-major axis. We fix its periastron to 17.5 AU, which leads to an orbital eccen- tricity of 0.43 and a semi-major axis of 30.6 AU. With such an orbit, analytical predictions indicate that the perturber should ex- cite planetesimals eccentricities up to 0.4. Therefore, this orbit is chosen to test numerically (see Table 2). 3.2.2. Outer perturber The second case considered is that of a planet exterior to the ring. There is indeed growing evidence for planets at large orbital separations, i.e several tens to few hundreds of AU away from their host star (see e.g. Luhman et al. 2007; Kalas et al. 2008; Marois et al. 2008, 2010) and the mass constraints set by direct imaging are loose given the age of the system (see Sect. 4.1.3). Therefore, we also investigate the ability of an external perturber to shape a disk. We consider coplanar outer planetary companions, and ex- plore the impact of the eccentricity, mass and periastron on the disk asymmetry. While an inner edge is in general considered as a clue for the presence of inner perturbers, it is obviously more delicate to assume that an outer edge is formed in the same manner, since a disk instrinsically has an outer limit. Therefore, the outer edge is not assumed here to be formed due to reso- nance overlap, and the planet periastron is fixed instead, to en- sure that it does not cross the disk. In order to explore a great variety of situations despite the CPU-time limitations, we con- sider a rough parameter space consisting in all the possible com- binations between masses mp = 0.1 − 1 − 2 MJup, periastrons qp = 150 − 200 − 250 AU and eccentricities ep = 0.2 − 0.4 − 0.6 (see Table 2). Note that the perturbers are being put on an initially eccentric orbit, which requires some discussion, since we assumed that the disk is initally symmetric. Indeed, this pictures a situation where the process putting the perturber on its eccentric orbit leaves the Faramaz et al.: Can eccentric debris disks be long-lived? considering the geometry of an ellipse: the offset δ of the cen- tre of symmetry from one of the focii of an ellipse is simply the product of its semi-major axis a by its eccentricity e, i.e., δ = ae. For a disk from our simulations, δ can be obtained by calculat- ing the centre of symmetry of the disk, using the positions of the test particles in the heliocentric frame: δ is the distance of this centre of symmetry to the star. The disk semi-major axis a is determined as follows: the disk is divided into superimposed angular sectors of 3◦. For each of these sectors, the radial distri- bution of the particles is fitted to a Gaussian. This gives us the radial position of the maximum density for each angular sector, and thus a set of points defining the global shape of the disk. It is then straightforward to retrieve a from this set of points by seek- ing for the major axis, i.e., the maximum distance between two opposite points. Finally, the disk eccentricity is simply e = δ/a. 4.1. Inner perturber We choose 4 illustrative results (see Table 2). For each, we show pole-on projections of the system at 1 Gyr on Fig. 7. We also summarize the outputs of our simulations in Table 2. Best candidates should have a significant orbital eccentricity of ∼ 0.4. The example of Case B is shown on Fig. 7 (upper right panel). However, scattering processes may be important, and their study allows us to place constraints on the mass of the perturber. Table 2. Summary of numerical experiments with an inner and an outer perturber, as well as for a brown dwarf and the stellar binary companion ζ1 Reticuli. Description of the disk at 1 Gyr: I) Steady state, e < 0.2, II) Steady state, e > 0.2, III) Scattered disk, IV) Resonant pattern V) Spiral pattern. The example cases highlighted further in Sect. 4 are labelled from A to G. Inner perturbers ep = 0.2 mp (MJup) ap (AU) 0.1 0.5 1 63.2 59.8 57.9 ep = 0.4 ep = 0.6 III,A II,B II III IV,C IV Outer perturbers mp (MJup) qp (AU) ep = 0.2 ep = 0.4 ep = 0.6 0.1 1 2 150 200 250 150 200 250 150 200 250 II,D V V II II I III,G I I V V,E V III,F II II III I II I I I I I V I I I I I I Other perturbers Perturber Brown Dwarf 42MJup ζ1 Reticuli 1M(cid:12) mp Orbital Parameters ap = 30.6 AU ; ep = 0.43 ap = 2046 AU ; ep = 0.815 Result II I 4.1.1. Scattered disks disk unperturbed. This may seem rather unrealistic, even though some scenarios can be envisaged. For instance, an inner per- turber could acquire its eccentricity via a planet-planet scattering event. However, this event should be such that a single perturber remains in the system. Otherwise, additional perturbations of a second planet would generate an orbital precession of the eccen- tric perturber, which in turn, could not sculpt the disk into an ec- centric shape. This is compatible with observational constraints, since as previously mentionned, radial velocity measurements rule out any massive perturber in the inner system. In the case of an outer perturber, an eccentric outer binary companion may be able to generate such initial conditions (see Sect. 6). In any case, retrieving realistic initial conditions relies on a complete study of the perturbations induced on the disk for multiple scenarios and, most probably, an extensive parameter space exploration. This study goes beyond the scope of the present paper, and shall be the subject of future work, which motivates our choice of simple initial conditions. 4. Numerical Investigation: Results We present here results obtained for both disk-planet set-ups we have considered: an inner and an outer planet, as well as for the case the perturber is the stellar companion ζ1 Ret. For each case, we will try to find the one that gives an eccentric disk compatible with observational constraints. Eiroa et al. (2010) gives a lower limit for the eccentricity of the disk in ζ2 Reticuli of 0.3. Therefore, given the uncertainties in the estimation of the disk global eccentricity we compute from our simulations, a disk global eccentricity lower than 0.2 is dis- carded in our analysis. This global eccentricity will be evaluated Inner perturbers may lead to very significant scattering pro- cesses. Namely, they can fill the inner parts of the disk instead of producing a well-defined ring. Such effects are presented with Case A on Fig 7 (upper left panel). This is an effect appearing in the presence of rather low-mass perturbers. As a matter of fact, such perturbers do not scatter material efficiently enough. This material tends then to populate the inner parts of the system. As a consequence, there is a lower mass limit for inner per- turbers, and in the specific case of ζ2 Reticuli system, this lower limit is between 0.1 and 0.5 MJup. However, one cannot exclude that another more massive planet produces scattering of the ma- terial, blowing it out and leaving an inner hole, so a more cor- rect way to express this constraint would be that perturbers with masses as low as 0.1 MJup should be accompanied by another more massive planet to create such a pattern. But this scenario presents difficulties: while this second planet must be massive enough to clear the inner parts of the system of its material, it also must have a limited dynamical effect on the first planet that sculpts the disk: this second planet must be distant and not too massive for the orbit of the first planet to remained unperturbed. Otherwise, this orbit would precess and no longer lead to an ec- centric pattern. This is not the purpose of this paper to investigate this scenario, but based on the previous arguments, it would most probably work in a very limited parameter space. 4.1.2. Resonant patterns It is notable that in the case of very eccentric (ep = 0.6) in- ner perturbers of mass between 0.5 and 1 MJup, resonant clumpy structures may arise. An example is Case C, shown on Fig 7 (bottom left panel). In Fig. 5, we show a semi-major axis vs eccentricity diagram of the disk: it reveals two populations in resonance with the planet, namely the 3:2 and 2:1 mean-motion 7 Faramaz et al.: Can eccentric debris disks be long-lived? the evolution of the disk offset on 1 Gyr on Fig. 9. The outputs of our simulations are summarized in Table 2. 4.2.1. Good candidates Case D is shown on Fig. 8 (upper left panel). The corresponding evolution of the offset clearly shows that the disk is at steady state and significantly eccentric (see Fig. 9). The best candi- dates must have significant eccentricities: none of our 0.2 eccen- tric perturbers manages to create the desired eccentric structure, even in the limit case predicted analytically where a 0.1 MJup perturber has a pericentre qp = 150 AU and eccentricity ep = 0.2 (see Fig. 2). As a consequence, the best outer candidates have eccentricities ∼ 0.4 − 0.6. 4.2.2. Spiral patterns Case E is shown on Fig. 8 (upper right panel). The system shows a spiral pattern at 1 Gyr, which, according to analytical pre- dictions, corresponds to one precession timescale (see Fig. 2). Thus, our N-body integrations confirm what was noted by Wyatt (2005), i.e., the analytical formula appears to be a lower limit and several precession timescales are sometimes necessary for spiral patterns to vanish. The effect of spirals can also be seen on the evolution of the disk offset, which undergoes oscillations (see Fig. 9). If we had observational proof that the disk of ζ2 Reticuli has reached a steady-state and contains no spiral pattern, then the results of our numerical investigation would allow us to place a lower mass limit of 0.1 MJup on an outer perturber in a range of periastrons from 150 to 250 AU, based on dynamical timescales criterion, otherwise it takes longer than 1 Gyr to generate a steady-state eccentric disk. This limit still holds for larger pe- riastrons than the range explored, since dynamical timescales increase with distance. But here, since the slightly edge-on ori- entation of the disk and the resolution of the Herschel/PACS im- ages, it is extremely difficult to discard the presence of any spiral pattern in this disk, and no lower mass limit can be clearly put on an outer perturber. 4.2.3. Scattered disks Outer perturbers may lead to very significant scattering pro- cesses. We present such effects in Fig 8 (lower panels), where Case F and Case G are considered. One can see that these processes are even more significant when the mass of the perturber increases. Indeed, very logically, more massive perturbers tend to scatter small bodies more effi- ciently. The distance to the disk plays a major part in this effect too, since close perturbers also tend to scatter more material. Additionally, when a perturber is on an eccentric orbit, it ap- proaches even closer to the disk. In the most dramatic cases, the disk is completely destroyed. Consequently, for a given distance to the disk, there is an upper limit to the mass of an outer com- panion. For the Case F, one might question the contribution of the scattered inner material to the emission of the disk, and whether it would be visible on resolved images. In this case, the potential visibility of material on real observations rely on the sensitiv- ity and resolution of the instrument used for these observations, as well as on the distance of the object, the radiative properties of the material itself and the quantity of light it receives, i.e. on the host star properties. Therefore, only the production of syn- Fig. 5. Semi-major axis versus eccentricity diagram of the disk at 1 Gyr for the Case C perturber. Overdensities of planetesimals at ∼ 79 and ∼ 95 AU correspond to planetesimals respectively in 3:2 and 2:1 mean-motion resonance with the perturber. resonances, at ∼ 79 AU and ∼ 95 AU, respectively. These reso- nant populations are put in evidence in Fig. 6. Usually, the pres- ence of a population in 3:2 resonance is associated with capture during outward planetary migration, as it is the case in our own Solar System (Malhotra 1993, 1995). But here, the appearance of these structures is most probably due to the fact that we used the chaotic zone formula (Eq. 13) to determine the perturber semi- major axis, which was derived for perturbers on circular orbits. Therefore, we should expect that this formula works less effi- ciently with increasing orbital eccentricity of the perturber: the result is that the planet digs into the disk, and consequently, plan- etesimals unprotected against close encounters by mean-motion resonance are scattered out, leaving the resonant structures ap- parent. This is supported by the fact that these resonant struc- tures disappear if the constraint given by the chaotic zone for- mula is applied to the perturber's apastron instead of its semi- major axis, as was done in the case of an inner brown dwarf- type companion. Interestingly, the observation of such resonant structures in a system may provide other clues on the dynamical history of a perturber than an outward migration: it could mean that the planet was originally shaping the inner edge of the disk before being put on an eccentric orbit. However, these are thin structures, and if the disk is seen close to edge-on, as is the case for ζ2 Reticuli, these would most probably be hidden by the non resonant population. 4.1.3. Brown dwarf Additionally, we investigated the possibility for the presence of a brown dwarf in the inner parts of the system, on an orbit such that the perturber should excite planetesimals eccentricities up to 0.4. The disk at 1 Gyr is shown on Fig 7 (bottom right panel). Its global eccentricity is ∼ 0.2 − 0.25, which shows that very massive perturbers in the inner parts of the system can create the wanted pattern. 4.2. Outer perturber We choose 4 illustrative results (see Table 2), and for each, show pole-on projections of the system at 1 Gyr on Fig. 8, along with 8 Faramaz et al.: Can eccentric debris disks be long-lived? Fig. 6. Views from above the plane of disks at 1 Gyr in Case C, i.e., where resonant patterns appear. Planetesimals in 3:2 (left) and 2:1 (right) mean-motion resonance with the perturber are put in evidence in red. Fig. 7. Example views from above the plane of disks at 1 Gyr, under the influence of inner perturbers.TOP Left: the perturber scatters material in the inner parts of the system (Case A). Right: one of the best candidates (Case B). BOTTOM Left: resonant patterns appear (Case C). Right: a brown-dwarf perturber can be a good candidate. 9 Faramaz et al.: Can eccentric debris disks be long-lived? thetic images for direct comparison with observations can reveal whether this material is apparent or not, and refine constraints on the potential perturbers at work in the system. Additionally, the evolution of the offset clearly suggests that the asymmetry re- laxes asymptotically to a small but non-zero value, and indeed, the apparent ring structure appears to show little eccentricity. More specifically concerning the ζ2 Reticuli system, our re- sults allow us to place an upper mass limit of 2 MJup on an outer perturber in a range of periastrons from 150 to 250 AU. However, one should note that this upper mass limit is expected to increase for periastrons greater than those explored, since scattering processes are expected to be less efficient for a given if the companion is further from the disk. that almost 25% of binaries with orbital periods greater than 103 days have orbital eccentrities e(cid:63) = 0.825 ± 0.075. In the present case, d = 3713 AU and e(cid:63) = 0.815 lead to a(cid:63) = 2046 AU and an orbital period T(cid:63) ∼ 105 yrs. Thus, ζ1 Reticuli being on an eccentric orbit, if not a highly eccentric one, is in fact pos- sible. However, the derived orbit should also have an apoastron value of q(cid:63) = 379 AU and one might question the disk survival at ∼ 70 − 120 AU with a stellar-type perturber approaching so close to the system. Eq. 1 of Holman & Wiegert (1999) gives the critical semi- major axis acrit for orbital stability around a star perturbed by a binary. This is: 4.3. Stellar binary companion We investigate here the influence of the binary companion ζ1Reticuli on the debris disk surrounding ζ2 Reticuli. The aim is to determine whether it alone could generate the observed asymmetry. If this is to be the case, the companion must be on an eccentric orbit. The only observational constraint available so far on the relative orbit of the binary system is the projected distance between the two stars, namely 3713 AU (Mason et al. 2001). Consequently, we cannot exclude the possibility that its orbit is eccentric, with a present-day location near apastron. We investigate whether a binary companion on an eccentric orbit at such a distance could account alone for the observed structure of the disk, i.e., an elliptic ring of minimum global eccentricity 0.3, without any constraint on the binary eccentricity. Both analytical and numerical methods are used. We first examine the forced secular eccentricity applied to the planetesimals by an eccentric binary companion, which is assumed to be coplanar. The debris disk surrounding ζ2 Reticuli is approximately centered on a = 100 AU, and the binary per- turber is at 3713 AU from ζ2 Reticuli. This is only a projected distance, not a semi-major axis. Consider that the binary com- panion is currently located at distance r(cid:63) from ζ2 Reticuli. The equation of its orbit around ζ2 Reticuli reads: r(cid:63) = a(cid:63)(1 − e2 (cid:63)) 1 + e(cid:63) cos v(cid:63) , (14) where v(cid:63), a(cid:63) and e(cid:63) are the binary companion current true anomaly, semi-major axis, and orbital eccentricity, respectively. The observed distance d = 3713 AU is related to r(cid:63) by d = r(cid:63) cos ψ, where ψ is a projection angle. This gives: a(cid:63) = d(1 + e(cid:63) cos v(cid:63)) (1 − e2 (cid:63)) cos ψ . (15) Now, from this result Eq. 15 and Eq. 2, with ap and ep being substituted by a(cid:63) and e(cid:63): 2ef (cid:39) 5 2 1 + e(cid:63) cos v(cid:63) e(cid:63) cos ψ (16) a d . It is clear from this equation that the highest possible ef values will be obtained for cos v(cid:63) = −1 (binary currently at apoastron) and cos ψ = 1 (no projection factor). With these assumptions, one derives: 2ef,max (cid:39) 0.068 For 2ef,max to reach at least 0.3, e(cid:63) ≥ 0.815 is required. This seems highly eccentric and very unlikely at first sight. Yet, Duquennoy & Mayor (1991, see their Fig. 6.b) have shown e(cid:63) 1 − e(cid:63) (17) . 10 acrit = [(0.464 ± 0.006) + (−0.380 ± 0.010)µ +(−0.631 ± 0.034)e(cid:63) + (0.586 ± 0.061)µe(cid:63) + (−0.198 ± 0.074)µe2 +(0.150 ± 0.041)e2 (cid:63) (cid:63)]a(cid:63) ,(18) where µ = m(cid:63)/(mζ2 Reticuli + m(cid:63)) is the star mass ratio of value 1/2 if we assume here m(cid:63) = mζ2 Reticuli = 1M(cid:12). Material with a ≥ acrit will be on an unstable orbit, and most probably scattered out of the system. to our case leads to acrit = 66+236−66 AU. Uncertainties on acrit are rather large, and this result shows that within uncertainties, the disk could exist at the observed dis- tances, or, on the contrary, not exist at all. Therefore, this orbit is tested numerically. Applying it We consider a ring of 150,000 massless planetesimals uni- formly distributed between 70 and 140 AU from their solar mass host star, with proper initial eccentricities randomly distributed between 0 and 0.05, and initial inclinations between ±3◦. The test particles are perturbed by another solar mass star in orbit around the primary with a semi-major axis 2046 AU and eccen- tricity 0.815, coplanar to the disk. We use the SWIFT-RMVS N- body symplectic code of Levison & Duncan (1994) to compute their orbital evolution for 2 Gyr. Our results are consistent with the predictions of Holman & Wiegert (1999): the disk is truncated at ∼ 80 AU (see Fig. 10, left panel), which is incompatible with observational constraints of the debris disk of ζ2 Reticuli, since it radially extends up to ∼ 120 AU. Moreover, at 2 Gyr, the disk has a rather symmetric shape, and no clear offset from the star (see Fig. 10, right panel). These results suggest that the binary companion on an eccentric orbit can unlikely account for the disk structure, and that an unseen eccentric companion is more likely responsible for shaping the disk. Another possible way for a stellar binary companion to gen- erate high eccentricity orbiting bodies around a primary is the Kozai mechanism (Kozai 1962). This would happen if the disk and the binary orbit were mutually inclined by more than ∼ 39◦. In that case, the orbit of a particle in the disk would suffer coupled modulations in inclination and eccentricity, i.e., a par- ticle would periodically switch from a highly eccentric orbit, coplanar with the binary companion's orbit, to a circular or- bit, very inclined with respect to the orbit of the binary com- panion. However, we can discard this mechanism from being a possible explanation for the eccentric global structure of the disk of ζ2 Reticuli: indeed, the Kozai hamiltonian is invariant by rotation, i.e., if this mechanism is able to excite planetesimals eccentricities to high values, their longitudes of periastrons re- main uniformely distributed, while they should be preferentially aligned for an eccentric and offset disk to be generated. Faramaz et al.: Can eccentric debris disks be long-lived? Fig. 8. Example views from above the disk plane at 1 Gyr, under the influence of outer perturbers. TOP Left: one of the best candidates (Case D). Right: spirals are still apparent (Case E). BOTTOM Left: one might question the contribution of the scattered material to the disk emission (Case F). Right: scattering processes tend to destroy the structure (Case G). 5. Synthetic images In this section, we will use the results of our best fit N-body simulations as "seeds" to produce a realistic dust population and synthetic images for comparison with Herschel/PACS ob- servations. The procedure followed to create synthetic images is straightforward: a population of dust is created from the position of the parent planetesimals. The main difference between dust particles and planetesi- mals is that the former are small enough to be affected by stellar radiation pressure. Radiation pressure is usually described for a particle via its constant ratio β to stellar gravity. The dust parti- cles are supposed to be released by planetesimals which do not feel any radiation pressure. Hence the daugther particles assume an orbit that is very different from that of their parent bodies. It is well known that if the parent bodies move on circular orbits, the dust particles are unbound from the star as soon as β ≥ 0.5. In our case, however, dust particles may be released by plan- etesimals orbiting on more or less eccentric orbits, which may change this threshold slightly. As planetesimal eccentricities are expected to be moderate on average, β = 0.5 can nevertheless be considered as a reasonable approximation. Small grains are released from seed planetesimal positions, at the planetesimal velocity, and are then spread along the or- bits determined by these initial conditions and their β value. We are aware that this simple procedure cannot accurately evaluate the spatial distribution of the smallest grains. To do so, complex models, such as the DyCoSS code of Th´ebault (2012), the CGA of Stark & Kuchner (2009), or the LIDT-DD ode by Kral et al. (2013), have to be used to evaluate the complex interplay be- tween the rate at which grains at collisionally produced from par- ent planetesimals, the time they spend (because of their highly eccentric orbits) in empty collisionally inactive regions and the rate at which they can be affected or even ejected by close en- counters with the perturbing planet (see Sect. 4 of Thebault et al. 2012), not to mention the Poynting-Robertson drag these small grains are subject to. However, this caveat is acceptable for the present problem, because the role played by small micron-sized grains close to the blow-out size is very minor at wavelengths > 70 µm, so that our 11 Faramaz et al.: Can eccentric debris disks be long-lived? Fig. 9. Example of the time evolution on 1 Gyr of the disk offset coordinates (X (solid line) and Y (dashed line)) for outer perturbers. TOP Left: one of the best candidates (Case D). Since the orientation of the perturber orbit, i.e., its semi-major axis is along the X- axis with positive periastron, this corresponds, for a disk located at ac ∼ 100 AU, to a negative X-offset ∆X ∼ −30 AU and a zero Y-offset. Right: the oscillations of the offset coordinates reveal the spiral-winding regime (Case E). BOTTOM The offset seems to relax. Left: Case F. Right: Case G. Fig. 10. Semi-major axis vs eccentricity diagram (left) and pole-on projection(right) of a disk of planetesimals after 2 Gyr under the influence of a one solar mass star of semi-major axis 2046AU on an orbit of eccentricity 0.815. The disk is truncated at ∼ 80 AU and nearly circular. It has no clear offset: its center of symmetry is offset of ∼ 6 AU from the star along the perturber major axis. With a disk mean semi-major axis of ∼ 70 AU, this gives a global eccentricity of ∼ 0.08. synthetic images will not be strongly affected by errors regarding their spatial distribution. We set the dust grain sizes range from 0.5 µm to 1 mm, with a classical Dohnanyi (1969) power-law distribution (index -3.5), which covers well the β distribution from 0 to 0.5, since this parameter depends on grain size. Their emission is then computed using the radiative transfer code GRaTeR (see, e.g., Lebreton et al. 2012). To do this, the following parameters are required: distance of the star (12 pc), magnitude in band V (V = 5.24), and total luminosity 0.96 L(cid:12). Because the disk is optically thin, its mass is linked linearly to the flux emission intensity and it can be easily scaled to fit with the intensity observed (see Table 3). The mass needed for the disk to produce a flux as observed on Herschel/PACS will vary with the dust grain composition, and thus its density. But since we do not have constraints on the dust composition, as- 12 Faramaz et al.: Can eccentric debris disks be long-lived? Wavelength (µm) Stellar Flux (mJy) Disk Flux (mJy) 70 100 160 24.9 ± 0.8 13.4 ± 1.0 19.4 ± 1.5 8.9 ± 0.8 13.5 ± 1.0 ∼ 14.7 ∼ 4.7 Table 3. Stellar and disk fluxes at 70, 100 and 160 µm (Eiroa et al. 2010). The fluxes at 70 and 100 µm, along with the total flux of the star-disk system at 160 µm are PACS measurements. The individual star and disk fluxes at 160 µm result from predic- tions. trosilicate grains are used (Draine 2003), and the mass of the disk is simply scaled to obtain intensities compatible with obser- vational constraints for a given wavelength. Thermal emission images are produced with a resolution of 1(cid:48)(cid:48)/pixel at 70 and 100 µm, and 2(cid:48)(cid:48)/pixel at 160 µm. Before con- volving these images with the PSF, the star is added at the central pixel, with a flux intensity matching the predicted stellar photo- sphere flux density in each waveband. The position angles of the disk observed with Herschel/PACS, and of the disk in our syn- thetic images, but also the orientation of the telescope during the observations are taken into account (see Table 4). Our purpose is to match the observations. Disk observed Disk simulated PSF PSF ζ2 Reticuli ζ2 Reticuli - - Wavelength (µm) PA(◦) 110 110 127 127 281 281 70/160 100/160 70/160 100/160 Table 4. Disk position angle observed with Herschel/PACS, disk position angle on our synthetic images, and telescope orientation during Herschel/PACS observations and PSF. We choose among our simulations one that lead to a clear and significantly eccentric disk at 1 Gyr, namely our Case A (see Table 2, and upper left panels of Fig. 8 and Fig. 9), and aim to reproduce the Herschel/PACS image at 100 µm. The mass of the disk is scaled so that a total flux of 13.5 mJy as observed with Herschel/PACS at 100 µm is spread all over the disk. The last parameter needed is the system inclination. Our best fit gives 65.5◦ (see Appendix B). We present an unconvolved and convolved image of this disk on Fig. 11. The disk offset is clearly visible on unconvolved im- ages. However, there seem to be no difference between the sym- metric and asymmetric state once the images are convolved, be- cause of the contrast between the star and the disk emission. Moreover, the disk flux per pixel is one order of magnitude smaller than the fluxes observed in Herschel/PACS images. This is, however, not surprising: in order to estimate the total disk flux, the flux was measured in a small region of the disk before applying aperture correction. Even if the correct aperture correction for a point source was used, that would always be a lower limit, and the total disk flux will be underestimated. The parent ring also may be narrower, which would increase the flux per pixel. Therefore, we investigate the impact of the width and of the total flux of the disk on the features visible with PACS, and pro- duce convolved images of a dusty disk produced by an asymmet- ric eccentric parent ring of diverse total fluxes (1, 2 and 5 times the flux measured by Herschel/PACS), and of diverse widths (semi major axis centered on 100 AU, widths 5, 10 and 20 AU). To do so, particles from a range of semi-major axis from our N- body simulation output are being selected . An inclination angle of 65◦ is chosen, which is the most probable inclination derived for the system. The fluxes per pixel recovered on convolved images with a disk five times more massive than the mass initially derived from observations are more compatible with the observations. This provides a better constraint on the disk mass and total flux. We show this example in Fig. 12 (left panel): one can see that the asymmetric double lobe structure now appears clearly. It is worth noting that the width of the parent ring has a lim- ited influence on the flux per pixel compared to the mass of the disk. But it has an influence on the appearance of the disk: a narrow parent ring (< 10 AU) leads to the apastron lobe to be resolved, which is more consistent with Herschel/PACS images, although this lobe does not seem to be located as far from the star as it is on the Herschel/PACS images. Therefore, we also investigate the location of the disk, by producing convolved images of a ring of dust produced by a nar- row eccentric parent ring of width 5 AU and semi-major axis centered on 120, 130 and 140 AU . The disk total flux is set to be five times the flux derived from observations. The inclination angle is here again set to 65◦. As expected, the further the disk is located, the clearer the lobes appear (see Fig. 12). Finally, with all these insights, we conclude that the hypoth- esis of an eccentric dusty disk around ζ2 Reticuli is indeed com- patible with Herschel/PACS images, provided that the dust is produced by a narrow parent ring with width less than 10 AU and located slightly further away than derived by Eiroa et al. (2010), i.e., it has a semi-major axis distribution centred between 120 and 140 AU. This slightly changes the constraints derived on potential perturbers, but the forced eccentricity depends only on the ra- tio between the planet and planetesimals semi-major axis, and in a linear way at lowest order approximation. This means that in this approximation, the constraints can be completely scaled in a linear way, i.e., the potential semi-major axis for planets must also be increased by 20 -- 40 % and the disk should rather be centered at 120 -- 140 AU, while the constraint on the perturber eccentricity (e (cid:38) 0.3), remains identical. 6. Discussion & Conclusion In this paper, the case of ζ2 Reticuli is used as an example to discuss the shaping of a disk into an eccentric ring on Gyr timescales. We show that eccentric patterns in debris disks can be main- tained on Gyr timescales, but also that eccentric perturbers can produce other patterns than eccentric rings. The general results of our simulations show that both inner and outer perturbers can generate extremely significant scatter- ing processes. They can lead a disk to adopt structures that do not show any clear elliptic ring, namely the inner part of the disk is filled or the structure destroyed. These scattering processes endanger the survival of an eccentric ring and their investiga- tion with numerical experiments allows us to put constraints on potential perturbers. From these constraints, we derive an upper mass limit for outer perturbers in a certain range of periastrons, and a lower mass limit for inner perturbers. 13 Faramaz et al.: Can eccentric debris disks be long-lived? Fig. 11. Synthetic images at λ = 100 µm, unconvolved disk (no star) (left) and convolved disk and star image with the PSF (right). Disk at 1 Gyr in Case A (see Table 2), seen with inclination 65◦. The star is at the centre of the image, and on the convolved image, the flux scale is set to match the one on Herschel/PACS image. Fig. 12. Synthetic images of Case A (see Table 2), at 100 µm, and convolved with the PSF. The parent ring is centred at 100 AU (left) and 130 AU (right), has a width of 5 AU and is seen with an inclination of 65◦. The disk total flux is 5 times the flux measured by Herschel/PACS. The star is at the centre of the image, and the flux scale is set to match that of Herschel/PACS image. Moreover, the timescale for spiral structures to vanish is longer with smaller mass perturbers, and thus, investigation of this timescale with numerical experiments permits to place a lower mass limit on perturbers, provided that the presence of spiral structures in a disk can definitely be ruled out from obser- vations. The role of numerical experiments is crucial here, since the analytical timescale at the disk centre of distribution under- estimates the effective timescale in a real extended disk. The offset of the disk centre with respect to the star is mostly stable; however, we note that in rare cases, it seems to relax very slowly. While the former evolution is characteristic for pericen- tre glow dynamics, the latter one is surprising. The relaxation of the eccentric structure is not expected in the first order secular analysis described by Wyatt et al. (1999) and Wyatt (2005). It may be the result of higher order terms that have been neglected in the analytical study, and more probably to erratic short-term variations of the planetesimals' semi-major axes due to moder- ately distant approaches to the planet. These effects that can lead to scattering of the planetesimals are eliminated in the analytical averaging process of the perturbations, and thus cannot be pre- dicted analytically in the secular approximations used here. A more detailed study of this relaxation phenomenon of pericentre glow structures is nevertheless beyond the scope of the present paper, and will be the purpose of future work. Transient spiral structures, filled inner holes, sparesly pop- ulated scattered disks, and resonant clumpy structures are all 14 Faramaz et al.: Can eccentric debris disks be long-lived? possible outcomes when an eccentric perturber acts on a debris disk. They can be put in evidence with numerical simulations, but more importantly, used to put constraints on a perturber in a system (mass, eccentricity). Therefore, we provide a method to investigate and model eccentric ring structures, based on a complementary analytic and numerical approach, where one can derive potential orbits from analytics and put them to test numer- ically using N-body codes. This method can be easily applied to other systems, and it is expected to be useful in a near future. Indeed, Kaib et al. (2013) have pointed at the fact that wide binary star systems, i.e systems with separations greater than 1000 AU, can produce eccentric planets around a primary star on Gyr timescales. This is due to Galactic tides and passing stars perturbations, which are able, sooner or later, to put the sec- ondary star on a highly eccentric orbit. The proportion of wide binary systems is by no mean negligible (∼ 50%, Duquennoy & Mayor 1991), and although Gyr-old debris disks are faint and difficult to detect, this will be overcome with the unique capa- bilities of ALMA, JWST and SPICA. Therefore, old eccentric patterns in debris disks are expected to be commonly observed in the future. The ζ2 Reticuli disk is one such example of Gyr-old eccentric debris disk. Moreover, ζ2 Reticuli is part of a wide binary star system, which may provide an explanation for the presence of an eccentric perturber around ζ2 Reticuli. We show that the binary companion cannot be directly responsible for the eccentric ring structure, and show that the asymmetry is rather due to a closer companion, either interior or exterior to the disk. In all cases, the eccentric companion should have an eccentricity e (cid:38) 0.3 to produce such a pattern. Investigation of the disk structure due to scattering processes provides an upper mass limit of 2 MJup for an outer perturber located in a range of periastrons 150 − 250AU, whereas a lower mass limit of 0.1 MJup is associated with inner perturbers in the ζ2 Reticuli system. By producing synthetic images, we show that the original in- terpretation of the double-lobed feature around ζ2 Reticuli, i.e., the observed eccentric ring e (cid:38) 0.3, is clearly supported, al- though the disk is located slightly further (20-40%) than orig- inally derived. Moreover, we find that the dusty disk should be created by a narrow parent ring (width <10 AU), which should have a slight inclination with the line of sight, compatible with the most probable inclination derived for the system, and it should also have a significantly greater flux than that estimated from the Herschel/PACS measurements (at least five times). Aknowledgements: We thank the referee, A. Mustill, for very useful comments that contributed to this paper. Computations presented in this paper were performed at the Service Commun de Calcul Intensif de l'Observatoire de Grenoble (SCCI) on the super- computer funded by Agence Nationale pour la Recherche un- der contracts ANR-07-BLAN-0221, ANR-2010-JCJC-0504-01 and ANR-2010-JCJC-0501-01. B. Montesinos, C. Eiroa and J.P. Marshall are supported by Spanish grant AYA 2011-26202. A. Bonsor and S. Ertel acknowledge the support of the ANR- 2010 BLAN-0505- 01 (EXOZODI). The authors wish to thank the PNP/CNES for their financial support. This work has also greatly benefited from the software resulting from Thomas Tintillier's training project. References Allard, F., Homeier, D., & Freytag, B. 2011, in Astronomical Society of the Pacific Conference Series, Vol. 448, 16th Cambridge Workshop on Cool Stars, Stellar Systems, and the Sun, ed. C. Johns-Krull, M. K. Browning, & A. A. West, 91 Augereau, J. C., Nelson, R. P., Lagrange, A. M., Papaloizou, J. C. B., & Mouillet, D. 2001, A&A, 370, 447 Augereau, J. C. & Papaloizou, J. C. B. 2004, A&A, 414, 1153 Aumann, H. H., Beichman, C. A., Gillett, F. C., et al. 1984, ApJ, 278, L23 Backman, D. E. & Paresce, F. 1993, in Protostars and Planets III, ed. E. H. Levy Baraffe, I., Chabrier, G., Barman, T. S., Allard, F., & Hauschildt, P. H. 2003, & J. I. Lunine, 1253 -- 1304 A&A, 402, 701 Beust, H., Augereau, J.-C., Bonsor, A., et al. 2013, in prep. Beust, H. & Dutrey, A. 2006, A&A, 446, 137 Boley, A. C., Payne, M. J., Corder, S., et al. 2012, ApJ, 750, L21 Chiang, E., Kite, E., Kalas, P., Graham, J. R., & Clampin, M. 2009, ApJ, 693, Dohnanyi, J. S. 1969, J. Geophys. Res., 74, 2531 Draine, B. T. 2003, ARA&A, 41, 241 Duncan, M., Quinn, T., & Tremaine, S. 1989, Icarus, 82, 402 Duquennoy, A. & Mayor, M. 1991, A&A, 248, 485 Eiroa, C., Fedele, D., Maldonado, J., et al. 2010, A&A, 518, L131 Eiroa, C., Marshall, J. P., Mora, A., et al. 2013, A&A, 555, A11 Ertel, S., Wolf, S., Eiroa, C., et al. 2011, in EPSC-DPS Joint Meeting 2011, 678 Ertel, S., Wolf, S., & Rodmann, J. 2012, A&A, 544, A61 Ford, E. B., Kozinsky, B., & Rasio, F. A. 2000, ApJ, 535, 385 Gray, R. O., Corbally, C. J., Garrison, R. F., et al. 2006, AJ, 132, 161 Guilloteau, S., Dutrey, A., Pi´etu, V., & Boehler, Y. 2011, A&A, 529, A105 Henry, T. J., Soderblom, D. R., Donahue, R. A., & Baliunas, S. L. 1996, AJ, 111, 734 439 Holman, M. J. & Wiegert, P. A. 1999, AJ, 117, 621 Johnson, H. L., Mitchell, R. I., Iriarte, B., & Wisniewski, W. Z. 1966, Communications of the Lunar and Planetary Laboratory, 4, 99 Kaib, N. A., Raymond, S. N., & Duncan, M. 2013, Nature, 493, 381 Kalas, P., Graham, J. R., Chiang, E., et al. 2008, Science, 322, 1345 Kalas, P., Graham, J. R., & Clampin, M. 2005, Nature, 435, 1067 Kalas, P., Graham, J. R., Fitzgerald, M. P., & Clampin, M. 2013, ArXiv e-prints Kozai, Y. 1962, AJ, 67, 591 Kral, Q., Th´ebault, P., & Charnoz, S. 2013, A&A, 558, A121 Krist, J. E., Stapelfeldt, K. R., Bryden, G., & Plavchan, P. 2012, AJ, 144, 45 Krivov, A. V. 2010, Research in Astronomy and Astrophysics, 10, 383 Krymolowski, Y. & Mazeh, T. 1999, MNRAS, 304, 720 Lebreton, J., Augereau, J.-C., Thi, W.-F., et al. 2012, A&A, 539, A17 Lenzen, R., Hartung, M., Brandner, W., et al. 2003, in Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series, Vol. 4841, Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series, ed. M. Iye & A. F. M. Moorwood, 944 -- 952 Levison, H. F. & Duncan, M. J. 1994, Icarus, 108, 18 Lohne, T., Krivov, A. V., & Rodmann, J. 2008, ApJ, 673, 1123 Luhman, K. L., Patten, B. M., Marengo, M., et al. 2007, ApJ, 654, 570 Malhotra, R. 1993, Nature, 365, 819 Malhotra, R. 1995, AJ, 110, 420 Mamajek, E. E. 2012, ApJ, 754, L20 Mardling, R. A. & Lin, D. N. C. 2002, ApJ, 573, 829 Marois, C., Lafreni`ere, D., Doyon, R., Macintosh, B., & Nadeau, D. 2006, ApJ, 641, 556 Marois, C., Macintosh, B., Barman, T., et al. 2008, Science, 322, 1348 Marois, C., Zuckerman, B., Konopacky, Q. M., Macintosh, B., & Barman, T. Masana, E., Jordi, C., & Ribas, I. 2006, A&A, 450, 735 Mason, B. D., Wycoff, G. L., Hartkopf, W. I., Douglass, G. G., & Worley, C. E. 2010, Nature, 468, 1080 2001, AJ, 122, 3466 Mayor, M., Pepe, F., Queloz, D., et al. 2003, The Messenger, 114, 20 Mayor, M. & Queloz, D. 1995, Nature, 378, 355 Moro-Mart´ın, A. 2012, ArXiv e-prints Moro-Mart´ın, A. & Malhotra, R. 2002, AJ, 124, 2305 Mouillet, D., Larwood, J. D., Papaloizou, J. C. B., & Lagrange, A. M. 1997, MNRAS, 292, 896 Mustill, A. J. & Wyatt, M. C. 2009, MNRAS, 399, 1403 Noyes, R. W., Hartmann, L. W., Baliunas, S. L., Duncan, D. K., & Vaughan, A. H. 1984, ApJ, 279, 763 Quillen, A. C. 2006, MNRAS, 372, L14 Reche, R., Beust, H., Augereau, J.-C., & Absil, O. 2008, A&A, 480, 551 Reiners, A. & Schmitt, J. H. M. M. 2003, A&A, 398, 647 Rousset, G., Lacombe, F., Puget, P., et al. 2003, in Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series, Vol. 4839, Society of 15 Faramaz et al.: Can eccentric debris disks be long-lived? Photo-Optical Instrumentation Engineers (SPIE) Conference Series, ed. P. L. Wizinowich & D. Bonaccini, 140 -- 149 Shaya, E. J. & Olling, R. P. 2011, ApJS, 192, 2 Soummer, R., Pueyo, L., & Larkin, J. 2012, ApJ, 755, L28 Stapelfeldt, K. R., Holmes, E. K., Chen, C., et al. 2004, ApJS, 154, 458 Stapelfeldt, K. R., Krist, J. E., Bryden, G. C., & Plavchan, P. 2012, in American Astronomical Society Meeting Abstracts, Vol. 220, American Astronomical Society Meeting Abstracts 220, 506.03 Stark, C. C. & Kuchner, M. J. 2008, ApJ, 686, 637 Stark, C. C. & Kuchner, M. J. 2009, ApJ, 707, 543 Th´ebault, P. 2012, A&A, 537, A65 Th´ebault, P. & Augereau, J.-C. 2007, A&A, 472, 169 Thebault, P., Kral, Q., & Ertel, S. 2012, A&A, 547, A92 Th´ebault, P., Kral, Q., & Ertel, S. 2012, ArXiv e-prints Torres, C. A. O., Quast, G. R., da Silva, L., et al. 2006, A&A, 460, 695 Trilling, D. E., Bryden, G., Beichman, C. A., et al. 2008, ApJ, 674, 1086 van Leeuwen, F. 2007, A&A, 474, 653 Watson, C. A., Littlefair, S. P., Diamond, C., et al. 2011, MNRAS, 413, L71 Wisdom, J. 1980, AJ, 85, 1122 Wyatt, M. C. 1999, PhD thesis, Royal Observatory, Blackford Hill, Edinburgh The Search for Other Worlds, ed. S. S. Holt & D. Deming, 93 -- 102 Wyatt, M. C. 2005, A&A, 440, 937 Wyatt, M. C. & Dent, W. R. F. 2002, MNRAS, 334, 589 Wyatt, M. C., Dermott, S. F., & Telesco, C. M. 2000, in Astronomical Society of the Pacific Conference Series, Vol. 219, Disks, Planetesimals, and Planets, ed. G. Garz´on, C. Eiroa, D. de Winter, & T. J. Mahoney, 289 Wyatt, M. C., Dermott, S. F., Telesco, C. M., et al. 1999, ApJ, 527, 918 EH9 3HJ, UK Wyatt, M. C. 2004, in American Institute of Physics Conference Series, Vol. 713, 16 Appendix A: SPIRE images of ζ2 Reticuli Faramaz et al.: Can eccentric debris disks be long-lived? Fig. A.1. Herschel/SPIRE images of ζ2 Reticuli at 250, 350 and 500 µm (left -- right). The images were reduced using the standard reduction scripts ofn HIPE, version 8.2 and SPIRE CAL 8.1. Image orientation is North up, East left. The pixel scales are 6(cid:48)(cid:48), 10(cid:48)(cid:48) and 14(cid:48)(cid:48) at 250, 350 and 500 µm, respectively. The SPIRE beam FWHM in each band is denoted by the white circle in the bottom left corner of each image. Wavelength (µm) Stellar Flux (mJy) Disk+Star Flux (mJy) 250 350 500 59.72 ± 6.70 24.68 ± 6.89 20.29 ± 7.66 2.03 ± 0.03 1.04 ± 0.02 0.51 ± 0.10 Table A.1. Stellar predicted fluxes and SPIRE fluxes measurements of the star-disk system at 250, 350 and 500 µm (Eiroa et al. 2013). Appendix B: Inclination of ζ2 Reticuli Observations of the debris disk surrounding ζ2 Reticuli reveal a double-lobed asymmetric feature. The inclination of this system relative to the line of sight is a key parameter to interprete correctly the observations. If the system is seen pole-on, one would rather expect the observed feature to be the signature of resonant clumps, whereas an eccentric ring signature would be more plausible if the system is observed edge-on. In general, observations suggest that stellar and disk inclinations are aligned (Watson et al. 2011; Guilloteau et al. 2011). Under this assumption, one can estimate the disk inclination from the observed stellar inclination. Consequently, we aim here to measure the star's inclination i, i.e., the angle formed by its rotation axis with respect to the line of sight. With this convention, the system is seen pole-on if i = 0◦ and edge-on if i = 90◦. The method used requires the knowledge of the following stellar properties : the colour index (B−V), the radius R(cid:63), the projected rotational velocity, vrot sin(i), and finally R(cid:48) HK is the chromospheric flux in the H and K lines of Ca ii, and T(cid:63) is the star effective temperature. These properties for ζ2 Reticuli are summarized in Table B.1. HK, an activity indicator defined as F(cid:48) We first use the activity/rotation diagram built by Noyes et al. (1984), which plots log(R(cid:48) HK) versus log(Ro) and shows a rela- tionship between these two quantities for late-type stars. Ro = Prot/τc is the Rossby number, Prot being the rotational period of the star and τc a model-dependent, typical, convective time, called the 'turnover time'. Using Fig. 6(b) of Noyes et al. (1984) and the HK) = −4.79 found by Henry et al. (1996) for ζ2 Reticuli allows us to estimate log(Prot/τc) ∼ 0.185± 0.085. observed value of log(R(cid:48) Then, using equation (4) of Noyes et al. (1984) where x is defined with the star colour index (B − V) by x = 1 − (B − V), one HK/σT 4 (cid:63), where F(cid:48) 1.362 − 0.166x + 0.025x2 − 5.323x3 , x > 0 1.362 − 0.14x , x < 0 . (B.1) 17 can estimate τc: (cid:40) log τc = Table B.1. Stellar properties of ζ2 Reticuli. Faramaz et al.: Can eccentric debris disks be long-lived? References. (1) Johnson et al. (1966); (2) this study; (3) Eiroa et al. (2013); (4) Reiners & Schmitt (2003); (5) Henry et al. (1996). Reference Stellar Property (B − V) R(cid:63) (R(cid:12)) L(cid:63) (L(cid:12)) T(cid:63) (K) vrot sin i (km/s) log(R(cid:48) HK) log(Ro) τc (days) Prot (days) vrot (km/s) i(◦) Value 0.60 ∼ 0.965 ± 0.05 0.97 5851 2.7 ± 0.3 −4.79 ± 0.03 ∼ 0.185 ± 0.085 ∼ 14.20 ± 2.75 ∼ 3.42 ± 0.66 ∼ 65.5+24.5−31.5 ∼ 9.10 1 2 3 3 4 5 2 2 2 2 2 ζ2 Reticuli has a colour index (B − V) = 0.60 (Johnson et al. 1966), which gives τc = 9.10 days. The corresponding range of possible rotation periods is Prot ∼ 14.20 ± 2.75 days. Since the equatorial rotation velocity is defined as vrot = 2πR(cid:63)/Prot, knowing the stellar radius R(cid:63) allows us to obtain a range of possible values for vrot. Using T(cid:63) and Lbol and corrections prescribed by Masana et al. (2006), we find that for ζ2 Reticuli, R(cid:63) = 0.965R(cid:12) ± 0.05. The corresponding value of equatorial velocity is ∼ 3.42 ± 0.66 km/s. We confront this with an independent measurement of vrot sin i by Reiners & Schmitt (2003). They find vrot sin(i) = 2.7±0.3 km/s, which combined with the calculated rotational velocity, allows us to estimate the stellar inclination. We find that the inclination can range from 34◦ to 90◦ (see Fig B.1). This is rather consistent with an inclined disk. Fig. B.1. Possible inclination angles (taken from the pole) for ζ2 Reticuli as a function of vrot ranging between ∼ 2.76 and 4.08 km/s, and i is computed using the values of vrot sin(i) = 2.7 ± 0.3 km/s (Reiners & Schmitt 2003). The acceptable zone for the inclination is grey-shaded. However, two angles are required to fully constraint the stellar rotation and disk axis. Therefore a degree of freedom remains and different orientations may lead to the same inclination i. Namely, the range of possible orientations leading to a same inclination is the set of axes describing a solid angle 2π sin i about the line of sight. But the number of axes which leads to the same inclination increases with i, since it follows a sin i distribution. This means that the inclinations in a range [34◦; 90◦] are not equiprobable. The probability for a given inclination to be between i and i + di, provided it is in the range [34◦; 90◦], can thus be written : (B.2) (B.3) dP(i) = (cid:82) 90◦ sin idi 34◦ sin idi , (cid:82) i2 (cid:82) 90◦ sin idi i1 34◦ sin idi 18 and the probability to find inclinations between i1 and i2 in the range [34◦; 90◦] is : P(i ∈ [i1; i2]) = . Applying this to the case of ζ2 Reticuli, among the possible range of [34◦; 90◦], we have ∼ 50% chance that the observed −31.5 . Consequently, the disk is more probably seen inclination is in the range [65.5◦; 90◦]. Thus the system inclination is i = 65.5◦+24.5 almost edge-on, with a pure edge-on configuration not having been ruled out. Faramaz et al.: Can eccentric debris disks be long-lived? Because of the large uncertainties, this constraint does not really allow us to say with absolute confidence whether the disk exhibits an eccentric ring or resonant clumps. However, resonant structures are in general thin structures which tend to be hidden by non-resonant bodies and are difficult to detect, even in case of pole-on observations (Reche et al. 2008). This argument clearly tends to support the interpretation of Eiroa et al. (2010), i.e., an eccentric ring structure with e (cid:38) 0.3 seen edge-on, and extending from ∼ 70 to ∼ 120 AU, is observed. Appendix C: Constraints on ζ2 Ret set by direct imaging VLT/NaCo (Lenzen et al. 2003; Rousset et al. 2003) Ks-band data were retrieved from the ESO archive (ID 086.C-0732(A); PI: Lohne,71574). Two epochs were available in August 2010 and November 2010, the former missing photometric calibration so only the latter could be used to set detection limits on the presence of bound companions. Nevertheless, both data sets were reduced and no companion is detected. The data from November, 11th 2010 were obtained in field stabilized-mode with five manual offsets of the derotator to simulate field rotation, with the S27 camera providing a pixel scale of 27 mas/pixel. Twenty image cubes with a DITxNDIT of 1.5s x 42 were obtained, for a total observing time on target of 21min. The semi-transparent mask C 0.7 sep 10 with a diameter of 0.7(cid:48)(cid:48) and a central transmission of 3.5×10−3 was used. Each individual image was bad pixel-corrected and flat- fielded. Background subtraction was made for each cube using the closest sky images. Recentering of the images was done using a Gaussian fit of the attenuated central star. Data selection was made within each data cube using criteria based on the attenuated central star flux and the encircled energy between 0.4(cid:48)(cid:48) and 0.55(cid:48)(cid:48). The images were then binned every 6s, and derotated into a reference frame where the pupil is stabilized in order to simulate Angular Differential Imaging, ADI (Marois et al. 2006). In this reference frame, the total field rotation provided by the manual offsets plus the natural pupi/field rotation is 17◦. This data cube is then reduced using Principal Components Analysis (Soummer et al. 2012), retaining 4 components out of 105. The noise in the final reduced image was estimated using a sliding 9 pixel wide box to get a preliminary map of detection limits in magnitude. We corrected this map by computing the flux losses due to the PCA reductionThey were estimated by injecting fake planets in the data cube at a 10 − σ level and processing again the data. Last, these detection limits in magnitude were converted into detection limits in masses, using the COND models (Baraffe et al. 2003) or BT-settl models (Allard et al. 2011), assuming an age of 2 Gyr. The 2D-detection limits given the COND evolutionary models is presented in Fig. C.1 1 UJF-Grenoble 1 / CNRS-INSU, Institut de Plan´etologie et d'Astrophysique de Grenoble (IPAG) UMR 5274, Grenoble, F-38041, France 2 LESIA, Observatoire de Paris, 92195, Meudon, France 3 Instituto Nacional de Astrof´ısica, ´Optica y Electr´onica, Luis Enrique Erro 1, Sta. Ma. Tonantzintla, Puebla, Mexico 4 Universidad Aut´onoma de Madrid, Dpto. F´ısica Te´orica, M´odulo 15, Facultad de Ciencias, Campus de Cantoblanco, E-28049 Madrid, Spain 5 European Southern Observatory, Casilla 19001, Santiago 19, Chile 6 Dpt de Astrof´ısica, Centro de Astrobiolog´ıa (INTA-CSIC), ESAC Campus, P.O.Box 78, E-28691, Villanueva de la Canada, Madrid, Spain 7 Aurora Technology B.V., ESA-ESAC, P.O. Box 78, 28691, Villanueva de la Canada, Madrid, Spain 8 Jet Propulsion Laboratory, California Institute of Technology, 4800 Oak Grove Drive, Pasadena, CA 91109, USA 9 NASA Goddard Space Flight Center, Exoplanets and Stellar Astrophysics, Code 667, Greenbelt, MD 20771, USA 10 Rutherford Appleton Laboratory, Chilton OX11 0QX, UK 11 Department of Physics and Astrophysics, Open University, Walton Hall, Milton Keynes MK7 6AA, UK 12 Christian-Albrechts-Universitat zu Kiel, Institut fur Theoretische Physik und Astrophysik, Leibnizstr. 15, 24098 Kiel, Germany 19 Faramaz et al.: Can eccentric debris disks be long-lived? Fig. C.1. Map of the detection limits in Jupiter masses set by the COND evolutionary models. The contours range from 60 to 150 MJup with a step of 10 MJup. 20
0905.3685
1
0905
2009-05-22T13:52:33
Detection of Earth-impacting asteroids with the next generation all-sky surveys
[ "astro-ph.EP" ]
We have performed a simulation of a next generation sky survey's (Pan-STARRS 1) efficiency for detecting Earth-impacting asteroids. The steady-state sky-plane distribution of the impactors long before impact is concentrated towards small solar elongations (Chesley and Spahr, 2004) but we find that there is interesting and potentially exploitable behavior in the sky-plane distribution in the months leading up to impact. The next generation surveys will find most of the dangerous impactors (>140m diameter) during their decade-long survey missions though there is the potential to miss difficult objects with long synodic periods appearing in the direction of the Sun, as well as objects with long orbital periods that spend much of their time far from the Sun and Earth. A space-based platform that can observe close to the Sun may be needed to identify many of the potential impactors that spend much of their time interior to the Earth's orbit. The next generation surveys have a good chance of imaging a bolide like 2008TC3 before it enters the atmosphere but the difficulty will lie in obtaining enough images in advance of impact to allow an accurate pre-impact orbit to be computed.
astro-ph.EP
astro-ph
Detection of Earth-impacting asteroids with the next generation all-sky surveys Peter Veres1, Robert Jedicke2, Richard Wainscoat2, Mikael Granvik2, Steve Chesley3, Shinsuke Abe4, Larry Denneau2, Tommy Grav5, ABSTRACT We have performed a simulation of a next generation sky survey's (Pan- STARRS 1) efficiency for detecting Earth-impacting asteroids. The steady-state sky-plane distribution of the impactors long before impact is concentrated to- wards small solar elongations (Chesley & Spahr 2004) but we find that there is interesting and potentially exploitable behavior in the sky-plane distribution in the months leading up to impact. The next generation surveys will find most of the dangerous impactors (>140 m diameter) during their decade-long survey missions though there is the potential to miss difficult objects with long synodic periods appearing in the direction of the Sun, as well as objects with long orbital periods that spend much of their time far from the Sun and Earth. A space-based platform that can observe close to the Sun may be needed to identify many of the potential impactors that spend much of their time interior to the Earth's orbit. The next generation surveys have a good chance of imaging a bolide like 2008 TC3 before it enters the atmosphere but the difficulty will lie in obtaining enough images in advance of impact to allow an accurate pre-impact orbit to be computed. Subject headings: Pan-STARRS; Asteroids; Near-Earth Objects; Meteors; Im- pact Processes 9 0 0 2 y a M 2 2 . ] P E h p - o r t s a [ 1 v 5 8 6 3 . 5 0 9 0 : v i X r a 1Faculty of Mathematics, Physics and Informatics, Comenius University, Mlynska Dolina, 842 48 Bratislava, Slovakia 2University of Hawaii, Institute for Astronomy, 2680 Woodlawn Drive, Honolulu, HI 96822-1897, USA 3Jet Propulsion Laboratory, California Institute of Technology, Pasadena, CA 91109, USA 4Institute of Astronomy, National Central University, No. 300, Jhongda Rd, Jhongli City, Taoyuan County 320, Taiwan 5Department of Physics and Astronomy, Bloomberg 243, Johns Hopkins University, 3400 N. Charles St., Baltimore, MD, 21218-2686, USA – 2 – 1. Introduction Throughout most of human history it was not understood that the Earth has been battered by large asteroids and comets and that the impacts and subsequent environmental changes have serious consequences for the survival and evolution of life on the planet. But in the past ∼50 years more than 170 impact structures have been identified on the surface of the Earth (Earth-impact database 2008). Were it not for the Earth's protective atmo- sphere, oceans, erosion and plate tectonics, the surface of the Earth would be saturated with impact craters like most other atmosphereless solid bodies in our solar system. While the impact probability is now relatively well understood as a function of the impactor size (e.g. Brown et al. 2002; Harris 2007) this work addresses specific questions related to discover- ing impacting asteroids before they hit the Earth. In particular, we build upon the work of Chesley & Spahr (2004) and determine the sky-plane distribution of impacting asteroids before impact and the effectiveness of the next generation large synoptic sky surveys at identifying impactors. The first surveys to target near-Earth objects (NEO) (Helin & Shoemaker 1979), as- teroids and comets with perihelion < 1.3 AU, provided the first look at their orbit and size distribution and allowed the first determination of the impact rate from NEO statistics (Shoemaker 1983) rather than crater counting on the Moon. These pioneers heightened the awareness of the impact risk and gave rise to the current generation of CCD-based aster- oid and comet surveys such as Spacewatch (Gehrels 1986), LINEAR (Stokes et al. 2000), LONEOS (Koehn & Bowell 1999), NEAT (Pravdo et al. 1999), and the current leader in discovering NEOs, the Catalina Sky Survey (CSS) (Larson et al. 1998). These programs benefitted from the elevated impact risk perception when in 1998 the U.S. Congress followed the recommendations of Morrison (1992) and mandated that the U.S. National Aeronautics and Space Administration (NASA) search, find and catalog ≥ 90% of NEOs with diameters larger than 1 km within 10 years. That goal will probably be achieved within the next few years. The residual impact risk is mainly due to the remaining undiscovered large aster- oids and comets (Harris 2007) but Stokes et al. (2004) suggest that the search should be expanded to identify ≥ 90% of potentially hazardous objects (PHO) by 2020. 1 Stokes et al. (2004) showed that the extended goal cannot be achieved in a reasonable time frame with existing survey technology. The search needs to be done from space (rapid completion but at high risk and high cost) or from new ground-based facilities (slower com- pletion but lower risk and lower cost). Their recommendation dovetailed nicely with the 1A PHO is an object with absolute magnitude H ≤ 22 (∼ 140 m diameter) on an orbit that comes within 0.05 AU of the Earth's orbit. – 3 – Astronomy and Astrophysics Survey Committee (2001) Decadal Report that made a strong case for the development of a large synoptic survey telescope (LSST) that would provide the necessary depth and sky coverage to identify the smaller PHOs while also satisfying the goals of other fields of astronomy. There are currently a few candidates for a large synoptic survey telescope. The most ambitious is an 8.4 m system being designed by the eponymous LSSTC (the LSST Corpora- tion) that anticipates beginning survey operations in 2016 in Chile. With a ∼9 deg2 field of view and 15 s exposures, simulations suggest that their system could identify &90% of PHOs in 15 years (Ivezi´c et al. 2007). A more modest LSST known as the Panoramic Survey Tele- scope and Rapid Response System (Pan-STARRS) will be composed of four 1.8m telescopes (PS4) and is expected to be located atop Mauna Kea in Hawaii. A prototype single telescope for Pan-STARRS known as PS1 should begin operations in mid-2009 from Haleakala, Maui. With a ∼7 deg2 field of view, the excellent seeing from the summit of Mauna Kea, and the use of orthogonal transfer array CCDs (Burke et al. 2007) for on-chip image motion com- pensation, the Pan-STARRS system will be competitive with and completed earlier than the LSSTC's system. The next generation survey telescopes have the potential to be prolific discoverers of PHOs but Earthlings aren't so much concerned with statistical impact risk calculated from PHO orbital distributions as they are interested in whether an impact event will occur. The statistical risk of a house burning down may seem inconsequential until you consider the actuality of your house being incinerated. Similarly, while Harris (2008) has calculated that expected fatalities due to an unanticipated asteroid impact have dropped from ∼1,100/year before the onset of modern NEO surveys to only ∼80/year now, as a species we would like to know whether one of the fatality inducing impacts will take place this century. Thus, this work concentrates on the detection of objects that may impact the Earth in the next hundred years. Following Chesley & Spahr (2004) we concentrate on the subset of PHOs that are in fact destined for a collision with the Earth. They showed that long before impact the im- pactors' steady state sky-plane distribution is concentrated on the ecliptic and at small solar elongation. We extend their analysis and find that the sky-plane distribution of impactors has interesting and potentially useful structure in the time leading to collision. We also study the capabilities of one next-generation survey (PS1) at identifying the impactors well before collision. In particular, we will answer the following questions: How different are the orbital characteristics of the impactor population and current NEO and PHO models? What is the survey efficiency for identifying asteroids on a collision course with the Earth as a function of their diameter? How much warning time will be provided before the impact? – 4 – How accurate is the orbital solution prior to impact? How does the MOID2 evolve in time and is the current definition of a PHO consistent with flagging dangerous objects? What are the orbital properties of objects that are not found? Are there methods to improve the effi- ciency of identifying impactors? Given the size-frequency distribution of NEOs what is the probability that PS1 will actually identify an impactor and what will be its most probable size? 2. Synthetic Earth-impacting asteroids Our synthetic impactor population model is described in detail in Chesley & Spahr (2004) and Grav et al. (2009). Here we provide a brief summary of the technique. We created ∼130,000 impactors based on the NEO population developed by Bottke et al. (2000) and Bottke et al. (2002) hereafter referred to as the Bottke NEO Model. The model incorporates objects from both asteroidal and cometary source regions but has at least two problems that affect its utility for creating an impactor population: 1) it assumes that the orbit distribution of NEOs is independent of their diameter and 2) it provides the (a, e, i, H) (semi-major axis, eccentricity, inclination, and absolute magnitude) distribution for NEOs on a coarse grid that is not suited to the narrower range of orbital elements of the impacting asteroids. However, there are few options to use as starting points for developing an impactor population and we will compare our impactor population's orbit distribution to the known small impactor population to understand the limitations of our technique. To generate the impactors NEOs were randomly selected from the Bottke NEO model and assigned random longitudes of ascending node and arguments of perihelion. Orbits with a MOID small enough to permit an impact were saved as potential impactors and then filtered according to their likelihood of impact to obtain the final set of impactors. The likelihood is the fraction of time that an object spends in close proximity to the Earth's orbit. i.e. orbits with a small velocity relative to the Earth tend to have shorter impact intervals and higher intrinsic impact probabilities. Higher likelihoods received higher weighting in the selection. If an orbit was chosen as an impactor then a year of impact was randomly selected between 2010 and 2110 - the date of collision is already randomly fixed by the longitude of the node at impact. To this point, the process assumed a two-body asteroid orbit with no planetary perturbations. The final step was to ensure an impact under the influence of all the perturbations in a complete solar system dynamical model. This was done by differentially adjusting the two-body argument of perihelion (ω) and orbital anomaly to reach a randomly 2Minimum Orbital Intersection Distance – 5 – selected target plane coordinate on the figure of the Earth. The final result is an osculating element set that leads to an Earth impact when propagated with the full dynamical model. The full set of impactors generate about three impacts per day uniformly distributed over the globe with an average separation of about 70 km.3 This technique preferentially selects objects on Earth-like orbits out of the Bottke NEO model but Brasser & Wiegert (2008) show that objects do not remain long in these types of orbits. This is not a problem except in the sense addressed above - that the Bottke NEO model is provided on a relatively coarse grid - because the NEO model already accounts for NEO 'residence times' on all types of NEO orbits. However, since we assume a flat distribution of NEO orbit elements within the (a, e, i) bin corresponding to Earth-like orbits it is likely that we generate fractionally more of the extremely Earth-like orbits than exist in reality. As shown by Chesley & Spahr (2004) and in Fig. 1 there are important differences between the impactor population and the NEOs. The impactors have orbits with lower semi-major axis, inclination and eccentricity. This has the effect of decreasing the Earth encounter and impact velocity (v∞ and vimp respectively) for the impactors relative to the NEOs. The decreased impact velocity has implications for modelling the impact crater size- frequency distribution on the Earth and Moon since lower impact energies per unit mass require larger impactors to create equivalent-size craters. As mentioned above, the Bottke NEO model assumed that the orbit distribution of the NEOs is independent of size. While this assumption is probably fine for large objects it must break down at smaller sizes due to the effect of non-gravitational forces such as the Yarkovsky effect (e.g. Bottke et al. 1998; O'Brien & Greenberg 2005). Where the transition occurs is not clear but Fig. 1 compares the (a, e, i) distribution of our impactor population to sporadic fireballs from the IAU Meteor Database of photographic orbits (Lindblad et al. 2003). We removed fireballs due to 17 major meteor showers4 (Jenniskens 2006) by requiring that the meteors not be associated with the parent meteor body using the dimensionless orbit 3Gallant et al. (2006) use a superior (but much more time consuming) technique to generate an even more unbiased impactor population from the Bottke NEO model and confirm that the latitude and longitude distribution of impact locations is flat to within a few percent when averaged over all impactors and times of year. 4From the IAU Meteor Database: Quadrantids, Lyrids, Pi Puppids, Eta Aquarids, Arietids, Daytime Zeta Perseids, June Bootids, Southern Delta Aquarids, Perseids, Draconids, Orionids, Southern Taurids, Northern Taurids, Leonids, Puppid/Velids, Geminids, Ursids. similarity D-criterion of Valsecchi et al. (1999)5: – 6 – D2 = [U2 − U1]2 + w1[cos θ2 − cos θ1]2 + ∆ξ 2, (1) where cos θ = 1 − U 2 − 1/a 2U , ∆ξ 2 = min[ w2∆φ2 A + w3∆λ2 ⊕A, w2∆φ2 B + w3∆λ2 ⊕B ], 2 φA = 2 sin(cid:16) φ2 − φ1 (cid:17), φB = 2 sin(cid:16) π + φ2 − φ1 (cid:17), λ⊕A = 2 sin(cid:16) λ⊕2 − λ⊕1 (cid:17), λ⊕B = 2 sin(cid:16) π + λ⊕2 − λ⊕1 2 2 2 (cid:17). The 1 and 2 subscripts refer to the two bodies whose orbits are being compared, U is the unperturbed geocentric speed just prior to impact, (θ, φ) define the direction of the radiant in a frame moving with the Earth about the Sun, and λ⊕ is the ecliptic longitude of the Earth at the time of meteoroid impact. The weighting factors wi were set to 1 as per Valsecchi et al. (1999). All objects with D-criterion relative to a parent body of ≤ 0.2 were discarded (577 meteors) leaving 2002 sporadic background fireballs. The impactor and NEO orbit distributions have already been discussed briefly above and in detail by Chesley & Spahr (2004). The differences between the impactor population and the fireballs are perhaps more interesting where it is important to keep in mind that the comparison in Fig. 1 is between the bias corrected impactor population and the ob- served bolide population. We believe that the apparent difference between the impactor and bolide distributions is a consequence of the uncorrected observational selection effects in the bolide data along with modifications in the distributions due to the Yarkovsky effect (e.g. Farinella et al. 1998). The bolide detection technique has a strong bias towards high kinetic energy events that preferentially detects objects on cometary-like orbits (Ceplecha et al. 1998). Indeed, if we consider 'comets' to be objects from our sporadic bolide data with a > 4 AU or e > 0.9 or i > 90◦ then ∼13.4% of the objects are of 'cometary' origin. This value is about twice the cometary fraction of 6 ± 4% suggested by the Bottke NEO model and consistent with the expected 'comet' enhancement in the bolide data. On the other hand, it is only half the 25% cometary contribution suggested by Stuart & Binzel (2004). 5We obtain essentially identical results using other D-criterion formulations (Southworth & Hawkins 1963; Drummond 1981; Galligan 2001) with corresponding but different upper limits on the D-criterion value. – 7 – No meteor or fireball had ever been detected before entering the Earth's atmosphere before we generated our synthetic population of Earth-impacting asteroids6. Then, on 7 Oc- tober 2008, asteroid 2008 TC3 was discovered by CSS using their 1.5 m telescope. Rapid fol- lowup by other observatories made it almost immediately clear that the few-meter-diameter (H = 30.7) asteroid would enter the Earth's atmosphere within a day and explode over northern Sudan. The substantial and largely self-organized follow-up effort resulted in 789 astrometric observations from amateur and professional observatories worldwide being sub- mitted to the Minor Planet Center. Due to the parallax induced by the distance between the observatories and the proximity of the asteroid, the observations allowed an accurate pre-impact orbit to be computed despite the short observational timespan (Table 1). The pre-impact orbit is very consistent with our predicted impactor orbit distribution as shown in Fig. 1. Note that, in particular for the eccentricity, the pre-impact orbit matches the expected impactor population better than the debiased NEO population or the observed bolide population. Orbital elements and their uncertainties for 2008 TC3 a [AU] 1.2712175 0.0000031 ω [◦] 194.1308964 233.954719 0.000040 0.0000011 e 0.2856863 0.0000023 i [◦] 2.331633 0.000017 Ω [◦] M0 [◦] 328.58963 0.00015 Elements 1-σ unc Table 1: Keplerian elements and their 1-σ uncertainties for 2008 TC3 for the epoch 2008 Oct 6.11535 TT, about one day before atmospheric entry on 2008 Oct 7.1146 UTC. The orbital solution made use of 574 astrometric observations (42 of which were discarded as outliers) and assumed an uncorrelated astrometric uncertainty of 0.5 arcseconds for every measurement. The orbital solution was obtained using the OpenOrb software (Granvik et al. 2008). 2.1. Time evolution of the impactor population's MOIDs Recall that a PHO is defined as an object with H ≤ 22 and a MOID < 0.05 AU with respect to the Earth's orbit. We investigated the evolution of the MOID for our impactor population as a function of time before impact as shown in Fig. 2. All the impactors's osculating orbits were integrated using the JPL's N-body integrator incorporating the effects of the Sun, eight planets, and the dwarf planets Pluto, Ceres, Vesta and Pallas. MOIDs were then calculated at different times before impact using the synthetic objects's integrated 6There were suspected radar detections of exoatmospheric meteoroids in the late 1970's (Kessler et al. 1980) – 8 – osculating orbits at the time of interest, not the orbit derived from synthetic observations of the object. We see that as the time of impact approaches the MOID decreases such that one month before impact essentially all objects have a MOID less than the Earth's capture radius (b): b = R ·s1 + v2 e v2 ∞ , (2) where ve represents the escape velocity from the surface of the Earth. Fig. 3 shows that ∼99% of all impactors are identified as PHOs 45 years in advance of impact. Even 100 years before impact ∼98% of the objects have a MOID<0.05 AU though the fraction of non-PHO impactors is increasing rapidly as the time before impact increases. There are a few objects with MOID>0.05 AU that would not be identified as impactors or PHOs using a single MOID determination from a derived osculating orbit at the time of discovery. These objects suffer from a close approach to Jupiter that converts them from a harmless object into an Earth-impactor. Modern impact monitoring sites such as JPL (Milani et al. 2005) and NEODyS (Chesley & Milani 1999) integrate the orbits of all objects to identify these unusual but dangerous cases. At all eight times before impact the cumulative fractional distribution of MOIDs (fC) in Figs. 2 exhibit a nearly constant slope for fC . 0.9. Under the simple assumption that the impactors are randomly distributed on the impact plane at the Earth we would expect fC ∝ MOID2. Thus, we fit the distribution to log fC = log f ′ C + m log[MOID/(10−4AU)] C is (roughly) the cumulative fraction of objects with MOID < 10−4 AU. We find where f ′ an average slope over all impact times of m = 0.88 ± 0.01, much less than 2 and quite different from the result of Tancredi (1998) who found a slope of ∼1.23 for clones of the extremely Earth orbit-like 1991 VG. We believe that the difference between the expected quadratic and our measured unit slope is due to the Earth's gravitational focussing. The cumulative fraction of impactors with MOID< 10−4 AU as a function of time is shown in Fig 4a. Clearly, the interpration of fC as a cumulative fraction breaks down for fC & 1 but we see that essentially all the impactors have MOID< 10−4 AU about 73 days before impact. Fig 4b provides a quantitative assessment of the time evolution of the mode of the MOID distribution from Fig. 2. The combination of the fits to the data in Figures 4 allow a rough determination of the time evolution of the fraction of impactors with a given MOID. 2.2. Sky-plane distribution of Earth-impacting asteroids In this section we analyze the sky-plane distribution of Earth-impacting asteroids in the years, months, and days leading up to impact. The topocentric location and brightness – 9 – of the impactors (as seen from Mauna Kea in all cases) was calculated for a ∼10K subset of impactors for every day in the 100 years leading to impact using the OpenOrb software package (Granvik et al. 2008). The steady-state7 sky plane distribution of Earth-impactors in Figure 5a shows all ob- jects that would be visible to PS1 with V < 22.7 assuming that H = 20 (diameter ∼350 m). The figure clearly shows the high sky-plane density 'sweet spots' at small solar elongations (± ∼ 90◦) identified by Chesley & Spahr (2004). The objects in the sweet spots are close to the Earth and therefore relatively bright despite their large phase angles, but are too close to the Sun to be easily observed. Twenty years before impact 962 objects (∼9.6%) lie within the sweet spot region that encompasses topocentric solar elongations from 60◦ to 90◦. Of these, 552 (∼5.5%) have an ecliptic latitude (β) in the range −10◦ ≤ β ≤ +10◦ and 826 (∼8.3%) have β < 20◦. The dis- tribution in ecliptic latitude (Fig. 6) is distinctly broader than provided by Chesley & Spahr (2004) who showed cumulative detections made by a simulated version of the LINEAR sur- vey (Stokes et al. 2000). The broad distribution in ecliptic latitude suggests that there is merit in extending searches for PHOs in the sweet spots to ecliptic latitudes of ±20◦. Since the southern parts of these extended sweet spots are close to the horizon from the northern hemisphere, and the northern parts of the extended sweet spots are close to the horizon from the southern hemisphere, this in turn suggests that there is merit in conducting searches for PHOs from both hemispheres or from space (e.g. Hildebrand et al. 2007). The middle and bottom distributions in Fig. 5 show the location of the same sample of objects 20 years before impact but with brighter magnitude limits of V < 20.7 and V < 18.7. Another way to interpret these figures is that they show the detectability (for the same limiting magnitude of V < 22.7) of H = 22 (diameter 140 m) and H = 24 (55 m) impactors at this particular time. The smaller objects must be closer to the Earth in order to be above the system's limiting magnitude but this also has the effect of increasing the phase angle which further decreases the apparent magnitude. Thus, smaller objects must be closer and have smaller phase angle to be detected. While searches for larger objects are efficient in the sweet spots they should be supplemented by opposition surveys to find smaller objects. e.g. 20 years before impact 2008 TC3 had an apparent magnitude of V ≈ 31 with an ecliptic opposition-centered longitude of λopp ≈ −92◦ and latitude of β ≈ 2◦ - not detectable but in the sweet spot region. Figures 7a-d show the development of the sky-plane distribution of impactors as a 7Although Figure 5a shows the sky-plane distribution of Earth-impacting asteroids 20 years before impact we have verified that the distribution is the same for earlier pre-impact times. – 10 – function of time before impact. • 1 day before impact the impactors are located in two main concentrations; one centered on the opposition direction and the other centered on the Sun. The impactors are widely spread in ecliptic latitude because they are close to the Earth. About half as many impactors approach from the λopp > 0 (a.m./morning) side compared to the λopp < 0 (p.m./evening) side. Objects that approach from the morning side are moving more slowly around the Sun than the Earth, meaning that the Earth runs into them. They have perihelia that are well inside the Earth's orbit, or have higher inclinations. Objects that impact on the evening side catch up with the Earth in its orbit around the Sun. They have perihelia that are closer to the Earth's orbit, or have higher eccentricities. • 30 days before impact, the concentration that will impact from the opposition direction has moved east and is centered close to λopp = 20◦. The impactors that will approach from close to the Sun direction are also further east with some of them becoming observable in the evening sweet spot. • 60 days before impact many of the impactors that will approach the Earth from outside the Earth's orbit are scattered around λopp = 40◦ and more of the impactors that will approach from inside the Earth's orbit are becoming visible near the evening sweet spot. This trend continues 90 days and 120 days before impact with the outside impactors moving further east and more of the inside impactors becoming more visible in the evening sweet spot. • More than 180 days before impact the sky plane distribution of the impactors is more complex. The observability of impactors decreases significantly because of their small solar elongations. 180 days, 1 year and 1.5 years before impact many impactors are close to the direction of the Sun or are far from the Earth. Observability improves from 2 years onwards as the distribution more closely mimics the steady state distribution of impactors in the sky. The difficulty in observing many of the impactors in the period between 0.5 and 1.5 years before collision with the Earth is important. In the event that an impact is predicted, precision astrometry will be needed during this period to determine where and when the impact will occur so that life-saving actions can be taken if necessary. – 11 – 3. Next generation sky survey impactor detection performance Our goals in this section are to determine • the efficiency of a next-generation sky survey at identifying Earth-impacting asteroids • the likely warning time for impending impacts • the reason(s) why some impactors remain undetected • methods for discovering the impactors that are most difficult to detect • the probability that a next-generation survey will identify an impacting object as a function of its diameter There have been numerous efforts in the recent past to simulate the performance of indi- vidual ground and/or space-based sky surveys at detecting NEOs (e.g. Bowell & Muinonen 1994; Jedicke et al. 2003; Harris & Bowell, 2004; Stokes et al. 2004; Ivezi´c et al. 2007; Moon et al. 2008) but only Chesley & Spahr (2004) concentrated specifically on identifying impactors and their work also simulated the performance of the last generation of asteroid surveys. As discussed above, the impactors have a different orbit distribution from the NEO or even PHO population and pose different challenges to a survey and its moving object detection system. These include some pathological cases in the distribution of synodic peri- ods for the objects, the fact that they are infrequently close enough and above the limiting magnitude of the detection system, and that at small topocentric distance their apparent rate of motion on the sky can be very small (mimicking more distant objects) and rapidly varying due to the topocentric motion of the observer. For all these reasons, when simulating the detection of impactors it is probably not sufficient to assume that all the objects will be detected and identified as imminently hazardous. A high-fidelity simulation is required that models the entire detection process from imaging through orbit determination and hazard assessment. The Pan-STARRS surveying mode has not yet been decided upon and, realistically, will asymptotically approach its final configuration during the first year of operation. On average, each month PS1 will survey π/2 steradians or ∼5,000 square degrees in an 'opposition' survey with an overlap between months of about π/4 steradians. The opposition fields are imaged in g, r, and i each lunation near new moon. The redder (z, y) filter observations are obtained near quadrature closer to full moon to take advantage of their reduced sensitivity to scattered moonlight. Each field will be imaged twice (15-30 minutes apart) in each filter in each month and imaged in two lunations/year due to the overlapping survey area from month to month. – 12 – We employed the Pan-STARRS Moving Object Processing System (MOPS) to process synthetic detections generated in a pseudo-realistic Pan-STARRS survey. The simulated survey covers a large region (3600 deg2) of the sky around opposition (within roughly ±30◦ from opposition in longitude and ±30◦ from the ecliptic) and two ∼ 600 deg2 regions in the sweet spots defined by Chesley & Spahr (2004) within ±10◦ of the ecliptic and from about 60◦ to 90◦ solar elongation. It is essentially a solar system specific sub-set of the sky that the Pan-STARRS system is expected to cover when it becomes operational. The survey simulator is described in detail in Jedicke et al. (2005) and Denneau et al. (2007). It incorporates a crude weather simulation and a realistic survey pattern that at- tempts to observe fields at high altitude and on the meridian. We use a full N-body ephemeris determination to calculate the exact (RA,Dec) of each impactor in each synthetic field and then add noise to the astrometric position according to the expected PS1 S/N-dependent astrometric error model. The photometry for each object is similarly degraded and then we make a cut at S/N=5 in order to simulate the statistical loss of detections near the system's limiting magnitude (V = 22.7). Each field is observed twice each night within ∼15 minutes to allow the formation of 'tracklets' - pairs of detections at nearly the same spatial location that might represent the same solar system object. Fields are re-observed 3 times per luna- tion (weather permitting) and tracklets are linked across nights to form 'tracks' that are then tested for consistency using an initial orbit determination (IOD). Detections in tracks with small astrometric residuals in the IOD are subsequently differentially corrected to obtain a final orbit. The major item that is not simulated is the effect of the camera fill-factor and this could have a major impact on the survey's efficiency. We will estimate this effect below but the simulation is otherwise one of the highest fidelity moving object simulations ever attempted. We anticipate that there may be concerns regarding simulating the performance of detector systems especially in the sense that the simulations tend to be optimistic compared to the actual detector. In particular, in this case we have 1. used a realistic survey scenario but it is not the survey that will eventually be imple- mented by PS1, 2. used a simplistic weather model, 3. used a S/N-dependent astrometric error that is more appropriate to future surveys with better catalogs (at the onset of operations PS1 will probably use the USNO-B catalog (Monet et al. 2003) which will limit absolute astrometry to worse than 0.1′′), 4. not incorporated false detections, – 13 – 5. used an early version of the MOPS with reduced efficiency compared to the most recent version (∼80% vs. ∼100%), 6. not accounted for the camera fill-factor (the fraction of 'live' pixels on the detector compared to the footprint of the focal plane on the sky). We think that the first three factors are relatively unimportant. The implemented survey scenario is quite good and surveys most of the sky in which solar system objects will appear. The weather model is simple but has the desired effect of disrupting the cadence of observations and sometimes eliminating lunations from consideration due to there being too many bad nights. We have also tested MOPS performance under conditions of larger astrometric uncertainty and find that it still performs well. The fact that no false detections were used in the simulation is also unimportant. In targetted smaller simulations (e.g. Milani et al. 2008) we tested the MOPS using a full density complement of false and synthetic asteroid detections and found no degradation in performance at the 5-σ level. In those tests the minimum ratio of false:synthetic detections was 1:1 on the ecliptic where the density of synthetic asteroid detections is highest. Since the density of asteroids decreases rapidly with latitude the false:synthetic ratio increases dra- matically toward the ecliptic poles. Still, few tracklets incorporate false detections because those detections are spatially uncorrelated (in the simulation). It is likely that real survey systems will produce correlated false detections near chip gaps, due to CCD defects, diffrac- tion spikes, etc., and this will produce more false tracklets than our simulation. However, the false tracklets will not link to other tracklets on other days and, if they do, will not pass quality control checks on the derived orbit. Furthermore, MOPS makes use of the detection's morphology when combining detections into tracklets - the distance between the detections must be consistent with trailing observed in the detections themselves. Indeed, we have employed MOPS to identify asteroids in real data from the CFHT telescope (Masiero et al. 2009) and from Spacewatch (e.g. Larsen et al. 2007) with excellent performance. The last two factors are most important. The reduced efficiency of the MOPS software used in this simulation will have a proportional effect on the number of simulated impactor discoveries - about a 20% reduction. The actual system performance will therefore be about 25% better than reported here. On the other hand, the largest negative impact on the simulation will be the PS1 camera fill-factor. The effective camera fill-factor (f ) is reduced by the metal lines between individual OTA cells, the gaps between the CCDs, the use of some cells for guide star acquisition, dead cells, bad cells (e.g. due to charge transfer efficiency problems or dark noise) and the removal of portions of the image by the PS1 funding agency to excise fast moving satellites. – 14 – We expect the overall fill-factor due to all these effects to result in f ∼ 0.88. As described above, for solar system discoveries we require 6 detections on 3 nights within a lunation. Thus, the impact of the fill-factor on the detection efficiency (ǫ) for fast moving objects like the impactor population is simply ǫ = f 6 ∼ 0.46 if the inactive area is uncorrelated on the focal plane or ǫ = f 3 ∼ 0.68 if it is correlated. That is, if the first detection of an object appears on an inactive area then its second detection is also likely to appear on an inactive area in the 'correlated' case and unlikely to do so in the 'uncorrelated' case. Losing 1 3 to 1 of potential discoveries due to the fill-factor is unfortunate but the loss is mitigated for slow moving distant objects because they appear near opposition in successive lunations. PS1 has two opportunities to discover the object in each of two successive lunations so the efficiency for finding these objects will be & 90%. The problem is worse for objects like NEOs and impactors that spend fewer lunations above the detection threshold. 2 To determine the Pan-STARRS (PS1) survey efficiency for detecting impacting asteroids as a function of time and size we divided our sample of ∼130,000 synthetic impactors into six independent sets that were assigned different absolute magnitudes (H) as shown in Table 2. We can assign any H to the impactors because the Bottke NEO model assumed that the size distribution of the NEOs was independent of the orbital elements. The six selected sizes span almost two orders of magnitude in diameter. The number of objects of each size was selected to provide a good number of derived objects (after processing through MOPS) and yet not so many objects as to be prohibitively expensive in processing time. Each of these six independent sets of impactors were then run through a simulation of the PS1+MOPS system. Figure 8 shows that the efficiency of the PS1 survey at detecting impactors increases as a function of time and impactor diameter. In just four years of PS1 operations it could identify ∼85% of all 1 km diameter objects that will impact the Earth in the next 100 years. Interestingly, it has a ∼74% chance of identifying a 1 km diameter that would impact during the 4 year time span of the survey. At first glance this is surprising since PS1 can detect a 1 km diameter asteroid at opposition at a geocentric distance of ∼2.67 AU. The objects should be visible long before impact in the large volume of space surveyed by the PS1 system. We will explore the reasons for the ineffectiveness of the survey at detecting these large hazardous objects below. The impactor discovery efficiency increases nearly linearly with time for objects of 200 m diameter to ∼40% efficiency in just four years. While this behavior cannot increase indef- initely we estimate that after 12 years the efficiency may be ∼80%. This is an interesting – 15 – value because the Pan-STARRS PS4 system will have 4× the collecting area8 of the PS1 system modelled here and thus the efficiency curve for 200 m diameter objects for PS1 corre- sponds to the 100 m diameter efficiency curve for PS4. Thus, during its anticipated 10 year survey mission PS4 alone may reach nearly 90% completion for objects ≥140 m diameter or larger that will impact in the next 100 years. Table 2: Size frequency distribution of synthetic impactors used in the PS1 survey simulation. The conversion between diameter and absolute magnitude assumes an albedo of p = 0.14 for NEOs (Stuart & Binzel 2004). absolute diameter (meters) magnitude number 1000 500 200 100 50 20 Total 17.75 19.25 21.25 22.75 24.25 26.25 1193 3816 4804 10001 25001 85000 129815 The impact warning time was defined by Chodas & Giorgini (2008) as the time between impact and when the probability of an impact on the Earth is calculated to be more than 50%. However, we believe that the experience and response to the discovery of (99942) Apophis shows that the impact awareness time for an object is considerably longer than the impact warning time. Once an object is discovered that has a non-zero probability of impact it is observed obsessively at every opportunity and impact calculations are regularly refined to monitor the impact probability. We define the impact awareness time as the period between the identification of an impactor as a PHO (MOID < 0.05 AU) and its time of impact. As discussed above, almost all the impactors in this study have MOID ≪ 0.05 AU long before impact. Fig. 9 (top) shows the MOID for derived objects at discovery (when at least 3 nights of observations have been obtained in the course of a single lunation). Almost all the objects will immediately be flagged as PHOs and therefore start the impact awareness time clock. The error on the MOIDs, the difference between the MOID for the synthetic object 8Since the PS4 system employs 4 separate cameras viewing the same field the PS4 image stacks will have ∼100% fill-factor. – 16 – and the MOID for the derived object, are small after just a single lunation with typical ∆MOID∼MOID∼ 10−3 AU. As more observations are acquired for these objects the impact probability should monotonically increase to 100%. Figure 10 shows the impact awareness time as a function of the impactor size. Since the synthetic impactors were designed to hit the Earth at random times over the next 100 years, and large objects can be detected long before impact, it is no surprise that the impact awareness time for large discovered synthetic objects is evenly distributed over 100 years (Figure 10 top). The apparent decrease in the fraction of 1000 m objects with larger impact awareness times is an artifact of statistics - fitting a line to the distribution shows that it is consistent with being flat at the 1-σ level. Surprisingly, the impact awareness times for small discovered synthetic objects is also evenly distributed over 100 years (Figure 10 bottom). This is because when PS1 discovers an object and observes it on ≥3 nights over an ∼ 10 day period the orbit is good enough to accurately determine the MOID and determine the impact awareness time. Of course, it may be difficult to obtain followup observations of the smallest objects to refine the orbit and the impact probability calculation. It is important to keep in mind that Fig. 10 does not include the large spike at zero warning time due to the undiscovered objects. For example, Figure 8 shows that ∼60% of 200 m diameter objects remain undiscovered at the end of the four years of PS1 surveying. Thus, the most probable awareness (warning) time for the smaller objects is zero. But when they are discovered before the apparition in which they impact the warning time can be many decades. As discussed above, to be 90% effective at eliminating the risk of an unanticipated impact in the next 100 years for objects >140 m diameter requires a PS1-like system to survey much longer or a more powerful system like PS4 (Kaiser & the Pan-STARRS Team 2005) or the system being developed by the LSSTC (Ivezi´c et al. 2007). While it is interesting to compute the efficiency with which a next generation survey such as PS1 can find impacting asteroids the reality of the situation is that it is unlikely that any large objects are on a collision course with the Earth in the next 100 years. The expected number of detected impactors, N(D), is simply the product of the efficiency, ǫ(D), for detecting impactors of diameter D and the size-frequency distribution of the population, n(D). The efficiency was already shown in Fig. 8 for D > 20 m. We use Brown et al. (2002)'s determination of the annual cumulative number of objects striking the Earth's atmosphere: N(> D) = 37(D/meters)−2.7 which implies that objects in the size range of 2008 TC3 strike the Earth about twice per year. Fig. 11 shows that unless the Earth is extremely unlucky PS1 will not detect a large (D > 20 m) impacting asteroid. A best case linear extrapolation of ǫ(D) to D < 20 m using just the 20 m and 50 m diameter points suggests that the likelihood of obtaining 3 nights of detections in the discovery of even smaller impactors is extremely – 17 – unlikely. The efficiency is decreasing faster with smaller diameter than the number of objects is increasing. While it is unlikely that PS1 will single-handedly obtain enough detections to determine a pre-impact orbit, the possibility of detecting a small asteroid prior to impact (such as the impactor 2008 TC3), or precovering the asteroid after impact using an orbit derived from the bolide trajectory, is also interesting to meteoriticists. With sufficient advance notice it would be possible to organize a ground-based team to observe the asteroid's atmospheric entry, train space-based platforms on the impact location, and prepare for recovery of meteorites. Post- impact precovery of the object could be useful to help determine the pre-impact size of the object and to verify the pre-impact orbit determination for the bolides (e.g. Abe 2008). Simulating the automated detection by PS1+MOPS of smaller impactors is difficult. At the current time MOPS operates by linking together tracklets on 3 nights taken over the course of ∼ 10 days. With typical impact velocities of ∼ 15 km/s the best case scenario requires that PS1+MOPS first detect the object at a distance of ∼ 13 × 106 km (approx- imately 10 days before impact). At PS1's assumed limiting magnitude of V = 22.7 this requires that the object be > 3m in diameter. Even if this situation were to unfold, varia- tions in the object's apparent position and velocity on the sky due to the topocentric motion of the observatory would likely render the object difficult to link within the MOPS (though we have not yet studied this scenario in detail). Instead, we consider the possibility that PS1 will detect an impacting asteroid prior to impact but on too few nights to determine a pre-impact orbit. It is unlikely that the smallest objects (< 20 m diameter) will be discovered, and have good enough derived orbits to guarantee impact, in any apparition except for the one in which they strike the Earth. Thus, we determined the sky-plane distribution and rate of motion 1, 5 and 10 days before impact for all the objects in Table 2. Once again, since the Bottke NEO model's orbit distribution is independent of size, and assuming that we can extend the orbit distribution of the large objects down to 1 m diameter (we have already discussed that while there is reason to believe that the Yarkovsky effect will modify the orbit distribution there is as of yet no debiased orbit distribution for the small impactors), we can assign any size or H to the objects and quickly determine the ensemble's apparent magnitude distribution. The efficiency for detecting these objects is then simply the ratio between the number that meet all our search criteria (within the surveying region, V < 22.7, assuming that surveying takes place on only 75% of nights due to weather, and with a rate of motion within the detectability range) and the total number. Fig. 12 shows that the maximum efficiency for the smallest objects is about 0.6%. – 18 – The distributions of the rates of motion of the meteoroids before impact are shown in Fig. 13. Typical rates of motion are much slower on the final approach trajectory than on fly-by apparitions. The detectable rates of motion are limited to 0 ≤ ω ≤ 12 deg/day where the upper limit is enforced by the Pan-STARRS funding agency. The figure shows clearly that the restricted rate of motion does not have a strong effect on detection efficiency when the impactors are on their approach to the Earth. In the discussion above and elsewhere in this work we have ignored the effects of trailing loss - when objects move during the exposure time they leave 'trails' on the image rather than point sources and the trails have, in the past, been more difficult to identify than point sources. While the PS1 point source detection limit is expected to be V ∼ 22.7 the efficiency of our faint trail detection algorithm is not yet known. It is thought that it should be efficient to per-pixel S/N ≪ 1 implying that the detection efficiency could be quite good even for intrinsically faint objects. Since we have not yet measured the trailing losses we ignore them. This means that the small impactor detection efficiencies reported here and in Fig. 12 are upper limits. Figure 11 shows the expected number of detections of small impactors (bolides) within 10 days of impact during the 4 year PS1 survey as a function of diameter using the Brown et al. (2002) size-frequency distribution (SFD). The cumulative probability of PS1 detecting a bolide in the 1-10 m size range is 25 ± 15% (the error is determined from the SFD only). In other words, PS1 has ∼ 25% probability of detecting another 2008 TC3-like object during its four year survey mission. Considering that we have argued that PS1 is the first of the next generation sky-surveys and that this powerful system only has a ∼ 25% probability of detecting another 2008 TC3- like object in four surveying years, how could a 'last-generation' survey like CSS have dis- covered 2008 TC3? In addition to the impactor, the CSS has also discovered several other small asteroids making very close approaches to the Earth (e.g. 2008 UA202). The CSS's telescope system has (Ed Beshore, personal communication) a smaller etendue (1.5 m × 1 deg2) than the PS1 system (1.8 m × 7 deg2) at a significantly worse site (CSS FWHM ∼ 2′′ vs ∼ 1′′ for PS1; worse weather conditions; brighter sky), and twice the camera readout time (∼ 12 s vs. ∼ 6s). Thus, we conservatively estimate that the PS1 system will be ∼ 10× more efficient than the CSS. The CSS survey strategy sacrifices limiting magnitude to cover as much sky as possible (T. Spahr, personal communication) - an old recipe for finding more NEOs (Bowell & Muinonen 1994). The result is that the CSS 1.5 m telescope covers about 3600 deg2 per lunation to V ∼ 21 mag while our PS1 simulation covers ∼3800 deg2 to a limiting magnitude of V ∼ 22.7. Based on these arguments our simulations suggest that the CSS made a low-probability discovery of 2008 TC3 - they were lucky. Another alternative – 19 – is that the SFD of the bolides from Brown et al. (2002) is in error - we consider this to be unlikely because of other corroborating research (e.g. Ivanov 2006). Like the CSS discovery of 2008 TC3, the problem is that PS1 will detect the object on only a single night resulting in an orbit with large uncertainties. PS1 can report single night detection pairs to the Minor Planet Center (MPC) but will then depend on rapid announcement by the MPC of likely impactors and prompt followup by other observatories around the world. We have considered the possibility that detections in the impactor's tracklet could pro- vide information to flag it as an imminent impactor based on rapid changes in the detections' trail length or orientation due to the effects of topocentric parallax. While the 15-30 min time difference between the detections in the tracklets probably does not provide enough leverage to make this method viable for Pan-STARRS it could be employed in the future by a dedicated impactor survey. Since ground-based all-sky bolide impact monitoring networks (e.g. Oberst 1998; Weryk 2008; Bland 2008) cover an extremely small fraction of the Earth's surface (∼1%; P. Brown, personal communication) it is exceedingly unlikely that an object imaged by PS1 will also be detected by a ground-based sensor. On the other hand, the detection of bolides from wide-coverage space-based systems (e.g. Koschny et al. 2004) is possible and may provide an opportunity to search through PS1 detections for pre-impact observations. Figure 14 shows the apparent sky-plane distribution for 3 m diameter bolides 1 day before impact. These small objects (e.g. 1-20 m) are visible for only a short time before impact and their apparent brightness is strongly affected by the phase angle so that they can only be detected near opposition. The location of 2008 TC3 3 days before impact is clearly within the PS1 opposition field. PS1 would have detected 2008 TC3 since it was brighter than V=22.7 mag (assuming H = 30.4 and G = 0.15) for 4–5 days before impact (with V = 18.9 mag at discovery 1 day before the impact) and the sky-plane motion was below the PS1 trailing cut-off of 12◦/day. Although the small impactors may be located over the entire sky there are two promi- nent clumps in Fig. 14 - towards the Sun and around the opposition point (the anti-solar direction). This result is consistent with the helion and antihelion sources (Poole 1997) of sporadic background bolides that are the result of the velocity vector summation of the Earth and the sporadic meteors. Jones & Brown (1993) had earlier suggested that these two sources of sporadic meteors were most consistent with a population of small meteoroids derived from a Jupiter Family Comet (JFC) parent population. We note that their study did not correct for strong selection effects in the radar detection of the meteors that would favor – 20 – the high velocity cometary sub-component of the helion and anti-helion sources. Considering that the Bottke NEO model employed here is dominated by asteroids but still produces these two sources it suggests that at least some component of the sporadic meteors in the helion and anti-helion regions may be of asteroidal origin. We believe that our results encourage optimism that the suite of next generation ground- based sky surveys will be able to eliminate ≥90% of the risk of an unanticipated impact of an object > 140 m diameter. Still, it is interesting to identify a survey strategy that could reduce the risk even further. Figure 15 shows that the undiscovered large (1 km) impactors rarely appear in the PS1 survey regions and are strongly clustered in the direction towards the Sun. This is not because the objects are on orbits interior to the Earth's orbit (e.g. Zavodny et al. 2008) but is due to their orbital period being close to one year. The synodic period with respect to Earth Psyn = P −1 asteroid−1 is thus long. The four year PS1 survey has a low efficiency for detecting objects with long synodic periods and the solution is to survey over a longer time period or to survey the sky at smaller solar elongations. The suite of next generation surveys (e.g. , Pan-STARRS, LSSTC) will survey the sky for about a decade and will increase the completion statistics at all impactor sizes. Surveying closer to the Sun is realistic only from future space-based platforms (e.g. Tedesco et al. 2000; Jedicke et al. 2003; Hildebrand et al. 2007; Mottola 2008). Earth − P −1 Figure 16 shows the orbital period for 1 km diameter impactors that strike the Earth during the 4 year survey mission. The orbital period of the undetected impactors peaks at 1 year as explained above. At first glance it may be surprising that a powerful survey system could miss these large impactors given the large distance at which 1 km diameter objects can be detected. The explanation is that all the undetected large impactors strike the Earth in the first 2 years; before the survey had sufficient time to explore the entire volume of the solar system to the distance at which 1 km diameter objects are detectable. 4. Conclusions/Discussion We have used a synthetic population of Earth-impacting asteroids to determine their sky-plane distribution and detectability with one of the next generation all-sky surveys, PS1. We find that • while the steady state sky plane distribution of the impactors is concentrated towards small solar elongation as shown by Chesley & Spahr (2004) their sky-plane distribution exhibits interesting behavior in the couple years leading up to impact. This behavior could be exploited to identify these objects well in advance of impact. – 21 – • the requirement that a PHO have a MOID<0.05 AU identifies 98% of all impactors even 100 years in advance of impact. The remaining 2% have much larger MOID and the error on the MOID will not identify these objects as potential impactors since they become PHOs only after encounter with another solar system object, usually Jupiter. However, numerical exploration of their orbit evolution as routinely performed by impact monitoring sites will generally reveal these hazardous objects long before impact. • the MOIDs determined for objects with detections obtained during just a single luna- tion are very good • the impactor model that we have developed is probably a good proxy for the unbi- ased orbit distribution of the large bolide population based on 1) the good agreement between the impactor orbit distribution and the orbit elements for impacting asteroid 2008 TC3 and 2) its sky-plane location at the time of discovery. • the next generation of all-sky surveys will identify a large fraction (& 90%) of impactors >140 m diameter in a decade-long survey designed to find them. • the impact awareness time before impact can be many decades for impactors of any size as long as they are discovered before impact. The most likely impact awareness time for smaller impactors is zero. • the next generation surveys will probably image a small impactor before impact but it will likely be observed over too short a time span (e.g. one night) to allow either an accurate pre-impact orbit to be computed or its identification as an imminent impactor. • a search for pre-impact detections of a bolide is possible • the most difficult objects to discover are those with long synodic periods relative to the Earth. • next generation surveys like PS1 will be efficient at detecting large impactors when their impact time is more than about 2 years after the start of the survey. – 22 – Acknowledgements This work was performed in collaboration with Spacewatch, LSSTC, CMU's AUTON lab, and the AstDyS team (Andrea Milani, Giovanni Gronchi and Zoran Knezevic). Steve Chesley's work was conducted at the Jet Propulsion Laboratory, California Institute of Technology, under a contract with the National Aeronautics and Space Administration. The design and construction of the Panoramic Survey Telescope and Rapid Response System by the University of Hawaii Institute for Astronomy is funded by the United States Air Force Research Laboratory (AFRL, Albuquerque, NM) through grant number F29601-02- 1-0268. MOPS is also supported by a grant (NNX07AL28G) to Robert Jedicke from the NASA NEOO program. Andrea Milani, Giovanni Gronchi and Zoran Knezevic of the Ast- DyS group provided critical orbit determination software to the MOPS team. The MOPS is currently being developed in association with the Large Synoptic Survey Telescope Cor- poration (LSSTC). The LSSTC's research and development effort is funded in part by the National Science Foundation under Scientific Program Order No. 9 (AST-0551161) through Cooperative Agreement AST-0132798. Additional funding to the LSSTC comes from private donations, in-kind support at Department of Energy laboratories and other LSSTC Institu- tional Members. Veres was supported by the National Scholarship Programme of the Slovak Republic, the European Social Fund, a Grant from Comenius University (No. UK/399/2008) and VEGA Grant No. 1/3067/06. – 23 – REFERENCES Abe, S. 2008. Meteoroids and Meteors - Observations and Connection to Parent Bodies. In (Eds. Mann, I., Nakamura, A. and Mukai, T.) Small Bodies in Planetary Systems, Lect. Notes Phys. 758. Astronomy and Astrophysics Survey Committee, Board on Physics and Astronomy, Space Studies Board, and National Research Council 2001. Astronomy and Astrophysics in the New Millennium. National Academy Press. Bland, P.A., Spurn´y, P., Shrben´y, L., Borovicka, J., Bevan, A.W.R., Towner, M.C., McClaf- ferty, T., Vaughan, D., Deacon, G., 2008. The Desert Fireball Network: First Results of Two Years Systematic Monitoring of Fireballs over the Nullarbor Desert of South Western Australia. Asteroids, Comets, Meteors conference, held in Baltimore, July 14 - 18, 2008, LPI Contribution No. 1405, paper id. 8246. Bottke, W.F., Rubincam, D.P., Burns, J.A., 1998. Dynamical Evolution of Meteoroids via the Yarkovsky Effect. AAS, DPS meeting #30, #10.02, Vol. 30, p.1029. Bottke, W.F., Jedicke, R., Morbidelli, A., Petit, J., Gladman, B., 2000. Understanding the Distribution of Near-Earth Asteroids. Science 288, 2190-2194. Bottke, W.D., Morbidelli, A., Jedicke, R., Petit, J. ,Levison, H.F., Michel, P., Metcalfe, T.S.,2002. Debiased Orbital and Absolute Magnitude Distribution of the Near-Earth Objects. Icarus 156, 399-433. Bowell, E., Muinonen, K., 1994. Earth-crossing asteroids and comets: Groundbased search strategies. In: Gehrels, T. (Ed.), Hazards due to Comets and Asteroids, , Univ. of Arizona Press, Tucson, pp. 149-197. Brasser, R., & Wiegert, P. 2008, MNRAS, 386, 2031. Brown, P., Spalding, R.E., ReVelle, D.O., Tagliaferii, E., Worden, S.P., 2002. The flux of small near-Earth objects colliding with the Earth. Nature 420, 294-296. Burke, B.E., Tonry, J.L., Cooper, M.J., Doherty, P.E., Loomis, A.H., Young, D.J., Lind, T.A., Onaka, P., Landers, D.J., Daniels, P.J., Daneu, J.L., (2007. In: Blourke, Morley, M. (Eds.), Orthogonal transfer arrays for the Pan-STARRS gigapixel camera. Sensors, Cameras, and Systems for Scientific/Industrial Applications VIII, Proceedings of the SPIE 6501, 650107. Ceplecha, Z., Borovicka, J., Elford, W.G., ReVelle, D.O., Hawkes, R.L., Porubcan, V., Simek, M., 1998. Meteor Phenomena and Bodies. Space Science Reviews 84, 327-471. – 24 – Chesley, S., Milani, A., 1999. NEODyS: an online information system for near-Earth objects. AAS, DPS meeting #31, #28.06. Chesley, S.R., Spahr T.B., 2004. Earth-impactors: orbital characteristics and warning times. In: Belton, M.J.S., Morgan, T.H., Samarashinha, N.H., Yeomans,D.K. (Eds.), Mit- igation of Hazardous Comets and Asteroids, , Cambridge Univ. Press, Cambridge, pp.22-37. Chodas, P.W., Giorgini, J.D., 2008. Impact warning times for Near-Eearth asteroids. As- teroids, Comets, Meteors conference, held in Baltimore, July 14 - 18, 2008, LPI Contribution No. 1405, paper id. 8371. Denneau, L., Kubica, J., Jedicke, R., 2007. The Pan-STARRS Moving Object Pipeline. Astron. Data Analysis Software and Systems XVI Conference Series, Vol. 376, Shaw, R.A., Bell, D.J. (Eds.), p.257. Drummond, J.D., 1981. A test of comet and meteor shower associations. Icarus 45, 545-553. Earth Impact Database, 2008. http://www.unb.ca/passc/ImpactDatabase/index.html Farinella, P., Vokrouhlicky, D., and Hartmann, W. K. 1998. Meteorite Delivery via Yarkovsky Orbital Drift. Icarus 132, 378-387. Gallant, J., Gladman, B., and ´Cuk, M. 2006. Current bombardment of the Earth-Moon system: Emphasis on cratering asymmetries. eprint arXiv:astro-ph/0608373. Galligan, D. P. 2001. Performance of the D-criteria in recovery of meteoroid stream orbits in a radar data set. MNRAS327, 623-628. Granvik, M., Virtanen, J., Muinonen, K., 2008. OpenOrb: Open-Source Asterod-Orbit- Computation Software Including Statistical Orbital Ranging. Asteroids, Comets, Me- teors 2008. LPI Contribution No. 1405, paper id. 8206. Grav, T., Jedicke, R., Denneau, L., Chesley, S., Holman, M.J., Spahr, T.B., 2009. The Pan-STARRS Synthetic Solar System Model: A tool for testing and efficiency deter- mination of the Moving Object Processing System. Icarus x, xx-xx. Gehrels, T., 1986. CCD Scanning. In: Lagerkvist, C.-I., Lindblad, B.A., Lundstedt, H., Rickman, H. (Eds.), Asteroids, Comets, and Meteors II, Univ. of Arizona Press, Tucson, pp. 19-20. Harris, A.W., Bowell, E.L.G., 2004. LSST Solar System Survey - Cadence and Sky Coverage Requirements. Bulletin of the Am. Astron. Soc. 36, p.1530. – 25 – Harris, A.W., 2007. What Spaceguard did. Nature 453, 1178-1179. Harris, A.W., 2008. An Update of the Population of NEAs and Impact Risk. AAS, DPS meeting #39, #50.01. Helin, E.F., Shoemaker, E.M., 1979. The Palomar planet-crossing asteroid survey, 1973-1978. Icarus 40, 321-328. Hildebrand, A. R., Tedesco, E. F., Carroll, K. A., Cardinal, R. D., Matthews, J. M., Kuschnig, R., Walker, G. A. H., Gladman, B., Kaiser, N. R., Brown, P. G., Lar- son, S. M., Worden, S. P., Wallace, B. J., Chodas, P. W., Muinonen, K., Cheng, A., Gural, P., 2007. The Near Earth Object Surveillance Satellite (NEOSSat) Mission Enables an Efficient Space-Based Survey(NESS Project) of Interior-to-Earth-Orbit (IEO) Asteroids. 38th Lunar and Planetary Science Conference, LPI Contribution No. 1338, p. 2372. Ivanov, B.A., 2006. Earth/Moon impact rate comparison: Searching constraints for lunar secondary/primary cratering proportion. Icarus 183, 504-507. Ivezi´c, Z., Tyson, J.A., Juri´c, M., Kubica, J., Connolly, A., Pierfederici, F., Harris, A.W., Bowell, E. & the LSST Collaboration, 2007. LSST: Comprehensive NEO detection, characterization, and orbits. In: Milani, A., Valsecchi, G.B., Vokrouhlick´y, D. (Eds.), Near Earth Objects, our Celestial Neighbors: Opportunity and Risk, IAU Symposium #236, Cambridge Univ., Cambridge, pp. 353-362. Jedicke, R., Morbidelli, A., Petit J., 2003. Earth and space-based NEO survey simulations: prospects for achieving the spaceguard goal. Icarus 161, 17-33. Jedicke, R., Denneau, L., Grav, T., Heasley, J., Kubica, J., Pan-STARRS Team, 2005. Pan- STARRS Moving Object Processing System. Bulletin of the Am. Astron. Soc. 37, p.1363. Jedicke, R.,Magnier, E.A., Kaiser, N., Chambers, K.C., 2006. The Next Decade of Solar System Discovery with Pan-STARRS. In: Milani, A., Valsecchi, G.B., Vokrouhlick´y, D. (Eds.), Proceedings IAU Symposium No. 236, Cambridge Univ. Press, Cambridge, pp.341-352. Jenniskens, P. 2006. Meteor Showers and their Parent Comets. Meteor Showers and their Parent Comets, by Peter Jenniskens, pp. . ISBN 0521853494. Cambridge, UK: Cam- bridge University Press, 2006.. Jones, J., & Brown, P. 1993, MNRAS, 265, 524 – 26 – Kaiser, N., Pan-STARRS Team, 2005. The Pan-STARRS Survey Telescope Project. Bulletin of the Am. Astron. Soc. 37, p. 1409. Kessler, D. J., Landry, P. M., Gabbard, J. R., & Moran, J. L. T. 1980, Solid Particles in the Solar System, 90, 137. Koehn, B.W., Bowell, E., 1999. Enhancing the Lowell Observatory Near-Earth-Object Search, AAS, DPS meeting #31, #12.02. Koschny, D., di Martino, M., Oberst, J., 2004. Meteor observation from space - The Smart Panoramical Optical Sensor (SPOSH). In: Triglav- Cekada, M., Trayner, C.(Eds.), Proceedings of the International Meteor Conference, International Meteor Organiza- tion, pp. 64-69. Larsen, J. A., and 15 colleagues 2007. The Search for Distant Objects in the Solar System Using Spacewatch. AJ133, 1247-1270. Larson, S., Brownlee, J., Hergenrother, C., Spahr, T., 1998. The Catalina Sky Survey for NEOs. Bulletin of the Am. Astron. Soc. 30, p. 1037. Lindblad, B.A., Neslusan, L., Svoren, J., Porubcan, V., 2003. IAU Meteor Database of photographic orbits version 2003. Earth, Moon and Planets 93, 249-260. Masiero, J., Jedicke, R., Pravec, P., Durech, J., Gwyn, S., Denneau, L., Larsen, J., 2009. The Thousand Asteroid Light Curvey Survey. Submitted to Icarus. Milani, A., Gronchi, G. F., Farnocchia, D., Knezevi´c, Z., Jedicke, R., Denneau, L., and Pierfederici, F. 2008. Topocentric orbit determination: Algorithms for the next gen- eration surveys. Icarus 195, 474-492. Milani, A., Chesley, S. R., Sansaturio, M. E., Tommei, G., and Valsecchi, G. B. 2005. line of variation searches for impactors. Icarus 173, Nonlinear impact monitoring: 362-384. Monet, D. G., and 28 colleagues 2003. The USNO-B Catalog. AJ125, 984-993. Moon, H.-K., Byun, Y.-I., Raymond, S.N., Spahr. T., 2008. Realistic survey simulations for kilometer class near Earth objects. Icarus 193, 53-73. Morrison, D., 1992. The Spaceguard Survey: Report of the NASA International Near-Earth- Object Detection Workshop. NASA, Washington, D.C.. – 27 – Mottola, S., Borner, A., Grundmann, J. T., Hahn, G. J., Kazeminejad, B., Kuhrt, E., Michaelis, H., Montenegro, S., Schmitz, N., Spietz, P.,2008. AsteroidFinder: A Space- Based Search for IEOs. Asteroids, Comets, Meteors conference, held in Baltimore, July 14 - 18, 2008, LPI Contribution No. 1405, paper id. 8140. O'Brien, D.P., Greenberg, R., 2005. The collisional and dynamical evolution of the main-belt and NEA size distributions. Icarus 178, 179-212. Oberst, J., Molau, S., Heinlein, D., Gritzner, C., Schindler, M., Spurn´y, P., Ceplecha, Z., Rendtel, J., Betlem, H., 1998. The "European Fireball Network": Current status and future prospects. Meteoritics & Planetary Science 33, 49-55. Poole, L.M.G., 1997. The structure and variability of the helion and antihelion sporadic meteor sources. Monthly Notices of the Royal Astron. Soc. 290, 245-259. Pravdo, S.H., Rabinowitz, D.L., Helin, E.F., Lawrence, K.J., Bambery, R.J., Clark, C.C., Groom, S.L., Levin, S., Lorre, J., Shaklan, S.B., Kervin, P., Africano, J.A., Sydney, P., Soohoo, V., 1999. The Near-Earth Asteroid Tracking (NEAT) Program: an Au- tomated System for Telescope Control, Wide-Field Imaging, and Object Detection. The Astronomical Journal 117, 1616-1633. Simek, M., 1995. Diurnal and Seasonal Variations of Sporadic Meteor Parameters Summer and Winter Periods. Earth, Moon, and Planets 68, 545-553. Shoemaker, E.M., 1983. Asteroid and comet bombardment of the earth. In: Annual review of earth and planetary sciences, Volume 11, Annual Reviews Inc., Palo Alto, pp. 461-494. Sitarski, G., 1968. Digital Computer Solution of the Equation of Position of a Comet on a Keplerian Orbit. Acta Astronomica 18, 197-205. Southworth, R.B., Hawkins, G.S., 1963. Statistics of meteor streams. Smithsonian Contri- butions to Astrophysics 7, 261-285. Stokes, G.H., Evans, J.B., Viggh, H.E.M., Shelly, F.C., Pearce, E.C., 2000. Lincoln Near- Earth Asteroid Program (LINEAR). Icarus 148, 21-28. Stokes, G. H., Evans, J.B., Larson, S.M., 2002. Near-Earth asteroid search programmes. In: Bottke, W., Cellino, A., Paolicchi, P., Binzel, R.P. (Eds.), Asteroid III, Univ. of Arizona Press, Tucson, pp. 45-54. Stokes et al. 2004. http://neo.jpl.nasa.gov/neo/report.html. – 28 – Stuart, J.S., Binzel, R.P., 2004. Bias - corrected population, size distribution, and impact hazard for the near - Earth objects. Icarus 170, 295-311. Tancredi, G. 1998, Celestial Mechanics and Dynamical Astronomy, 69, 119 Tedesco, E.F., Muinonen K., Price S.D., 2000. Space-based infrared near-Earth asteroid survey simulation. Planetary and space science 48, 801-816. Valsecchi, G.B., Jopek, T.J., Froeschle, Cl., 1999. Meteoroid stream identification: a new approach - I. Theory. MNRAS 304, 743-750. Weryk, R.J., Brown, P.G., Domokos, A., Edwards, W.N., Krzeminski, Z., Nudds, S.H., Welch, D.L., 2008. The Southern Ontario All-sky Meteor Camera Network. Earth, Moon, and Planets 102, 241-246. Zavodny, M., Jedicke, R., Beshore, E. C., Bernardi, F., and Larson, S. 2008. The orbit and size distribution of small Solar System objects orbiting the Sun interior to the Earth's orbit. Icarus198, 284-293. This preprint was prepared with the AAS LATEX macros v5.2. – 29 – 2008 TC3 impactors sporadic background fireballs NEO model 0.10 0.08 0.06 0.04 0.02 0.00 0 1 2 Semimajor Axis (AU) 3 4 5 2008 TC3 impactors sporadic background fireballs NEO model 0.12 0.10 0.08 0.06 0.04 0.02 0.4 Eccentricity 0.6 0.8 1.0 impactors sporadic background fireballs NEO model 40 80 Inclination 120 160 Fig. 1.- Semi-major axis (top), eccentricity (middle) and inclination (bottom) distribution for (solid) our synthetic impactor population, (dotted) sporadic background fireballs and (dashed) the Bottke NEO model. The semi-major axis distribution is not shown beyond 5.1 AU but these data are included in the eccentricity and inclination distributions. The corresponding orbital elements for 2008 TC3 are indicated by the solid vertical line in each figure. n o i t c a r F n o i t c a r F n o i t c a r F 0.00 0.0 0.2 0.20 2008 TC3 0.15 0.10 0.05 0.00 0 – 30 – R C E n m i R C E e d o m R C E x a m orbital epochs 100 years 50 years 10 years 1 year before impact 1 0.1 n o i t c a r F 0.01 1E-3 1E-4 1E-7 1E-6 1E-5 1E-4 1E-3 0,01 0,1 1 orbital epochs 30 days 10 days 5 days 1 days before impact MOID (AU) R C E e d o m R C E x a m R C E n m i 1 0.1 n o i t c a r F 0.01 1E-3 1E-4 1E-7 1E-6 1E-5 1E-4 1E-3 0,01 0,1 1 MOID (AU) Fig. 2.- Fractional differential (solid) and cumulative (dashed) distribution of MOID for synthetic impactors at different times before impact. The Earth capture radius (ECR) is calculated for the highest (max ECR), most probable (mode ECR) and lowest (min ECR) impactor encounter velocity. The bin size is log(MOID/AU) = 0.07 in both figures. – 31 – ) % ( e r u l i a F y t i l i b a b o r P O H P 2.0 1.5 1.0 0.5 0.0 1 10000 100 10 Time Before the Impact (days) 1000 Fig. 3.- The probability that an Earth-impacting asteroid will not be identified as a PHO (MOID> 0.05AU) as a function of time before impact. The superimposed line is a fit (R2 = 0.99) to an empirically defined function: p = 1.3 × 10−4 + t0.92. – 32 – 10 1000 Time Before the Impact (days) 100 10 1 0.1 0.01 1E-3 0.1 1 1E-3 1E-4 1E-5 10000 max ECR mode ECR min ECR t p e c r e n t I U A 4 - 0 1 = D O M I ) U A ( I D O M e b a b o r P l t s o M 1 10 100 1000 10000 Time Before the Impact (days) Fig. 4.- (top) The constant (f ′) in a fit (log f = log f ′ + m log[MOID/(10−4AU)] with f < 0.9) to the cumulative fraction (f ) of impactors with MOID as a function of time before i.e. f ′ is the cumulative fraction of objects with MOID≤ 10−4 AU impact from Figure 2. at each of the specified time steps. The superimposed solid line is a fit (R2 = 0.954) to log f ′ = a[log2(t/days) − log2(t0/days)] with t being the an empirically defined function: time before impact, a = −0.079 and t0 = 72.8 days. (bottom) The most probable MOID (mode) for the impactors as a function of time before impact. The superimposed solid line is a fit (R2 = 0.90) to an empirically defined function: MOID (AU)= −0.014 − 1.5 × 10−3 ln [((t/days) + 105)]. The Earth capture radius is shown for the highest (max ECR), most probable (mode ECR) and lowest (min ECR) encounter velocity for the impactors. 100 – 33 – 0 -100 50 0 -50 e d u t i t a L c i t p i l c E 50 0 -50 50 0 -50 100 Opposition Centered Ecliptic Longitude -100 0 Fig. 5.- Topocentric opposition centered ecliptic sky plane distribution of H = 20 (∼350 m diameter) Earth-impacting asteroids 20 years before impact for a survey with a limiting magnitude of (top) V = 22.7 (middle) V = 20.7 and (bottom) V = 18.7. This is sufficiently in advance of impact to represent the impactors's steady-state sky plane distribution. – 34 – n o i t c a r F 0.15 0.10 0.05 0.00 -50 0 50 Ecliptic Latitude Fig. 6.- Ecliptic latitude distribution of Earth-impacting asteroids with V < 22.7 twenty years before impact. 100 – 35 – 0 -100 50 0 -50 e d u t i t a L c i t p i l c E 50 0 -50 50 0 -50 100 Opposition Centered Ecliptic Longitude -100 0 Fig. 7.- a) Topocentric opposition centered ecliptic sky plane distribution of Earth- impacting asteroids with V ≤ 22.7 (top) 1 day, (middle) 30 days, and (bottom) 60 days before impact. 100 – 36 – 0 -100 50 0 -50 e d u t i t a L c i t p i l c E 50 0 -50 50 0 -50 100 Opposition Centered Ecliptic Longitude -100 0 Fig. 7.- b) The opposition centered ecliptic sky plane distribution of Earth-impacting asteroids with V ≤ 22.7 (top) 90 days, (middle) 120 days, and (bottom) 150 days before impact. 100 – 37 – 0 -100 50 0 -50 e d u t i t a L c i t p i l c E 50 0 -50 50 0 -50 100 Opposition Centered Ecliptic Longitude -100 0 Fig. 7.- c) The opposition centered ecliptic sky plane distribution of Earth-impacting asteroids with V ≤ 22.7 (top) 180 days, (middle) 1 year, and (bottom) 1.5 years before impact. – 38 – 50 0 -50 100 -100 Opposition Centered Ecliptic Longitude 0 e d u t i t a L c i t p i l c E Fig. 7.- d) The opposition centered ecliptic sky plane distribution of Earth-impacting asteroids with V ≤ 22.7 2 years before impact. – 39 – t s s e n e e p m o C l impactor diameter 1000m 500m 200m 100m 50m 20m 1.0 0.8 0.6 0.4 0.2 0.0 0.0 0.5 1.0 1.5 2.0 2.5 3.0 Survey Time (Years) 74% 65% 43% 26% 7% 0.3% 3.5 4.0 t s s e n e e p m o C l 1 0.1 0.01 1E-3 1E-4 1E-5 1 10 100 1000 Impactor Diameter (m) Fig. 8.- (left) PS1 survey completeness (efficiency) as a function of time for different size impactors assuming an albedo of 0.14. The percentages on the right side of the figure adjacent to the efficiency curves are the efficiency for detecting an impactor that strikes the Earth during the 4 year survey. (right) The efficiency for discovering objects impacting during the 4 year survey as a function of impactor diameter. A linear extrapolation to D < 20 m using the two leftmost data points is shown with a dashed line. – 40 – 0.075 0.050 0.025 n o i t c a r F 0.1 0.01 1E-3 1E-4 ) U A ( I D O M 1E-5 1E-6 1E-7 1E-7 1E-6 1E-5 1E-4 1E-3 Derived MOID (AU) 0.01 0.1 0.025 0.050 0.075 Fraction 9.- (lower-left) The the MOID Fig. the derived MOID for impactors detected in the (MOIDsynthetic−MOIDderived) vs. simulation. (right) Distribution of the error in the MOID. (top) Distribution of the derived MOID. distribution error in of the – 41 – impactor diameter 1000m 500m 200m 80 100 40 60 Impact Awareness Time (years) impactor diameter 100m 50m 20m 80 100 40 60 Impact Awareness Time (years) 0.15 0.10 0.05 n o i t c a r F 0.00 0 20 0.15 0.10 0.05 n o i t c a r F 0.00 0 20 Fig. 10.- Distribution of impact awareness times (defined in the text) for discovered syn- thetic impactors of different sizes. Note that the awareness time is zero for undiscovered impactors and they are not included in this figure. – 42 – future impacts instant impacts 3 C T 8 0 0 2 1 10 100 Impactor Diameter (m) 1000 1 0.1 0.01 1E-3 1E-4 1E-5 s r o t c a p m I d e r e v o c s D i f o r e b m u N Fig. 11.- (solid) Expected number of detections of large (> 20 m diameter) impactors that will strike the Earth in the next 100 years that are observed 3 nights in a single lunation during a four year PS1 survey and (dashed) the number of small (1-20 m diameter) impactors detected by the PS1 system in the lunation prior to impact. The probable size range of the impacting asteroid 2008 TC3 is indicated by the shaded region. – 43 – 1day 5days 10days before the impact 100 10 Diameter (m) 0.1 0.01 1E-3 1E-4 1 t s s e n e e p m o C l Fig. 12.- Estimated four year PS1 survey detection efficiency 1, 5, and 10 days before impact for small impactors on their final approach to the Earth. – 44 – 1 day 5 days 10 days before the impact 0.01 0.1 1 10 Apparent Motion (deg/day) n o i t c a r F 0.16 0.14 0.12 0.10 0.08 0.06 0.04 0.02 0.00 1E-3 Fig. 13.- Distributions of the apparent rate of motion for small impactors 1, 5, and 10 days before impact. There is no restriction on apparent magnitude. The upper limit on the detectable rate of motion imposed by the Pan-STARRS funding agency is 12 deg/day. – 45 – 2008 TC3 e d u t i t a L c i t p i l c E 50 0 -50 50 0 -50 100 0 -100 Opposition Centered Ecliptic Longitude Fig. 14.- (top) Sky-plane probability distribution for 3 m diameter impactors (bolides) 1 day before impact with no restrictions on brightness and apparent motion. (bottom) Same as above but with V < 22.7 and apparent motion < 12 deg/day. Approximate PS1 search regions are shown as solid rectangles (the opposition region is in the center and the sweet spots are on the left and right). The position of 2008 TC3 at discovery by the Catalina Sky Survey is highlighted. – 46 – e d u t i t a L c i t p i l c E 50 0 -50 50 0 -50 100 0 -100 Opposition Centered Ecliptic Longitude Fig. 15.- (top) Sky-plane position probability distribution for 1 km diameter impactors that were not discovered during the simulated 4 year PS1 survey mission with no restriction on the apparent magnitude. Darker regions represent where the objects are most likely to be found. (bottom) Same as above but imposing a PS1 limiting magnitude cutoff at V = 22.7. Approximate PS1 search regions are shown as solid rectangles. – 47 – 1km size population undetected found 1 2 3 4 Orbital period (years) 100 80 60 40 20 0 r e b m u N Fig. 16.- Orbital period distribution for found (dotted) and undetected (solid) 1 km diam- eter impactors during the 4 year PS1 survey mission.
1701.01661
1
1701
2017-01-06T15:24:09
Formation and evolution of exoplanets in different environments
[ "astro-ph.EP" ]
The ultimate goal of exoplanetologists is to discover life outside our Earth and to fully understand our place in the Universe. Even though we have never been closer to attaining this goal, we still need to understand how and where the planets (efficiently) form. In this manuscript I briefly discuss the important role of stellar metallicity and chemistry on the formation and evolution of exoplanets.
astro-ph.EP
astro-ph
Formation and evolution of exoplanets in different environments Vardan Adibekyan1 1Instituto de Astrofísica e Ciências do Espaço, Universidade do Porto, CAUP, Rua das Estrelas, 4150-762 Porto, Portugal; vadibekyan@ astro.up.pt Abstract. The ultimate goal of exoplanetologists is to discover life outside our Earth and to fully understand our place in the Universe. Even though we have never been closer to attaining this goal, we still need to understand how and where the planets (efficiently) form. In this manuscript I briefly discuss the important role of stellar metallicity and chemistry on the formation and evolution of exoplanets. 1. Introduction Ever since the first giant exoplanet was discovered orbiting a Sun-like star about twenty years ago (Mayor & Queloz 1995), the search has been ongoing for small, rocky planets around other stars, evocative of Earth and other terrestrial planets in the Solar System. As of today, there are more than 3500 planets detected1 and several thousand candidates (Coughlin et al. 2016) waiting for validation. These discoveries helped us to understand that extra-solar planets are very common in our Galaxy. The diversity of the discovered planets is astonishing, and most of detected planets brought us more questions than answers. While the Universe is full of surprises, we (exoplanetologists) are drawing closer to the answer to the most daring questions of humankind: are we alone in the Universe and what is our place in there? In fact, the last "simplistic" calculations2 of Behroozi & Peeples (2015) shows that the chance that we are the only civilization the Universe will ever have is less than 8%. In this manuscript first I start by briefly presenting how different are the proper- ties of so-far detected exoplanets and how the two completely (ideologically) different theories are getting close to explain the formation and evolution of these planets. In the second part of the paper I discuss the importance of chemical conditions of the environment where the exoplanets form. 2. The zoo of exoplanets and their formation scenarios Among the few thousands of the detected exoplanets, we observed many words that are very different from what we have in our Solar System and from what we could imagine. We observed a planet with an extremely eccentric orbit of 0.97 (HD20782b -- O'Toole 1exoplanet.eu 2The authors did not consider requirements of individual elemental abundances for planet formation (e.g. Adibekyan et al. 2012b, 2015). 1 2 Vardan Adibekyan et al. 2009), a planet in a circumbinary orbit surrounded by four suns (PH1b3/Kepler- 64b -- Schwamb et al. 2013), a dense "superplanet"4 with a radius of only ∼ 1R(cid:88) and mass of ∼ 22M(cid:88) (CoRoT-3b -- Deleuil et al. 2008), a very hot planet with a surface day-side temperature of more than 9000 K orbiting its pulsating hot sun in less than six hours (Kepler-70b/KOI-55.01 -- Charpinet et al. 2011). The detection of these weird exoplanets5 puts to shame many old science fiction stories and make harder the job of new science fiction writers. Left: The distribution of discovered planets in the period -- mass diagram. Figure 1. Right: Mass of the known planets as a function of the discovery year. Different symbols represents planets discovered by different detection techniques. Some of the planets of our Solar System are shown for reference. The distribution of exoplanets in the period -- mass diagram is shown in Fig. 1. First, the plot shows that different planet detection techniques occupy different regions of this diagram. Second, one can see that the detected planets, while being very diverse in their properties (see previous paragraph), are clustered in three main groups: hot- Jupiters (Mp ∼1-2M(cid:88) and P (cid:46) 10 days), hot/warm super-Earths/Neptunes (Mp ∼10M⊕ and P (cid:46) 100 days), and gas and ice giants (Mp ∼1-2M(cid:88) and P ∼ 1000 days). We note that this diagram is strongly constrained by the biases and detection limits of different techniques. In particular the detection limits of these techniques and current instrumen- tation is responsible for the empty bottom-right triangle of the figure. However, some of the observed features, such as the "period-valley" (a lack of giant planets with periods between 10-100 days: Udry et al. 2003) or the sub-Jovian desert (a lack of sub-Jupiter mass planets at orbital periods shorter than 3 days: Szabó & Kiss 2011) are probably physical and give important insights for our understanding of exoplanet formation and evolution. For a recent excellent review on the architecture of exoplanetary systems we refer the reader to Winn & Fabrycky (2015). A logical question now to ask is how do these very different planets form? Cur- rently two main mechanisms are proposed for the formation of exoplanet that are con- 3This planet was first discovered by two citizen scientists. 4Note that this sub-stellar object can be a low-mass brown-dwarf. 5https://en.wikipedia.org/wiki/List_of_exoplanet_extremes 100101102103104105106Period(day)101100101102103104Mass(M⊕)ImagingTransit/TTVRV19952000200520102015YearofdiscoveryImagingTransit/TTVRV Exoplanets and metallicity 3 ceptually different. In the so called core-accretion (CA) model low-mass planets form from the coagulation of very small solid bodies (Pollack et al. 1996). If before the dissipation of the protoplanetary disk, a core of about 5-10 M⊕ is formed then, it can undergo runaway accretion of gas and form a giant planet. In the so called gravitational instability (GI), in the gaseous disk (usually massive and cold), localized instabilities collapse into giant planets (Boss 1998). These two models have experienced substantial development and modifications, and the most recent and advanced ones (e.g. Nayak- shin 2016; Bitsch et al. 2015; Levison et al. 2015) include important phenomena such us pebble accretion (e.g. Johansen & Lacerda 2010) and/or migration in the disk (e.g. Alibert et al. 2004). Planetary population synthesis calculations (Ida & Lin 2004) based both on CA (e.g. Mordasini et al. 2009; Hasegawa & Pudritz 2013) and GI followed by tidal downsizing (TD -- e.g. Forgan & Rice 2013; Nayakshin 2016) reproduce many of the properties of the observed exoplanets. We refer the reader to Mordasini et al. (2015) for a recent review on the Global models of planet formation and evolution. 3. Exoplanets and stellar metallicity The correlation between stellar metallicity and the occurrence rate of giant planets is a firmly established fact (e.g. Gonzalez 1997; Santos et al. 2001), however, the exact functional form of this dependence is not fully established yet (see left panel of Fig. 2; Mortier et al. 2013). This observational result got its theoretical support first in CA (e.g. Mordasini et al. 2009) and then in the TD (Nayakshin 2016)6. Despite the large amount of observational data, it is still not clear if the planet-metallicity correlation holds for low-mass/small-sized planets (see right panel of Fig. 2; Sousa et al. 2011; Mayor et al. 2011; Wang & Fischer 2015; Buchhave & Latham 2015; Zhu et al. 2016). This is probably because it is hard to detect these light planets (especially at large distances) and it is very difficult to create a comparison sample of stars without low-mass planets. Left - Mortier et al. (2013): Frequency of giant planets as a function of Figure 2. metallicity and mass of the HARPS + CORALIE sample. Different functional forms are shown in different colors . Right - Mayor et al. (2011): The metallicity ([Fe/H]) distribution of stars hosting giant gaseous planets (black), planets less massive than 30 M⊕ (red), and for the global combined sample stars (blue). The latter histogram has been divided by 10 for the sake of visual comparison. 6Note that most of the GI based models do not predict a strong correlation between giant planet frequency and metallicity (e.g. Boss 1998). 4 Vardan Adibekyan The importance of metallicity is not only limited to the formation efficiency of planets. Metallicity also determines the maximal mass of the exo-Neptunes (Courcol et al. 2016), the presence or absence of gaseous atmosphere of small-sized planets (Dawson et al. 2015), and the mass of the core (heavy elements) of giant planets (e.g. Miller & Fortney 2011). Figure 3. The orbital semi-major axis of low-mass and small-size planets orbiting FGK dwarf stars. Planets detected by the RV and Transit techniques are shown with filled circles and empty star-symbols, respectively. Blue color corresponds to planets orbiting metal-rich stars ([Fe/H] ≥ -0.1) and red color corresponds to planets around metal-poor stars ([Fe/H] < -0.1). The habitable zone (Kopparapu et al. 2013) for stars of different masses is highlighted by blue shade. It is also interesting to note that the final orbital separation of planets shows a dependence on metallicity of the system (Adibekyan et al. 2013b; Beaugé & Nesvorný 2013; Mulders et al. 2016; Adibekyan et al. 2016). Adibekyan et al. (2016), based on the previous results that low-mass and small-sized planets orbiting around metal-rich stars do not have long orbits (Adibekyan et al. 2013b), suggested that planets in the "habitable zone" should be preferentially less metallic than our Sun. Fig. 3 shows the orbital distance of low-mass (detected with RV) and small-radius (detected by transit method) planets against the mass of their host stars. The plot is based on the data of Adibekyan et al. (2016) and illustrates their findings. Here we should note that (Mulders et al. 2016) observed several Kepler planet candidates orbiting their metal-rich stars at long periods7. However, they also observed that the planet occurrence rate is two times higher for metal-poor systems when compared to the systems with super-solar metallicities. 4. Exoplanets and stellar chemistry In stellar astrophysics, the iron content is usually used as a proxy for overall metallicity and most of the aforementioned studies followed this trend. Several works, however, 7Note that the sample of (Mulders et al. 2016) consists of Kepler planet candidates and not only confirmed planets. 00.20.40.60.81Distance(AU)0.60.70.80.91.01.1Mass(Mfl)HD40307g Exoplanets and metallicity 5 searched for chemical peculiarities of planet hosting stars in terms of abundances of in- dividual elements. While many contradictory results can be found in the literature (e.g. Bodaghee et al. 2003; Robinson et al. 2006; Brugamyer et al. 2011; Suárez-Andrés et al. 2016a,b), the enhancement of α-elements of iron-poor planet hosts was shown to be robust (Haywood 2008; Kang et al. 2011; Adibekyan et al. 2012a,b). Interest- ingly, Adibekyan et al. (2012a) showed that even low-mass/small-radius planets show α-enhancement at low-iron regime. The right panel of Fig. 4 depicts the α-enhancement (here Si abundance is used as a proxy for α-elements) of iron-poor planet hosts for the HARPS sample of Adibekyan et al. (2012c). The enhancement in α-elements rela- tive to iron is typical for the thick disk stars (e.g. Fuhrmann 1998; Adibekyan et al. 2013a). In fact, the HARPS data suggests that the planet formation frequency is about 5.5 times higher in the thick disk (12.3±4.1%) when compared to the Galactic thin disk (2.2±1.3%) in the metallicity range of -0.7 < [Fe/H] < -0.3 dex. Gonzalez (2009) recommended to use a so-called refractory index "Ref", which quantifies the mass abundances of refractory elements (Mg, Si and Fe) important for planet formation, rather than [Fe/H]. The importance of this index increases in the Fe- poor region (Adibekyan et al. 2012c; Gonzalez 2014b) when one compares statistics of planets around the thin disk and thick disk stars. Figure 4. [Si/Fe] versus [Mg/Si] and [Si/Fe] versus [Fe/H] for stars with and with- out (gray dots) planets. The red squares refer to the Jovian hosts and the blue circles refer to the stars hosting exclusively Neptunians and super-Earths (M < 30 M⊕). Three stars that are hosting low-mass/small-radius planets with precise radius and mass determinations are presented with a symbol of star (Santos et al. 2015). The position of the Sun is marked with the modern sun symbol. The studies of individual heavy elements and specific elemental ratios in stars with planets are very important because they are expected to control the structure and composition of terrestrial planets (e.g. Grasset et al. 2009; Bond et al. 2010; Thiabaud et al. 2014; Dorn et al. 2015). In particular, Mg/Si and Fe/Si ratios are important to constrain the internal structure of terrestrial planets (Dorn et al. 2015). Recently, Santos et al. (2015) tested these models on three terrestrial planets (see Fig. 4) and showed that 0.40.20.00.2[Fe/H]CoRoT7Kepler93Kepler100.10.00.1[Mg/Si]0.050.000.050.100.150.200.25[Si/Fe]CoRoT7Kepler93Kepler10 6 Vardan Adibekyan the iron mass fraction inferred from the mass-radius relationship is in good agreement with the iron abundance derived from the host star's photospheric composition. 5. Exoplanets and Galactic chemical evolution As discussed in the previous sections exoplanet formation efficiency, the type of the planets formed and their orbital characteristics depend on metallicity and chemical con- ditions. Putting all these results together one can reach to an interesting conclusion (or perhaps a speculation): i) Planets orbiting their stars in the circumstellar habitable zone have sub-solar metallicities (Fig. 3). ii) These iron-poor stars are usually enhanced in α- elements (i.e. high Si/Fe ratio) and at the same time have high Mg/Si ratio (Fig. 4). iii) High [Mg/Si] and low [Fe/Si] abundance ratio should produce a planet of a composition and structure that is different than ours (Dorn et al. 2015). Metallicity and abundance of different elements important for planet formation varies with time and location in our Galaxy and in Universe in general. Several studies during the last decade tried to predict prevalence of terrestrial planets in our Galaxy and in the so called "Galactic habitable zone"8 (e.g. Lineweaver et al. 2004; Gowanlock et al. 2011; Gonzalez 2014a; Gobat & Hong 2016). Some other studies extended these works to the observable Universe (e.g. Behroozi & Peeples 2015; Zackrisson et al. 2016). We refer the reader to these interesting works for more information and details about the recent interpretations of evolution of life across space and time. 6. Conclusion Formation efficiency, composition, structure, and even "habitability" of planets depend on the chemical conditions of the environment they form i.e. time and place in the Galaxy. Acknowledgments. I thank the science organizing committee of Non-Stable Uni- verse: Energetic Resources, Activity Phenomena and Evolutionary Processes for the invitation to participate to this interesting conference. I acknowledge the support from Fundação para a Ciência e Tecnologia (FCT) through national funds and from FEDER through COMPETE2020 by the following grants UID/FIS/04434/2013 & POCI-01- 0145-FEDER-007672, PTDC/FIS-AST/7073/2014 & POCI-01-0145-FEDER-016880 and PTDC/FIS-AST/1526/2014 & POCI-01-0145-FEDER-016886. I also acknowl- edges the support from FCT through Investigador FCT contracts of reference IF/00650/2015. Finally, I would like to thank Pedro Figueira for his interesting comments and discus- sion. References Adibekyan, V., Figueira, P., & Santos, N. C. 2016, Origins of Life and Evolution of the Bio- sphere, 46, 351. 1509.02429 Adibekyan, V., Santos, N. C., Figueira, P., Dorn, C., Sousa, S. G., Delgado-Mena, E., Israelian, G., Hakobyan, A. A., & Mordasini, C. 2015, A&A, 581, L2. 1508.04970 8The region within the Galaxy where planets can form around stars and retain liquid water on their surface for a significant amount of time. Exoplanets and metallicity 7 Adibekyan, V. Z., Delgado Mena, E., Sousa, S. G., Santos, N. C., Israelian, G., González Hernández, J. I., Mayor, M., & Hakobyan, A. A. 2012a, A&A, 547, A36. 1209.6272 Adibekyan, V. Z., Figueira, P., Santos, N. C., Hakobyan, A. A., Sousa, S. G., Pace, G., Delgado Mena, E., Robin, A. C., Israelian, G., & González Hernández, J. I. 2013a, A&A, 554, A44. 1304.2561 Adibekyan, V. Z., Figueira, P., Santos, N. C., Mortier, A., Mordasini, C., Delgado Mena, E., Sousa, S. G., Correia, A. C. M., Israelian, G., & Oshagh, M. 2013b, A&A, 560, A51. 1311.2417 Adibekyan, V. Z., Santos, N. C., Sousa, S. G., Israelian, G., Delgado Mena, E., González Hernández, J. I., Mayor, M., Lovis, C., & Udry, S. 2012b, A&A, 543, A89. 1205.6670 Adibekyan, V. Z., Sousa, S. G., Santos, N. C., Delgado Mena, E., González Hernández, J. I., Israelian, G., Mayor, M., & Khachatryan, G. 2012c, A&A, 545, A32. 1207.2388 Alibert, Y., Mordasini, C., & Benz, W. 2004, A&A, 417, L25. astro-ph/0403574 Beaugé, C., & Nesvorný, D. 2013, ApJ, 763, 12. 1211.4533 Behroozi, P., & Peeples, M. S. 2015, MNRAS, 454, 1811. 1508.01202 Bitsch, B., Lambrechts, M., & Johansen, A. 2015, A&A, 582, A112. 1507.05209 Bodaghee, A., Santos, N. C., Israelian, G., & Mayor, M. 2003, A&A, 404, 715. astro-ph/ 0304360 Bond, J. C., O'Brien, D. P., & Lauretta, D. S. 2010, Astrophys J, 715, 1050. 1004.0971 Boss, A. P. 1998, ApJ, 503, 923 Brugamyer, E., Dodson-Robinson, S. E., Cochran, W. D., & Sneden, C. 2011, ApJ, 738, 97. 1106.5509 Buchhave, L. A., & Latham, D. W. 2015, ApJ, 808, 187. 1507.03557 Charpinet, S., Fontaine, G., Brassard, P., Green, E. M., Van Grootel, V., Randall, S. K., Silvotti, R., Baran, A. S., Østensen, R. H., Kawaler, S. D., & Telting, J. H. 2011, Nat, 480, 496 Coughlin, J. L., Mullally, F., Thompson, S. E., Rowe, J. F., Burke, C. J., Latham, D. W., Batalha, N. M., Ofir, A., Quarles, B. L., Henze, C. E., Wolfgang, A., Caldwell, D. A., Bryson, S. T., Shporer, A., Catanzarite, J., Akeson, R., Barclay, T., Borucki, W. J., Boyajian, T. S., Campbell, J. R., Christiansen, J. L., Girouard, F. R., Haas, M. R., Howell, S. B., Huber, D., Jenkins, J. M., Li, J., Patil-Sabale, A., Quintana, E. V., Ramirez, S., Seader, S., Smith, J. C., Tenenbaum, P., Twicken, J. D., & Zamudio, K. A. 2016, ApJS, 224, 12. 1512.06149 Courcol, B., Bouchy, F., & Deleuil, M. 2016, MNRAS, 461, 1841. 1604.08560 Dawson, R. I., Chiang, E., & Lee, E. J. 2015, MNRAS, 453, 1471. 1506.06867 Deleuil, M., Deeg, H. J., Alonso, R., Bouchy, F., Rouan, D., Auvergne, M., Baglin, A., Aigrain, S., Almenara, J. M., Barbieri, M., Barge, P., Bruntt, H., Bordé, P., Collier Cameron, A., Csizmadia, S., de La Reza, R., Dvorak, R., Erikson, A., Fridlund, M., Gandolfi, D., Gillon, M., Guenther, E., Guillot, T., Hatzes, A., Hébrard, G., Jorda, L., Lammer, H., Léger, A., Llebaria, A., Loeillet, B., Mayor, M., Mazeh, T., Moutou, C., Ollivier, M., Pätzold, M., Pont, F., Queloz, D., Rauer, H., Schneider, J., Shporer, A., Wuchterl, G., & Zucker, S. 2008, A&A, 491, 889. 0810.0919 Dorn, C., Khan, A., Heng, K., Connolly, J. A. D., Alibert, Y., Benz, W., & Tackley, P. 2015, A&A, 577, A83. 1502.03605 Forgan, D., & Rice, K. 2013, MNRAS, 432, 3168. 1304.4978 Fuhrmann, K. 1998, A&A, 338, 161 Gobat, R., & Hong, S. E. 2016, A&A, 592, A96. 1605.06627 Gonzalez, G. 1997, MNRAS, 285, 403 -- 2009, MNRAS, 399, L103 -- 2014a, ArXiv e-prints. 1403.6761 -- 2014b, MNRAS, 443, 393. 1406.0861 Gowanlock, M. G., Patton, D. R., & McConnell, S. M. 2011, Astrobiology, 11, 855. 1107.1286 Grasset, O., Schneider, J., & Sotin, C. 2009, Astrophys J, 693, 722. 0902.1640 Hasegawa, Y., & Pudritz, R. E. 2013, ApJ, 778, 78. 1310.2009 Haywood, M. 2008, A&A, 482, 673. 0804.2954 Ida, S., & Lin, D. N. C. 2004, ApJ, 604, 388. astro-ph/0312144 8 Vardan Adibekyan Johansen, A., & Lacerda, P. 2010, MNRAS, 404, 475. 0910.1524 Kang, W., Lee, S.-G., & Kim, K.-M. 2011, ApJ, 736, 87. 1105.3083 Kopparapu, R. K., Ramirez, R., Kasting, J. F., Eymet, V., Robinson, T. D., Mahadevan, S., Terrien, R. C., Domagal-Goldman, S., Meadows, V., & Deshpande, R. 2013, ApJ, 765, 131. 1301.6674 Levison, H. F., Kretke, K. A., & Duncan, M. J. 2015, Nat, 524, 322. 1510.02094 Lineweaver, C. H., Fenner, Y., & Gibson, B. K. 2004, Science, 303, 59. astro-ph/0401024 Mayor, M., Marmier, M., Lovis, C., Udry, S., Ségransan, D., Pepe, F., Benz, W., Bertaux, J. ., Bouchy, F., Dumusque, X., Lo Curto, G., Mordasini, C., Queloz, D., & Santos, N. C. 2011, ArXiv e-prints. 1109.2497 Mayor, M., & Queloz, D. 1995, Nat, 378, 355 Miller, N., & Fortney, J. J. 2011, ApJ, 736, L29. 1105.0024 Mordasini, C., Alibert, Y., Benz, W., & Naef, D. 2009, A&A, 501, 1161. 0904.2542 Mordasini, C., Mollière, P., Dittkrist, K.-M., Jin, S., & Alibert, Y. 2015, International Journal of Astrobiology, 14, 201. 1406.5604 Mortier, A., Santos, N. C., Sousa, S., Israelian, G., Mayor, M., & Udry, S. 2013, A&A, 551, Mulders, G. D., Pascucci, I., Apai, D., Frasca, A., & Molenda- Zakowicz, J. 2016, AJ, 152, 187. A112. 1302.1851 1609.05898 Nayakshin, S. 2016, ArXiv e-prints. 1609.07503 O'Toole, S. J., Tinney, C. G., Jones, H. R. A., Butler, R. P., Marcy, G. W., Carter, B., & Bailey, J. 2009, MNRAS, 392, 641. 0810.1589 Pollack, J. B., Hubickyj, O., Bodenheimer, P., Lissauer, J. J., Podolak, M., & Greenzweig, Y. 1996, icarus, 124, 62 Robinson, S. E., Laughlin, G., Bodenheimer, P., & Fischer, D. 2006, ApJ, 643, 484. astro-ph/ 0601656 Santos, N. C., Adibekyan, V., Mordasini, C., Benz, W., Delgado-Mena, E., Dorn, C., Buchhave, L., Figueira, P., Mortier, A., Pepe, F., Santerne, A., Sousa, S. G., & Udry, S. 2015, A&A, 580, L13. 1507.08081 Santos, N. C., Israelian, G., & Mayor, M. 2001, A&A, 373, 1019. astro-ph/0105216 Schwamb, M. E., Orosz, J. A., Carter, J. A., Welsh, W. F., Fischer, D. A., Torres, G., Howard, A. W., Crepp, J. R., Keel, W. C., Lintott, C. J., Kaib, N. A., Terrell, D., Gagliano, R., Jek, K. J., Parrish, M., Smith, A. M., Lynn, S., Simpson, R. J., Giguere, M. J., & Schawinski, K. 2013, ApJ, 768, 127. 1210.3612 Sousa, S. G., Santos, N. C., Israelian, G., Mayor, M., & Udry, S. 2011, A&A, 533, A141. 1108.5279 Suárez-Andrés, L., Israelian, G., González Hernández, J. I., Adibekyan, V. Z., Delgado Mena, E., Santos, N. C., & Sousa, S. G. 2016a, A&A, 591, A69. 1605.03049 -- 2016b, ArXiv e-prints. 1611.10092 Szabó, G. M., & Kiss, L. L. 2011, ApJ, 727, L44. 1012.4791 Thiabaud, A., Marboeuf, U., Alibert, Y., Cabral, N., Leya, I., & Mezger, K. 2014, Astron Astrophys, 562, A27. 1312.3085 Udry, S., Mayor, M., & Santos, N. C. 2003, A&A, 407, 369. astro-ph/0306049 Wang, J., & Fischer, D. A. 2015, AJ, 149, 14 Winn, J. N., & Fabrycky, D. C. 2015, ARA&A, 53, 409. 1410.4199 Zackrisson, E., Calissendorff, P., Gonzalez, J., Benson, A., Johansen, A., & Janson, M. 2016, ArXiv e-prints. 1602.00690 Zhu, W., Wang, J., & Huang, C. 2016, ArXiv e-prints. 1605.04310
1602.02890
1
1602
2016-02-09T08:22:30
Saturn's icy satellites investigated by Cassini - VIMS. IV. Daytime temperature maps
[ "astro-ph.EP" ]
The spectral position of the 3.6 micron continuum peak measured on Cassini-VIMS I/F spectra is used as a marker to infer the temperature of the regolith particles covering the surfaces of Saturn's icy satellites. This feature is characterizing the crystalline water ice spectrum which is the dominant compositional endmember of the satellites' surfaces. Laboratory measurements indicate that the position of the 3.6 micron peak of pure water ice is temperature-dependent, shifting towards shorter wavelengths when the sample is cooled, from about 3.65 micron at T=123 K to about 3.55 micron at T=88 K. A similar method was already applied to VIMS Saturn's rings mosaics to retrieve ring particles temperature (Filacchione et al., 2014). We report here about the daytime temperature variations observed on the icy satellites as derived from three different VIMS observation types. Temperature maps are built by mining the complete VIMS dataset collected in years 2004-2009 (pre-equinox) and in 2009-2012 (post equinox) by selecting pixels with max 150 km/pixel resolution. VIMS-derived temperature maps allow to identify thermal anomalies across the equatorial lens of Mimas and Tethys.
astro-ph.EP
astro-ph
Saturn's icy satellites investigated by Cassini - VIMS. IV. Daytime temperature maps Gianrico Filacchionea,∗, Emiliano D'Aversaa, Fabrizio Capaccionia, Roger N. Clarkb, Dale P. Cruikshankc, Mauro Ciarnielloa, Priscilla Cerronia, Giancarlo Belluccia, Robert H. Brownd, Bonnie J. Burattie, Phillip D. Nicholsonf, Ralf Jaumanng, Thomas B. McCordh, Christophe Sotine, Katrin Stephang, Cristina M. Dalle Orec aINAF-IAPS, Istituto di Astrofisica e Planetologia Spaziali, Area di Ricerca di Tor Vergata, via del Fosso del Cavaliere, 100, 00133, Rome, Italy bPSI Planetary Science Institute, Tucson, AZ, USA cNASA Ames Research Center, Moffett Field, CA 94035-1000, USA dLunar and Planetary Laboratory and Steward Observatory, University of Arizona, Tucson, AZ 85721, USA eJet Propulsion Laboratory, California Institute of Technology, 4800 Oak Groove Drive, Pasadena, CA 91109, USA fCornell University, Astronomy Department, 418 Space Sciences Building, Ithaca, NY 14853, USA gInstitute for Planetary Exploration, DLR, Rutherfordstasse 2, 12489, Berlin, Germany hThe Bear Fight Institute, Winthrop, WA 98862, USA Abstract The spectral position of the 3.6 µm continuum peak measured on Cassini-VIMS I/F spectra is used as a marker to infer the temperature of the regolith particles covering the surfaces of Saturn's icy satellites. This feature is characterizing the crystalline water ice spectrum which is the dominant compositional endmember of the satellites' surfaces. Laboratory measurements indicate that the position of the 3.6 µm peak of pure water ice is temperature-dependent, shifting towards shorter wavelengths when the sample is cooled, from about 3.65 µm at T=123 K to about 3.55 µm at T=88 K. A similar method was already applied to VIMS Saturn's rings mosaics to retrieve ring particles temperature (Filacchione et al., 2014). We report here about the daytime temperature variations observed on the icy satellites as derived from three different VIMS observation types: a) a sample of 240 disk-integrated I/F observations of Saturn's regular satellites collected by VIMS during years 2004-2011 with solar phase in the 20◦-40◦ range, corresponding to late morning-early afternoon local times. This dataset is suitable to exploit the temperature variations at hemispherical scale, resulting in average temperature T<88 K for Mimas, T<<88 K for Enceladus, T<88 K for Tethys, T=98-118 K for Dione, T=108-128 K for Rhea, T=118-128 K for Hyperion, T=128-148 K and T>168 K for Iapetus' trailing and leading hemispheres, respectively. A typical ±5 K uncertainty is associated to the temperature retrieval. On Tethys and Dione, for which observations on both leading and trailing hemispheres are available, in average daytime temperatures higher of about 10 K on the trailing ∗Corresponding author, email [email protected] Preprint submitted to Icarus June 1, 2021 than on the leading hemisphere are inferred. b) Satellites disk-resolved observations taken at 20- 40 km/pixel resolution are suitable to map daytime temperature variations across surfaces' features, such as Enceladus' tiger stripes and Tethys' equatorial dark lens. These datasets allow to disentangle solar illumination conditions from temperature distribution when observing surface's features with strong thermal contrast. c) Daytime average maps covering large regions of the surfaces are used to compare the inferred temperature with geomorphological features (impact craters, chasmatae, equatorial radiation lenses and active areas) and albedo variations. Temperature maps are built by mining the complete VIMS dataset collected in years 2004-2009 (pre-equinox) and in 2009-2012 (post equinox) by selecting pixels with max 150 km/pixel resolution. VIMS-derived temperature maps allow to identify thermal anomalies across the equatorial lens of Mimas and Tethys. A temperature T>115K is measured above Enceladus' Damascus and Alexandria sulci in the south pole region. VIMS has the sensitivity to follow seasonal temperature changes: on Tethys, Dione and Rhea higher temperature are measured above the south hemisphere during pre-equinox and above the north hemisphere during post-equinox epochs. The measured temperature distribution appears correlated with surface albedo features: in fact temperature increases on low albedo units located on Tethys, Dione and Rhea trailing hemispheres. The thermal anomaly region on Rhea's Inktomi crater detected by CIRS (Howett et al., 2014) is confirmed by VIMS: this area appears colder with respect to surrounding terrains when observed at the same local solar time. Keywords: Saturn, Satellites, surfaces, Spectroscopy, Ices 1. Introduction With the aim to retrieve Saturn's icy satellites surface temperature, a method based on the spectral properties at 3.6 µm of water ice at cryogenic temperatures has been developed and applied to the Cassini-VIMS, the Visual and Infrared Mapping Spectrometer, dataset. Considered as a mineral by many authors (Fletcher, 1970; Hobbs, 1974) water ice is capable of many rotovibrational modes: solid state lattice's rotational modes are driven by interactions with nearby molecules while vibrational modes appear at characteristic frequencies given by the molecular structure. Since all these modes are temperature-dependent, in principle is possible to derive the temperature of a water ice sample by analyzing the properties of its spectral reflectance. Water ice lattice undergoes temperature dependent phase-transitions: in vacuum the structure is amorphous for T < 137 K, cubic for 137 ≤ T ≤ 197 K and hexagonal for T > 197 K (Hobbs, 1974). The non-amorphous state is in general referred as crystalline, showing a diagnostic spectral feature at 1.65 µm not present in the amorphous state. Another temperature-induced effect is the variation of the absorption coefficient which decreases when the sample is cooled. As a consequence, cold ice absorbs more photons than warm ice. At the end of the fifties, the first positive infrared identification of water ice on the surfaces of Europa and Ganymede was reported (Kuiper, 1957). With the progress of infrared imaging spectroscopy techniques, advanced spaceborne instrumentation has been developed to study the icy 2 surfaces in the outer solar system: the Near Infrared Mapping Spectrometer (NIMS) on Galileo mission achieved the first compositional maps of the galilean icy satellites (Carlson et al., 1996) while a decade later VIMS, the Visual and Infrared Mapping Spectrometer (Brown et al., 2004) on Cassini has observed in detail the entire population of the icy satellites orbiting around Saturn, giving the first identification of crystalline water ice on the south polar region of the active moon Enceladus (Brown et al., 2006; Jaumann et al., 2008) and the first compositional maps (Jaumann et al., 2006). The analysis of VIMS data indicates that other species, like CO2 and organics correlated with C-H and possibly C-N or adsorbed H2 (Clark et al., 2012), appear mixed with water ice on the surfaces of Hyperion (Cruikshank et al., 2007, 2010), Iapetus (Cruikshank et al., 2008; Clark et al., 2012) and Phoebe (Clark et al., 2005; Buratti et al., 2008; Coradini et al., 2008). Exogenic dark material dispersed on Dione's trailing hemisphere (Clark et al., 2008; Stephan et al., 2010; Scipioni et al., 2013), across Rhea (Stephan et al., 2012; Scipioni et al., 2014) and on Iapetus leading hemisphere (Pinilla-Alonso et al., 2011) was also observed and mapped in detail. Extensive studies have been published about the analysis of disk-integrated I/F spectra of Saturn's satellites and resulting comparative properties, including hemispheric asymmetries observed on the surfaces of the regular satellites (Filacchione et al., 2007, 2010), the retrival of phase and light curves with corresponding photometric parameters (Pitman et al., 2010; Ciarniello et al., 2011) and the radial distribution of water ice and chromophores across Saturn's system of satellites and rings (Filacchione et al., 2012, 2013). All these remotely-sensed observations have boosted an intensive laboratory activity devoted to the characterization of water ice spectral properties. Relevant to the retrieval of water ice temperature was the investigation of the temperature-dependent 1.65 µm absorption feature seen in the reflectance of crystalline water ice (Fink and Larson, 1975). Further analyses, focused on water ice optical constants, have allowed to infer real and imaginary parts of the index of refraction in the VIS and NIR up to 2.8 µm at -7◦C (Warren, 1984), to characterize the water ice absorptions and FWHM as a function of temperature between 20-270 K (Grundy and Schmitt, 1998) and to use the 1.65 µm band as temperature proxy (Grundy et al., 1999; Leto et al., 2005). Taffin et al. (2012) have acquired and characterized water ice sample in the 1-1.8 µm spectral range by varying grains in the 80-700 µm size range and temperature from 80 to 140 K. Optical constants of amorphous and crystalline water ice for 20 ≤ T≤ 150 K in the 1.1-2.6 µm spectral range have been inferred by Mastrapa et al. (2008). A further spectral extension towards longer wavelengths (2.5-22 µm) for amorphous and crystalline forms (Mastrapa et al., 2009) has shown that the reflectance around 3.6 µm is influenced by temperature. This spectral range was never used before to infer the temperature of the ice sample. Located in between the intense absorption at 3.0 µm due to the O-H-O symmetric stretch vibration (ν1) and the weaker 4.6 µm feature generated by a combination of (ν2) and a weak libration mode, the shape of the continuum around 3.6 µm is strongly influenced by the thermal behavior of those two features. The resulting spectral variation caused by temperature changes is described later in section 2. In this work we are aiming to infer surface temperatures of Saturn's icy satellites from infrared observations at different spatial scales (hemispheric, regional and local). While the FP1 thermopile 3 detector radiometer channel of Cassini/CIRS (Flasar et al., 2004) is specifically designed to measure satellites' temperatures through fitting the thermal emission peak in the a 17-1000 µm spectral range, VIMS is a 0.35-5.0 µm imaging spectrometer devoted to infer surface composition (Brown et al., 2004). Despite being limited in spectral range, VIMS offers the advantage of having much better imaging capabilities than CIRS-FP1: with a nominal FOV of 1.83◦ and an IFOV of 500 µrad, VIMS is able to acquire hyperspectral cubes up to 64×64 spatial pixels while CIRS-FP1 channel relies on a single pixel with a circular FOV of 3.9 mrad. Moreover, CIRS mid-infrared Michelson interferometer channel uses two HgCdTe detectors (FP3, FP4), with a 0.273 mrad IFOV, designed to measure icy satellites daytime temperatures. Conversely, since VIMS operates only in reflectance conditions, e.g. on dayside hemispheres, only daytime temperatures can be measured, therefore preventing the possibility to infer nocturnal temperatures and derive accurate thermal inertia data. This task can be achieved only by CIRS. For a given composition, thermal inertia depends upon surface roughness and porosity: as a general rule a grainy water ice pack has high bond albedo and low thermal inertia. On contrary, packed grains show high bond albedo and high thermal inertia. By analyzing CIRS data Howett et al. (2010) have derived an average thermal inertia lower than 20 J m−2 K−1 s−1/2 across the surfaces of the Saturn's satellites which increases up to 89 J m−2 K−1 s−1/2 above thermal anomaly regions of Mimas and Tethys, as discussed later. The properties of the water ice optical constants and resulting I/F spectra as function of the temperature are the topics of the next section 2. In section 3 we discuss how this particular spectral behavior of the ice allows us to use the VIMS instrument as a thermal mapper and to infer the icy satellites daytime temperature. This method is applied in subsection 4 to disk-integrated I/F observations suitable to explore the daytime temperature variability of the regular satellites at hemi- spherical scale and to characterize the differences between their leading and trailing hemispheres. The detection of temperature variations across Enceladus south hemisphere and in Tethys equatorial radiation lens are discussed in section 5. Global temperature maps for Mimas, Enceladus, Tethys, Dione, Rhea, Hyperion and Iapetus during the pre Saturn's equinox (2004-2009) and post equinox (2009-2012) periods are discussed in section 6. 2. Water ice optical constants and temperature-dependent effects Crystalline water ice reflectance is characterized by a broad peak at about 3.6 µm that has temperature-dependent properties. Optical constants measured in transmittance by Mastrapa et al. (2009) in the range of temperature between 20 and 140 K show that in the 3.2-3.6 µm spectral range the real n index is almost constant while the imaginary k index minimum position is temperature- dependent, moving to shorter wavelengths when the temperature decreases (Fig. 1, right panel). The same trend is observed on reflectance measurements performed by Clark et al. (2012) on small grains of pure water ice illuminated at standard conditions (phase=30◦) for temperatures of the sample varying between T=88 K and T=172 K (Fig. 2). The analysis of these data demonstrates that the 3.6 µm reflectance peak shifts towards shorter wavelengths when the ice sample is cooled, moving from 3.675 µm at T=172 K to 3.58 µm at T=88 K. This means a shift of about 90 nm 4 Figure 1: Water ice optical constants at temperature between T=20 to 140 K by Mastrapa et al. (2009). Left panel: real index n. Right panel: imaginary index k. Note how the k index minimum shifts towards shorter wavelengths with temperature decrease. corresponding to about 6 VIMS spectral bands. 3. VIMS as a thermal mapper Starting from the experimental evidences discussed in the previous section, a 4th-degree poly- nomial fit in the 3.2-3.8 µm range has been applied to laboratory measurements shown in Fig. 2 to determine the wavelength where the peak occurs with the view toward using it as a marker to retrieve similar temperatures on the regular icy satellites of Saturn. While band depths and I/F level are both grain size-dependent, the position of the 3.6 µm peak is insensitive to grain size and illumination geometry: this effect is well shown on synthetic spectra of pure water ice grains of 10-100 µm diameters discussed in Filacchione et al. (2012), Fig. 5. Moreover, the 3.6 µm peak position is not affected by illumination geometry as shown on synthetic VIS-IR reflectance spectra of pure water ice grains of 30 µm diameter simulated at fixed phase angle g=30◦ and incidence angles i=15◦, 25◦, 35◦, 45◦, 55◦, 65◦, 75◦ (Fig. 3, left panel) and at fixed incidence angle i=80◦ at phase g=0◦, 20◦, 40◦, 60◦, 80◦, 120◦, 140◦ (Fig. 3, right panel). In both cases the synthetic spectra are simulated following the Monte Carlo ray-tracing formalism discussed in Ciarniello et al. (2014) by using water ice optical constants at T=90 K measured by Warren (1984); Mastrapa et al. (2009); Clark et al. (2012). At fixed phase, changes of the incidence angle have small effects on the continuum level and on band 5 Figure 2: Water ice reflectance measured at temperature 88 ≤ T ≤ 172 K. The best-fit wavelength of the 3.6 µm peak is reported for each temperature. The peak position at T=88 K and T 172 K is indicated by vertical black and red dashed lines, respectively. depths. Conversely, fixing the incidence angle and varying the phase, large variations are induced: in fact when the phase angle increases the 1.5, 2.0 and 2.9 µm absorptions become larger while the continuum level decreases. In both cases, however, the position of the 3.6 µm is not changing being it only dependent from temperature through the imaginary index of the optical constant, as previously shown in Fig. 1. The same fitting technique used for laboratory data has been applied to VIMS I/F data. The retrieval of the temperature is therefore performed through the comparison of the 3.6 µm peak wavelength between VIMS and laboratory data. The position of the peak has been derived by applying a 4th degree polynomial fit to VIMS I/F data in the 3.2-3.8 µm range and then determining the position of the peak on the fitted curve. A sampling equal to 1/3 of VIMS band width has been used to determine the peak wavelength, corresponding to a temperature's error of about ±5 K to be associated to all results reported hereafter. Data filtering, as discussed later in section 6, allow us 6 Figure 3: Water ice synthetic normalized reflectances of pure water ice grains of 30 µm diameter at fixed phase angle g=30◦ and incidence angles i=15◦, 25◦, 35◦, 45◦, 55◦, 65◦, 75◦ (left panel) and at fixed incidence angle i=80◦ at phase g=0◦, 20◦, 40◦, 60◦, 80◦, 120◦, 140◦ (right panel). Simulation performed with the Monte Carlo ray-tracing formalism discussed in Ciarniello et al. (2014) by using water ice optical constants at T=90 K (Warren, 1984; Mastrapa et al., 2009; Clark et al., 2012). to apply this method on pixels having a typical SNR>25 at 3.6 µm. The error resulting from the presence of other compositional endmembers is not included. As a general rule, the spectral position and shape of the 3.6 µm peak can be altered only in case of water ice mixed with materials having absorption features in this spectral range. As an example, tholins (Imanaka et al., 2012) and organic materials are characterized by a broad absorption feature in the 3.2-3.6 µm range which could affect the peak properties and therefore the retrieval of the temperature. Conversely, amorphous carbon (Zubko et al., 1996) is featureless at these wavelengths: while small amount of carbon mixed within water ice reduces significantly the reflectance level, being spectrally neutral it is not able to alter the peak position. The same temperature retrieval method has been already applied to several VIMS mosaics of Saturn's ring and the results were compared with CIRS observations occurring at the same time (Filacchione et al., 2014). In general the temperatures measured by VIMS with this method are higher than corresponding ones reported by CIRS. This is a consequence of the shallow layer (few microns) to which VIMS is sensitive compared to CIRS which measures temperature at greater depth, from a few millimeters to a centimeter. Since in this work data taken around noon time have been selected, the shallower skin depth sampled by VIMS could be the reason of the higher temperatures inferred. Moreover, both instruments are influenced by surface roughness and sub-pixel shadows but the discussion of these effects goes well beyond the scopes of this work. Also grain size and amount of contaminants embedded in water ice play a role in the 3.6 µm peak's properties and these effects 7 have yet to be investigated. Certainly, combining VIMS and CIRS measurements together will allow us to better understand the regolith's physical properties and heat transport mechanisms. The outer layer in fact is warmer than the material below it when measured around noon as the surface receive the maximum of energy that the subsurface will later get by conduction. It is noteworthy that the method based on the 3.6 µm peak fit greatly differs with respect to the derivation of the temperature from thermal infrared radiance fitting with a blackbody law. In the first case water ice is used as a proxy and spectral properties measured in laboratory conditions are used to infer the temperature of the satellites' surfaces. The major advantages of this method are the wide excursion of the peak as function of temperature, which allows to derive its spectral position with great precision and the good signal to noise ratio in the 3.6 µm range in VIMS spectra. The drawback is the effect of contaminants in changing the shape and position of the water ice feature which is difficult to characterize without specific laboratory measurements. In the second case the temperature is derived by fitting radiance in the thermal range with a blackbody function. While this second method is in principle more robust and working well for warm objects orbiting close to the Sun (Coradini et al., 2011), it can be scarcely used to derive icy satellites temperature from VIMS data because the thermal emission is lower than the solar contribution in the 3-5 µm range and signal to noise is poor. Up to the present time, only Goguen et al. (2013) have successfully used the blackbody fit to derive Baghdad Sulcus color temperature Tc=197±20 K from VIMS high spatial resolution data acquired above Enceladus' south pole region during the 14th April 2012 South Pole Cassini's flyby. In the next sections we discuss about the results obtained by applying the 3.6 µm fit method to three different VIMS datasets: 1. a collection of 240 disk-integrated observations of Saturn's regular satellites collected by VIMS between 2004 and 2011 with solar phase in the 20◦-40◦ range, corresponding to late morning- early afternoon local times is described in section 4. This dataset is suitable to analyze average surface temperature at global scale and to retrieve temperature variations across the leading and trailing hemispheres. 2. Disk-resolved observations (20-40 km/pixel) of Enceladus and Tethys, discussed in section 5, which are suitable to map temperature variations across geomorphological regional features, such as Enceladus' tiger stripes or Tethys equatorial dark lens. 3. Global temperature maps are derived for Mimas, Enceladus, Tethys, Dione, Rhea, Hyperion and Iapetus by using the entire VIMS geo-referenced dataset. These results are given in section 6. 4. Icy satellites disk-integrated temperatures. A first temperature retrieval for all regular satellites from Mimas to Iapetus is carried out on the dataset listed in Tables 1-7 which includes observations taken with solar phase in the 20◦-40◦ range, e.g., at late morning-early afternoon conditions, selected from the dataset already discussed in Filacchione et al. (2012). The filtering on phase is imposed in order to match the laboratory 30◦ phase illumination condition (Clark et al., 2012) for which the 3.6 µm peak position is calibrated. 8 SEQUENCE OBS ID S25-V1542757940 S25-V1542757954 S25-V1542757969 S25-V1542757998 S25-V1542758013 S25-V1542758042 S25-V1542758057 S25-V1542758071 S25-V1542758086 S25-V1542758101 S25-V1542758115 S25-V1542758130 S25-V1542758144 S25-V1542758159 S25-V1542758174 S25-V1542758188 S30-V1558932227 S30-V1558941481 S30-V1558941534 S30-V1558941566 S30-V1558941605 S30-V1558941644 S30-V1558941692 S30-V1558941731 S30-V1558941770 S30-V1558949456 S30-V1558949482 S30-V1558949509 S30-V1558949535 S30-V1558949561 S57-V1644780761 S66-V1675134165 S66-V1675134191 S66-V1675135681 S66-V1675135707 S66-V1675135741 S66-V1675135767 S66-V1675137246 S66-V1675137272 S66-V1675137332 S66-V1675138026 S66-V1675138052 START-END TIME (UT) 2006-324T23:19:59-23:20:14 2006-324T23:20:14-23:20:28 2006-324T23:20:29-23:20:43 2006-324T23:20:58-23:21:12 2006-324T23:21:12-23:21:27 2006-324T23:21:42-23:21:56 2006-324T23:21:56-23:22:11 2006-324T23:22:11-23:22:25 2006-324T23:22:26-23:22:40 2006-324T23:22:40-23:22:55 2006-324T23:22:55-23:23:09 2006-324T23:23:09-23:23:24 2006-324T23:23:24-23:23:38 2006-324T23:23:39-23:23:53 2006-324T23:23:53-23:24:08 2006-324T23:24:08-23:24:22 2007-147T04:09:45-04:09:59 2007-147T06:43:58-06:44:24 2007-147T06:44:51-06:45:16 2007-147T06:45:23-06:46:02 2007-147T06:46:03-06:46:41 2007-147T06:46:42-06:47:20 2007-147T06:47:29-06:48:08 2007-147T06:48:09-06:48:47 2007-147T06:48:48-06:49:26 2007-147T08:56:53-08:57:19 2007-147T08:57:20-08:57:45 2007-147T08:57:46-08:58:11 2007-147T08:58:12-08:58:38 2007-147T08:58:38-08:59:04 2010-044T18:48:38-18:49:58 2011-031T02:15:11-02:15:37 2011-031T02:15:38-02:16:03 2011-031T02:40:27-02:40:53 2011-031T02:40:54-02:41:19 2011-031T02:41:27-02:41:53 2011-031T02:41:54-02:42:19 2011-031T03:06:32-03:06:58 2011-031T03:06:59-03:07:24 2011-031T03:07:59-03:08:24 2011-031T03:19:32-03:19:58 2011-031T03:19:59-03:20:24 S 12 12 12 12 12 12 12 12 12 12 12 12 12 12 12 12 12 24 24 24 24 24 24 24 24 24 24 24 24 24 30 24 24 24 24 24 24 24 24 24 24 24 L 12 12 12 12 12 12 12 12 12 12 12 12 12 12 12 12 12 12 12 12 12 12 12 12 12 12 12 12 12 12 30 12 12 12 12 12 12 12 12 12 12 12 EXP (s) 80 80 80 80 80 80 80 80 80 80 80 80 80 80 80 80 80 80 80 120 120 120 120 120 120 80 80 80 80 80 80 80 80 80 80 80 80 80 80 80 80 80 PHASE (◦) 38.7 38.7 38.7 38.4 38.4 38.4 38.1 38.1 38.1 38.1 37.8 37.8 37.8 37.8 37.5 37.5 24.4 26.0 26.0 26.0 26.0 26.0 26.0 26.0 26.0 25.0 25.0 25.0 25.0 25.0 27.1 39.0 38.7 31.9 31.6 31.6 31.4 25.3 25.0 24.8 22.3 22.0 SUBSPC LON (◦) 225.1 225.1 225.1 225.3 225.3 225.3 225.5 225.5 225.5 225.5 225.7 225.7 225.7 225.7 225.9 225.9 72.5 109.9 110.2 110.6 110.9 111.2 111.5 111.8 112.2 155.5 155.8 156.2 156.5 156.9 117.5 84.1 84.0 83.3 83.3 83.3 83.2 83.0 83.0 83.0 83.1 83.1 SUBSPC LAT (◦) -47.4 -47.4 -47.4 -47.1 -47.1 -47.1 -46.8 -46.8 -46.8 -46.8 -46.5 -46.5 -46.5 -46.5 -46.1 -46.1 -34.1 -35.8 -35.8 -35.8 -35.8 -35.8 -35.8 -35.7 -35.7 -31.9 -31.8 -31.8 -31.7 -31.7 -10.6 -2.5 -2.5 -2.6 -2.6 -2.6 -2.6 -2.8 -2.8 -2.8 -2.8 -2.8 SSOL LON (◦) 206.1 206.1 206.1 206.3 206.3 206.3 206.6 206.6 206.6 206.6 206.8 206.8 206.8 206.8 207.1 207.1 67.1 113.3 113.6 113.9 114.2 114.4 114.7 115.0 115.3 149.1 149.3 149.6 149.8 150.1 141.7 46.2 46.4 52.8 53.1 53.1 53.3 59.7 60.0 60.2 63.1 63.4 SSOL LAT (◦) -13.1 -13.1 -13.1 -13.1 -13.1 -13.1 -13.1 -13.1 -13.1 -13.1 -13.1 -13.1 -13.1 -13.1 -13.2 -13.2 -10.9 -10.9 -10.9 -10.9 -10.9 -10.9 -10.9 -10.9 -10.8 -10.1 -10.1 -10.1 -10.1 -10.1 1.8 7.0 7.0 7.0 7.0 7.0 7.0 7.0 7.0 7.0 7.0 7.0 SPC ALT (km) 152554. 152554. 152554. 152444. 152444. 152444. 152339. 152339. 152339. 152339. 152238. 152238. 152238. 152238. 152143. 152143. 207129. 224918. 224968. 225018. 225068. 225119. 225169. 225219. 225269. 238232. 238445. 238662. 238880. 239102. 30537. 139223. 139266. 140809. 140892. 140892. 140977. 143351. 143462. 143574. 144851. 144971. Table 1: Mimas disk-integrated observations processed in this work. Columns S and L indicate the cube dimensions along sample and lines axis, respectively. By fitting the position of the 3.6 µm peak on those disk-integrated I/F spectra we have measured average daytime temperatures of T=88 K for Mimas, T<<88 K for Enceladus, T<88 K for Tethys, T=100 K for Dione, T=108 K for Rhea, T=113 K for Hyperion, T=138K and T>168K for Iapetus trailing and leading hemispheres, respectively. The distribution of the temperature as a function of the subspacecraft longitude on the 240 observations processed is shown in Fig. 4. Three different trends are visible in the scatterplot: 1) constant temperature is observed independently from the observed hemisphere on Mimas, Enceladus and Hyperion; 2) On Tethys, Dione and Rhea, for which observations taken on both leading and trailing hemispheres are available, temperatures are higher by about 10 K on the trailing hemisphere than on the leading hemisphere. In this case higher temperature is correlated with the prevalence of lower albedo terrain units on the trailing hemispheres of these moons; 3) On Iapetus the higher temperature is observed on the dark leading than on the bright trailing hemisphere. The lack of laboratory measurements for T< 88 K and T> 168 K allows us only to give upper limits to Mimas, Enceladus, Tethys temperatures and a lower limit to Iapetus leading hemisphere temperature, respectively. 9 SEQUENCE OBS ID S17-V1516162569 S17-V1516162727 S17-V1516162803 S17-V1516163887 S17-V1516164030 S17-V1516165299 S17-V1516165442 S17-V1516166711 S17-V1516166854 S17-V1516168270 S17-V1516168413 S17-V1516168802 S17-V1516169654 S17-V1516170506 S17-V1516171358 S55-V1637469704 S60-V1652861972 S60-V1652871322 S60-V1652871418 S60-V1652874331 S60-V1652874379 S60-V1652874427 S60-V1652874475 S70-V1696181352 S70-V1696181623 START-END TIME (UT) 2006-017T03:46:39-03:47:46 2006-017T03:49:17-03:50:24 2006-017T03:50:33-03:51:40 2006-017T04:08:37-04:10:48 2006-017T04:11:00-04:13:11 2006-017T04:32:09-04:34:20 2006-017T04:34:32-04:36:43 2006-017T04:55:40-04:57:52 2006-017T04:58:03-05:00:15 2006-017T05:21:39-05:23:51 2006-017T05:24:02-05:26:14 2006-017T05:30:31-05:31:59 2006-017T05:44:43-05:46:11 2006-017T05:58:55-06:00:23 2006-017T06:13:07-06:14:35 2009-325T03:58:33-03:59:21 2010-138T07:34:32-07:35:31 2010-138T10:10:22-10:11:01 2010-138T10:11:58-10:12:37 2010-138T11:00:31-11:01:10 2010-138T11:01:19-11:01:58 2010-138T11:02:07-11:02:46 2010-138T11:02:55-11:03:34 2011-274T16:39:25-16:40:22 2011-274T16:43:56-16:44:53 S 32 32 32 48 48 48 48 48 48 48 48 36 36 36 36 24 30 30 30 30 30 30 30 36 36 L 24 24 24 32 32 32 32 32 32 32 32 28 28 28 28 22 30 15 15 15 15 15 15 18 18 EXP (s) 80 80 80 80 80 80 80 80 80 80 80 80 80 80 80 80 60 80 80 80 80 80 80 80 80 PHASE (◦) 23.7 23.8 23.8 24.4 24.5 25.4 25.5 26.4 26.8 28.1 28.3 28.7 29.8 30.9 32.2 35.3 28.0 34.4 34.5 36.7 36.7 36.8 36.8 24.6 24.4 SUBSPC LON (◦) 226.3 226.8 227.0 229.4 229.8 232.4 232.6 234.8 235.4 237.7 237.9 238.5 239.8 240.9 242.1 38.9 48.0 69.7 70.0 76.5 76.6 76.7 76.9 50.3 50.8 SUBSPC LAT (◦) -0.3 -0.3 -0.3 -0.3 -0.3 -0.2 -0.2 -0.2 -0.1 -0.1 -0.1 -0.1 -0.1 -0.0 -0.0 -1.9 -1.0 -0.5 -0.5 -0.4 -0.4 -0.4 -0.4 -0.5 -0.5 SSOL LON (◦) 241.7 242.4 242.6 246.1 246.7 250.7 251.1 254.8 255.9 260.2 260.8 262.0 264.7 267.5 270.4 74.6 75.5 103.8 104.2 112.9 113.1 113.3 113.5 28.6 29.3 SSOL LAT (◦) -18.3 -18.3 -18.3 -18.3 -18.3 -18.4 -18.4 -18.4 -18.5 -18.5 -18.5 -18.5 -18.5 -18.5 -18.5 1.6 4.2 4.2 4.2 4.2 4.2 4.2 4.2 11.3 11.3 SPC ALT (km) 148814. 148805. 148805. 148929. 148973. 149454. 149515. 150235. 150512. 151732. 151923. 152323. 153352. 154515. 155913. 51731. 35699. 94938. 95611. 110604. 110891. 111177. 111463. 77778. 79700. Table 2: Enceladus disk-integrated observations processed in this work. Columns S and L indicate the cube dimensions along sample and lines axis, respectively. SEQUENCE OBS ID S09-V1489020423 S09-V1489020590 S09-V1489020757 S22-V1532362574 S30-V1558913033 S30-V1558913501 S30-V1558913605 S36-V1577490717 S36-V1577491678 S36-V1577492448 S36-V1577494179 S36-V1577495140 S36-V1577495910 S36-V1577496871 S36-V1577497641 S36-V1577498602 S36-V1577499372 S36-V1577500333 S36-V1577501103 S36-V1577502064 S36-V1577502834 S36-V1577503795 S36-V1577504565 S36-V1577505526 S36-V1577506296 S40-V1587465454 S40-V1587465468 S40-V1587465844 S40-V1587465858 S40-V1587466234 S40-V1587466248 S40-V1587466624 S40-V1587466638 S40-V1587467014 S40-V1587467028 S54-V1634216514 S54-V1634216781 S54-V1634217726 START-END TIME (UT) 2005-068T00:20:27-00:22:55 2005-068T00:23:14-00:25:42 2005-068T00:26:01-00:28:29 2006-204T15:45:00-15:45:52 2007-146T22:49:51-22:50:30 2007-146T22:57:39-22:58:56 2007-146T22:59:23-22:59:55 2007-361T23:15:52-23:18:52 2007-361T23:31:53-23:34:53 2007-361T23:44:43-23:47:43 2007-362T00:13:34-00:16:34 2007-362T00:29:35-00:32:35 2007-362T00:42:25-00:45:25 2007-362T00:58:26-01:01:26 2007-362T01:11:16-01:14:16 2007-362T01:27:17-01:30:17 2007-362T01:40:07-01:43:07 2007-362T01:56:08-01:59:08 2007-362T02:08:58-02:11:58 2007-362T02:24:59-02:27:59 2007-362T02:37:49-02:40:49 2007-362T02:53:50-02:56:50 2007-362T03:06:40-03:09:40 2007-362T03:22:41-03:25:41 2007-362T03:35:31-03:38:31 2008-112T10:00:19-10:00:31 2008-112T10:00:32-10:00:45 2008-112T10:06:49-10:07:01 2008-112T10:07:02-10:07:15 2008-112T10:13:19-10:13:31 2008-112T10:13:32-10:13:45 2008-112T10:19:49-10:20:01 2008-112T10:20:02-10:20:15 2008-112T10:26:19-10:26:31 2008-112T10:26:32-10:26:45 2009-287T12:19:07-12:19:28 2009-287T12:23:34-12:23:55 2009-287T12:39:19-12:39:40 S 48 48 48 24 24 24 24 32 32 32 32 32 32 32 32 32 32 32 32 32 32 32 32 32 32 24 24 24 24 24 24 24 24 24 24 42 42 42 L 36 36 36 24 24 24 24 32 32 32 32 32 32 32 32 32 32 32 32 32 32 32 32 32 32 12 12 12 12 12 12 12 12 12 12 22 22 22 EXP (s) 80 80 80 80 60 120 50 160 160 160 160 160 160 160 160 160 160 160 160 160 160 160 160 160 160 40 40 40 40 40 40 40 40 40 40 20 20 20 PHASE (◦) 23.5 23.6 23.7 24.3 39.1 37.3 36.9 35.8 35.9 35.9 36.0 36.0 36.1 36.2 36.2 36.3 36.4 36.5 36.5 36.6 36.7 36.8 36.9 37.1 37.2 36.4 36.4 36.1 36.1 35.8 35.8 35.5 35.5 35.3 35.2 27.6 26.5 22.9 SUBSPC LON (◦) 214.9 215.5 216.3 107.4 38.8 38.7 38.7 303.2 305.4 307.0 310.5 312.4 313.9 315.8 317.3 319.2 320.7 322.5 324.1 325.7 327.4 329.1 330.5 332.4 333.9 32.1 32.2 32.5 32.5 32.8 32.9 33.2 33.3 33.6 33.6 248.2 247.7 246.1 SUBSPC LAT (◦) 1.3 1.3 1.3 -37.5 2.2 0.4 0.0 21.3 21.2 21.1 21.0 20.9 20.8 20.8 20.7 20.6 20.6 20.5 20.5 20.5 20.4 20.4 20.3 20.3 20.3 -38.2 -38.2 -38.1 -38.1 -38.0 -38.0 -38.0 -38.0 -37.9 -37.9 2.5 2.4 2.3 SSOL LON (◦) 208.6 209.0 209.5 93.6 3.3 4.3 4.6 320.4 322.7 324.4 328.2 330.3 332.0 334.1 335.8 337.9 339.6 341.5 343.3 345.1 347.1 349.0 350.6 352.7 354.4 11.3 11.4 12.0 12.2 12.9 13.1 13.8 14.0 14.6 14.7 220.2 220.8 222.9 SSOL LAT (◦) -20.7 -20.7 -20.7 -16.4 -12.8 -12.8 -12.8 -9.8 -9.8 -9.8 -9.7 -9.7 -9.7 -9.7 -9.7 -9.7 -9.7 -9.7 -9.7 -9.7 -9.7 -9.7 -9.7 -9.7 -9.7 -7.7 -7.7 -7.7 -7.7 -7.7 -7.7 -7.7 -7.7 -7.7 -7.7 2.1 2.1 2.1 SPC ALT (km) 214413. 213687. 212736. 130244. 127606. 131768. 132841. 2307240. 2316120. 2322380. 2335900. 2343070. 2348740. 2355520. 2360860. 2367210. 2372200. 2377760. 2382740. 2387530. 2392440. 2397110. 2406700. 2405450. 2408900. 572815. 573143. 574770. 575093. 577005. 577321. 579190. 579498. 581022. 581324. 123558. 125888. 134467. Table 3: Tethys disk-integrated observations processed in this work. Columns S and L indicate the cube dimensions along sample and lines axis, respectively. 10 SEQUENCE OBS ID S13-V1501609705 S13-V1501605232 S15-V1507717998 S15-V1507718096 S15-V1507718735 S15-V1507718819 S15-V1507700309 S15-V1507700483 S15-V1507700542 S15-V1507733247 S15-V1507735292 S15-V1507703774 S15-V1507704605 S15-V1507706267 S15-V1507707228 S15-V1507708306 S15-V1507701040 S15-V1507738123 S17-V1514076858 S22-V1532405053 S22-V1532405086 S22-V1532405149 S22-V1532405609 S34-V1569831278 S34-V1569853207 S34-V1569853808 S34-V1569854409 S34-V1569854506 S34-V1569854700 S70-V1696219767 S70-V1696219855 S70-V1696219943 START-END TIME (UT) 2005-213T17:20:29-17:23:38 2005-213T16:05:56-16:07:39 2005-284T10:04:43-10:06:12 2005-284T10:06:21-10:07:50 2005-284T10:17:00-10:18:15 2005-284T10:18:24-10:19:39 2005-284T05:09:54-05:10:45 2005-284T05:12:48-05:13:39 2005-284T05:13:47-05:14:38 2005-284T14:18:52-14:20:44 2005-284T14:52:56-14:54:33 2005-284T06:07:39-06:08:17 2005-284T06:21:30-06:22:08 2005-284T06:49:12-06:49:50 2005-284T07:05:13-07:06:04 2005-284T07:23:11-07:24:02 2005-284T05:22:05-05:22:43 2005-284T15:40:07-15:40:41 2005-358T00:25:02-00:26:26 2006-205T03:32:59-03:33:25 2006-205T03:33:32-03:34:24 2006-205T03:34:35-03:35:01 2006-205T03:42:15-03:42:41 2007-273T07:39:27-07:42:44 2007-273T13:44:55-13:46:20 2007-273T13:54:56-13:56:21 2007-273T14:04:57-14:06:22 2007-273T14:06:34-14:07:59 2007-273T14:09:48-14:11:13 2011-275T03:19:40-03:20:56 2011-275T03:21:08-03:22:24 2011-275T03:22:36-03:23:52 S 42 24 32 32 36 36 24 24 24 36 40 24 24 24 24 24 24 32 36 24 24 24 24 48 36 36 36 36 36 34 34 34 L 30 24 32 32 24 24 24 24 24 36 28 18 18 18 24 24 18 32 18 12 12 12 12 48 18 18 18 18 18 17 17 17 EXP (s) 140 160 80 80 80 80 80 80 80 80 80 80 80 80 80 80 80 30 120 80 160 80 80 80 120 120 120 120 120 120 120 120 PHASE (◦) 39.9 37.4 21.9 21.9 22.0 22.0 20.2 20.2 20.2 22.8 22.8 20.5 20.6 20.7 20.8 20.9 20.3 22.7 21.2 30.1 30.0 29.9 29.6 32.1 23.1 23.0 22.9 22.9 22.8 28.8 28.8 28.7 SUBSPC LON (◦) 251.8 240.3 137.2 137.4 138.2 138.3 115.7 115.9 115.9 157.9 160.9 119.8 120.7 122.7 123.9 125.2 116.5 165.2 95.1 299.5 299.4 299.3 299.0 334.3 47.8 48.7 49.5 49.7 50.0 102.7 102.8 103.0 SUBSPC LAT (◦) -41.1 -43.0 -0.5 -0.5 -0.5 -0.5 -0.3 -0.3 -0.3 -0.8 -0.9 -0.4 -0.4 -0.4 -0.4 -0.4 -0.3 -1.1 -0.4 6.9 6.9 7.0 7.2 4.7 9.4 9.3 9.2 9.2 9.2 0.0 0.0 0.0 SSOL LON (◦) 210.9 204.0 147.3 147.5 148.5 148.6 120.4 120.7 120.8 170.5 173.6 125.7 127.0 129.4 131.0 132.6 121.5 177.9 105.3 280.4 280.5 280.6 281.0 2.7 35.9 36.8 37.7 37.9 38.3 129.4 129.5 129.6 SSOL LAT (◦) -20.8 -20.8 -20.1 -20.1 -20.1 -20.1 -20.1 -20.1 -20.1 -20.1 -20.1 -20.1 -20.1 -20.1 -20.1 -20.1 -20.1 -20.1 -19.2 -16.5 -16.5 -16.5 -16.5 -10.4 -10.4 -10.4 -10.4 -10.4 -10.4 11.4 11.4 11.4 SPC ALT (km) 273999. 287995. 281504. 280126. 272575. 271890. 500654. 498285. 497496. 119191. 99308. 455298. 444499. 423845. 410961. 397423. 491190. 72415. 597903. 261515. 261452. 261390. 261164. 63758. 183882. 187928. 191945. 192744. 194339. 199816. 199916. 200017. Table 4: Dione disk-integrated observations processed in this work. Columns S and L indicate the cube dimensions along sample and lines axis, respectively. SEQUENCE OBS ID S17-V1516202693 S17-V1516202765 S43-V1597403216 S43-V1597403307 S43-V1597404177 S43-V1597405905 S43-V1597405996 S43-V1597406099 S43-V1597407060 S43-V1597407151 S43-V1597407254 S43-V1597407345 S43-V1597407448 S43-V1597407539 S43-V1597407642 S43-V1597407688 S43-V1597407740 S43-V1597407786 S43-V1597407838 S43-V1597407884 S43-V1597407936 S43-V1597407982 S43-V1597408034 S43-V1597833610 S43-V1597833658 S43-V1597833706 S43-V1597833802 S43-V1597833850 S43-V1597833898 S43-V1597833946 S43-V1597833994 S43-V1597834042 S72-V1710091283 START-END TIME (UT) 2006-017T14:55:22-14:56:28 2006-017T14:56:34-14:57:40 2008-227T10:28:30-10:30:00 2008-227T10:30:01-10:31:30 2008-227T10:44:31-10:46:01 2008-227T11:13:19-11:14:49 2008-227T11:14:50-11:16:19 2008-227T11:16:33-11:18:03 2008-227T11:32:34-11:34:04 2008-227T11:34:05-11:35:34 2008-227T11:35:48-11:37:18 2008-227T11:37:19-11:38:48 2008-227T11:39:02-11:40:32 2008-227T11:40:33-11:42:02 2008-227T11:42:16-11:43:01 2008-227T11:43:02-11:43:46 2008-227T11:43:54-11:44:39 2008-227T11:44:40-11:45:24 2008-227T11:45:32-11:46:17 2008-227T11:46:18-11:47:02 2008-227T11:47:10-11:47:55 2008-227T11:47:56-11:48:40 2008-227T11:48:48-11:49:33 2008-232T10:01:41-10:02:20 2008-232T10:02:29-10:03:08 2008-232T10:03:17-10:03:56 2008-232T10:04:53-10:05:32 2008-232T10:05:41-10:06:20 2008-232T10:06:29-10:07:08 2008-232T10:07:17-10:07:56 2008-232T10:08:05-10:08:44 2008-232T10:08:53-10:09:32 2012-070T16:30:08-16:32:13 S 42 42 32 32 32 32 32 32 32 32 32 32 32 32 32 32 32 32 32 32 32 32 32 24 24 24 24 24 24 24 24 24 54 L 36 36 16 16 16 16 16 16 16 16 16 16 16 16 16 16 16 16 16 16 16 16 16 12 12 12 12 12 12 12 12 12 54 EXP (s) 40 40 160 160 160 160 160 160 160 160 160 160 160 160 80 80 80 80 80 80 80 80 80 120 120 120 120 120 120 120 120 120 40 PHASE (◦) 34.7 34.8 27.7 27.7 27.4 26.9 26.9 26.8 26.5 26.5 26.5 26.4 26.4 26.4 26.3 26.3 26.3 26.3 26.3 26.3 26.3 26.2 26.2 26.5 26.6 26.6 26.7 26.7 26.7 26.7 26.8 26.8 32.1 SUBSPC LON (◦) 309.1 309.0 348.9 348.9 349.4 350.4 350.5 350.5 351.1 351.2 351.2 351.3 351.3 351.4 351.4 351.5 351.5 351.5 351.6 351.6 351.6 351.6 351.7 5.9 5.9 5.9 6.0 6.0 6.0 6.0 6.0 6.1 148.2 SUBSPC LAT (◦) 0.3 0.3 -4.1 -4.1 -3.9 -3.7 -3.7 -3.7 -3.6 -3.6 -3.6 -3.6 -3.6 -3.6 -3.6 -3.5 -3.5 -3.5 -3.5 -3.5 -3.5 -3.5 -3.5 -30.6 -30.7 -30.7 -30.8 -30.8 -30.8 -30.9 -30.9 -30.9 -1.4 SSOL LON (◦) 338.7 338.8 321.2 321.3 322.1 323.7 323.8 323.9 324.8 324.8 324.9 325.0 325.1 325.2 325.3 325.3 325.4 325.4 325.5 325.5 325.5 325.6 325.6 357.7 357.7 357.8 357.9 357.9 358.0 358.0 358.0 358.1 177.0 SSOL LAT (◦) -18.5 -18.5 -5.3 -5.3 -5.3 -5.3 -5.3 -5.3 -5.3 -5.3 -5.3 -5.3 -5.3 -5.3 -5.3 -5.3 -5.3 -5.3 -5.3 -5.3 -5.3 -5.3 -5.3 -5.2 -5.2 -5.2 -5.2 -5.2 -5.2 -5.2 -5.2 -5.2 13.2 SPC ALT (km) 226163. 225845. 1638460. 1638800. 1642060. 1648300. 1648620. 1648990. 1652310. 1652610. 1652970. 1653270. 1653620. 1653920. 1654200. 1654350. 1654530. 1654680. 1654850. 1655000. 1655180. 1655330. 1655500. 427174. 427783. 428391. 429609. 430217. 430826. 431435. 432044. 432653. 58652. Table 5: Rhea disk-integrated observations processed in this work. Columns S and L indicate the cube dimensions along sample and lines axis, respectively. 11 SEQUENCE OBS ID S11-V1497069822 S11-V1497069911 S11-V1497070000 S11-V1497070437 S11-V1497070874 S11-V1497070963 S11-V1497071546 S11-V1497072041 S11-V1497072478 S11-V1497088451 S11-V1497089002 S11-V1497089933 S11-V1497090342 S11-V1497090451 S11-V1497090986 S11-V1497129440 S11-V1497130881 S11-V1497131666 S11-V1497140531 S11-V1497140620 S11-V1497141146 S11-V1497142550 S17-V1513877057 S17-V1513911329 S21-V1530174536 S21-V1530174748 S21-V1530176158 S21-V1530176823 S21-V1530178153 S21-V1530178818 S21-V1530179483 S21-V1530179598 S21-V1530179713 S21-V1530186404 S65-V1669622273 S65-V1669622327 S65-V1669622390 S65-V1669622444 S65-V1669622507 S65-V1669622615 S65-V1669622737 S65-V1669622845 S65-V1669626673 S65-V1669626736 S65-V1669626790 S65-V1669626853 S65-V1669626961 S65-V1669627191 S69-V1692997197 S69-V1692997353 S69-V1692997510 S69-V1692997590 START-END TIME (UT) 2005-161T04:16:15-04:17:32 2005-161T04:17:44-04:19:01 2005-161T04:19:13-04:20:30 2005-161T04:26:30-04:27:47 2005-161T04:33:47-04:35:04 2005-161T04:35:16-04:36:33 2005-161T04:44:59-04:46:16 2005-161T04:53:14-04:54:31 2005-161T05:00:31-05:01:48 2005-161T09:26:44-09:28:18 2005-161T09:35:55-09:37:29 2005-161T09:51:26-09:52:16 2005-161T09:58:15-09:59:49 2005-161T10:00:04-10:01:38 2005-161T10:08:59-10:10:33 2005-161T20:49:52-20:51:09 2005-161T21:13:53-21:15:10 2005-161T21:26:58-21:28:15 2005-161T23:54:43-23:56:00 2005-161T23:56:12-23:57:29 2005-162T00:04:58-00:06:15 2005-162T00:28:22-00:29:39 2005-355T16:55:02-16:55:53 2005-356T02:26:14-02:27:05 2006-179T07:57:52-07:59:35 2006-179T08:01:24-08:03:07 2006-179T08:24:54-08:26:37 2006-179T08:35:59-08:37:42 2006-179T08:58:09-08:59:52 2006-179T09:09:14-09:10:57 2006-179T09:20:19-09:22:02 2006-179T09:22:14-09:23:57 2006-179T09:24:09-09:25:52 2006-179T11:15:40-11:17:23 2010-332T07:10:54-07:11:48 2010-332T07:11:48-07:12:42 2010-332T07:12:51-07:13:45 2010-332T07:13:45-07:14:39 2010-332T07:14:48-07:16:35 2010-332T07:16:36-07:18:22 2010-332T07:18:38-07:20:25 2010-332T07:20:26-07:22:12 2010-332T08:24:14-08:25:08 2010-332T08:25:17-08:26:11 2010-332T08:26:11-08:27:04 2010-332T08:27:14-08:29:01 2010-332T08:29:02-08:30:48 2010-332T08:32:52-08:34:38 2011-237T20:10:31-20:12:49 2011-237T20:13:07-20:15:25 2011-237T20:15:44-20:16:53 2011-237T20:17:04-20:18:13 S 24 24 24 24 24 24 24 24 24 30 30 20 30 30 30 24 24 24 24 24 24 24 24 24 24 24 24 24 24 24 24 24 24 24 34 34 34 34 34 34 34 34 34 34 34 34 34 34 40 40 40 40 L 18 18 18 18 18 18 18 18 18 18 18 14 18 18 18 18 18 18 18 18 18 18 24 12 24 24 24 24 24 24 24 24 24 24 12 18 18 18 18 18 18 18 18 18 18 18 18 18 20 20 20 20 EXP (s) 160 160 160 160 160 160 160 160 160 160 160 160 160 160 160 160 160 160 160 160 160 160 160 160 160 160 160 160 160 160 160 160 160 260 80 80 80 80 160 160 160 160 80 80 80 160 160 160 160 160 80 80 PHASE (◦) 33.8 33.8 33.7 33.6 33.5 33.4 33.2 33.1 32.9 26.6 26.3 25.9 25.7 25.6 25.4 20.3 22.1 23.1 36.6 36.8 37.6 39.9 22.0 20.2 37.3 37.1 35.9 35.3 34.0 33.4 32.8 32.7 32.6 26.5 32.8 32.7 32.6 32.5 32.3 32.1 31.9 31.6 25.4 25.4 25.3 25.1 24.9 24.7 39.5 39.3 39.2 39.0 SUBSPC LON (◦) 146.2 146.3 146.5 146.8 147.3 147.4 148.0 148.5 148.9 163.7 164.2 165.0 165.4 165.5 165.9 209.2 212.1 213.6 237.8 238.2 239.9 245.1 125.8 151.2 16.7 16.8 17.0 17.1 17.3 17.4 17.5 17.5 17.5 18.5 152.2 152.7 152.7 152.8 154.9 155.0 156.6 157.8 193.3 193.8 194.4 195.5 196.1 198.3 14.6 16.8 17.9 18.5 SUBSPC LAT (◦) -5.3 -5.3 -5.3 -5.2 -5.1 -5.1 -5.0 -4.9 -4.9 -2.3 -2.2 -2.1 -2.0 -2.0 -1.9 9.6 10.5 11.0 18.3 18.4 18.8 20.1 7.9 3.2 13.0 12.9 12.3 12.0 11.4 11.1 10.8 10.7 10.6 7.5 -0.1 -0.1 -0.1 -0.1 -0.1 -0.1 -0.1 -0.1 -0.1 -0.1 -0.1 -0.1 -0.1 -0.1 0.1 0.1 0.1 0.1 SSOL LON (◦) 171.0 171.0 171.1 171.3 171.5 171.6 171.9 172.2 172.4 180.8 181.1 181.6 181.8 181.9 182.1 204.1 205.1 205.5 211.9 212.0 212.3 213.4 139.6 163.7 14.6 14.7 15.5 15.9 16.7 17.1 17.5 17.5 17.6 21.7 280.7 281.3 281.3 281.9 283.0 283.6 283.6 286.4 322.4 323.7 323.7 324.7 325.3 327.5 26.1 28.4 29.5 30.1 SSOL LAT (◦) -0.2 -0.2 -0.2 -0.2 -0.2 -0.2 -0.2 -0.2 -0.2 -0.2 -0.2 -0.2 -0.2 -0.2 -0.2 -0.2 -0.2 -0.2 -0.2 -0.2 -0.2 -0.2 -2.9 -2.4 -4.5 -4.5 -4.5 -4.5 -4.5 -4.5 -4.6 -4.6 -4.6 -4.7 7.1 7.1 7.1 7.1 7.1 7.1 7.1 7.1 7.1 7.1 7.1 7.1 7.1 7.1 10.9 10.9 10.9 10.9 SPC ALT (km) 384855. 384584. 384040. 382140. 380244. 379703. 377270. 375112. 372958. 303736. 301505. 298050. 296085. 295594. 293637. 174404. 172413. 171520. 166551. 166564. 166639. 167071. 762971. 560628. 311371. 310973. 308026. 306682. 304013. 302806. 301644. 301541. 301233. 292344. 96822. 97018. 97215. 97412. 97609. 98006. 98404. 98804. 112458. 112458. 112683. 113133. 113811. 114265. 67704. 68277. 68851. 69425. Table 6: Hyperion disk-integrated observations processed in this work. Columns S and L indicate the cube dimensions along sample and lines axis, respectively. 12 SEQUENCE OBS ID S07-V1483156810 S15-V1510202713 S15-V1510202822 S15-V1510203548 S15-V1510203657 S15-V1510204383 S15-V1510204492 S15-V1510205218 S15-V1510205327 S15-V1510206053 S15-V1510206162 S33-V1568157352 S33-V1568162035 S33-V1568162861 S33-V1568163183 S33-V1568163600 S33-V1568171295 S33-V1568171804 START-END TIME (UT) 2004-366T03:34:12-03:37:56 2005-313T04:16:22-04:17:56 2005-313T04:18:11-04:24:28 2005-313T04:30:17-04:31:51 2005-313T04:32:06-04:38:23 2005-313T04:44:12-04:45:46 2005-313T04:46:01-04:52:18 2005-313T04:58:07-04:59:41 2005-313T04:59:56-05:06:13 2005-313T05:12:02-05:13:36 2005-313T05:13:51-05:20:08 2007-253T22:40:51-22:47:26 2007-253T23:58:54-00:00:47 2007-254T00:12:40-00:17:45 2007-254T00:18:02-00:23:07 2007-254T00:24:59-00:30:04 2007-254T02:33:14-02:35:07 2007-254T02:41:43-02:42:51 S 38 30 30 30 30 30 30 30 30 30 30 48 36 42 42 42 36 32 L 34 18 18 18 18 18 18 18 18 18 18 48 36 42 42 42 36 32 EXP (s) 160 160 640 160 640 160 640 160 640 160 640 160 80 160 160 160 80 60 PHASE (◦) 38.2 39.7 39.7 39.7 39.7 39.8 39.8 39.9 39.9 39.9 39.9 33.3 33.4 33.5 33.5 33.5 33.6 33.6 SUBSPC LON (◦) 70.0 38.4 38.4 38.4 38.4 38.4 38.4 38.3 38.3 38.3 38.3 247.0 247.4 247.5 247.5 247.5 248.1 248.1 SUBSPC LAT (◦) -23.7 10.3 10.3 10.4 10.4 10.4 10.4 10.5 10.5 10.5 10.5 -10.7 -10.7 -10.7 -10.7 -10.7 -10.6 -10.6 SSOL LON (◦) 106.1 75.3 75.3 75.3 75.3 75.4 75.4 75.4 75.4 75.4 75.4 215.8 216.1 216.1 216.1 216.2 216.6 216.6 SSOL LAT (◦) -7.4 -4.7 -4.7 -4.7 -4.7 -4.7 -4.7 -4.7 -4.7 -4.7 -4.7 1.5 1.5 1.5 1.5 1.5 1.5 1.5 SPC ALT (km) 173986. 719296. 718907. 717933. 717544. 716571. 716183. 715211. 714823. 713852. 713561. 85545. 96436. 98699. 99407. 100256. 118217. 119206. Table 7: Iapetus disk-integrated observations processed in this work. Columns S and L indicate the cube dimensions along sample and lines axis, respectively. Figure 4: Wavelength of the 3.6 µm peak measured on the icy satellites as function of the sub-spacecraft longitude. On the left vertical axis is reported the water ice temperature scale. 13 5. Disk-resolved temperature maps With the goal of deriving temperature maps across specific geomorphological units, we have selected two VIMS observations which offer best view of Enceladus' south pole region and Tethys' equatorial radiation lens. 5.1. Enceladus Since the discovery of plumes (Porco et al., 2006) ejected from the tiger stripes fissures located on the south polar region of Enceladus, many studies have investigated the mechanisms originating this activity (Nimmo et al., 2007; Hedman et al., 2013; Porco et al., 2014), the composition of the plumes particles and gaseous species (Waite et al., 2012; Hansen et al., 2006; Postberg et al., 2011) and correlated them with the thermal properties of this region (Spencer et al., 2006; Howett et al., 2011b). One of the best observations of Enceladus south pole region obtained by VIMS is cube V1500059894 which was acquired from 2005-195T18:50:28 to 18:52:57 (UT) using an IR channel integration time of 160 msec/line. In this observation VIMS resolution is equal to 16.9 km/pixel while the solar phase is 46◦. A detailed discussion of the tiger stripes spectral features observed on this specific observation is reported in Brown et al. (2006). SPICE routines (Acton, 1996) and reconstructed spacecraft's attitude and trajectory kernels have been used to calculate several geometry parameters (latitude, longitude, local solar time or LST, spacecraft's distance, phase, incidence, and emission angles) for each VIMS pixel (center and 4 corners). This method has been applied to all VIMS data discussed in the rest of this work. In Figure 5 are shown the orthographic projection maps of I/F at 3.6 µm, the solar incidence angle, the ISS basemap and the temperature. Appearing dark on the I/F(3.6 µm) map, the four tiger stripes can be identified above the south pole. The area located at mid-latitudes towards the limb appears bright because photometric correction has been not applied to the VIMS data. The map of the solar incidence angle is shown in the top right panel of Fig. 5. In this figure the sub-solar point (corresponding to an incidence angle of 0◦) is rendered in black, and is located near lon=120◦ W and lat=-30◦ (i.e. on the right side of the image). The distribution of the temperature, correlated with the wavelength of the 3.6 µm peak, is shown in the bottom right panel: in this map the light blue pixels located in the south polar region around tiger stripes have a temperature T>115K. A few scattered points, marked in white, correspond to maximum T=128 K. In average the rest of the surface has T=95 K (dark blue pixels), including regions close to the subsolar point at noon local time. This result demonstrates that the method is insensitive to illumination and photometric effects, being capable to measure a ∆T ≈ 20 K between points having a ∆inc = 90◦, e.g. between tiger stripes located at morning terminator and subsolar point. Conversely, the spatial resolution influences the temperature retrieval: as a general rule thermal contrast is strongly reduced by lower spatial resolution because high temperature features are mixed with cold surfaces within the same pixel. This effect is well-evident when comparing our low spatial resolution (16.9 km/pixel) results with Goguen et al. (2013) findings, which have analyzed VIMS high spatial resolution (38 × 214 m) data collected during a 74 km low altitude Cassini's flyby above Baghdad Sulcus and have measured a color temperature Tc=197±10 K by fitting VIMS radiance in the 3-5 µm thermal range. 14 Figure 5: Enceladus disk-resolved observation V1500059894. Top left panel: I/F(3.6 µm) image mapped in ortho- graphic projection. Top right panel: solar incidence angle (in deg) map projection. Bottom left panel: ISS basemap. Bottom right panel: VIMS-derived temperature projection. 5.2. Tethys Apart from active heating, as in the case of Enceladus' south polar region previously discussed, other processes can cause anomalous thermal behavior on the surfaces of the icy moons. On Tethys, 15 as an example, significant temperature variation is observed across the equatorial region of the leading hemisphere (Howett et al., 2012). A similar effect has been reported also for Mimas (Howett et al., 2011a). These thermal anomalies are directly correlated with the irradiation of the surface by magnetospheric MeV electrons (Paranicas et al., 2010a,b), which are focused on the leading hemisphere equatorial regions, resulting in the formation of a dark "lens" visible in the IR/UV ratio ISS images (Schenk et al., 2011): as a consequence of the energy released by this space weathering, surface water ice grains are sintered together in a few centimeters-deep layer resulting in a local increase of thermal inertia inside the dark lens (Howett et al., 2012). Tethys equatorial lens appears well-resolved on VIMS observation V1567089225 shown in Fig. 6 taken from 2007-241T13:58:51 to 14:15:33 (UT) with an integration time of 640 msec/line. At the time of the observation VIMS spatial resolution is 35.1 km/pixel and solar phase 65◦. The equatorial belt corresponds to the dark area visible in the center of the I/F(3.6 µm) image and ISS basemap mapped in orthographic projection shown in Figure 6. On the same images is partially visible the 400 km-wide Odysseus impact basin centered along the north terminator at lon=130◦, lat=30◦. As shown in the top right panel, the subsolar point is located close to the limb at lat=-10◦. The temperature map, in the bottom right panel, reveals that south latitude regions close to the subsolar point have the highest temperature, T>115 K (light blue pixels). Black pixels identify cold areas at T≤88 K which are localized in the low-albedo equatorial lens and along the terminator, including the Odysseus crater floor. The remaining points at both north and south latitudes coded in blue are at T≈100 K. It is remarkable that on this image a large thermal gradient (∆T > 27 K) visible near the subsolar point (noon time) between the equatorial lens and the nearby southern regions. This effect could be the consequence of the large thermal inertia difference reported on these two regions by CIRS, corresponding to 5 and 25 J m−2 K−1 s−1/2 outside and inside the equatorial lens, respectively (Howett et al., 2012). 6. Averages temperature cylindrical maps By using data projection techniques the temperature distribution derived from VIMS observations has been mapped on the regular satellites' surfaces and correlated with geomorphological features, including albedo units, equatorial radiation lenses and impact craters. VIMS observations are filtered in two groups corresponding to the pre-equinox (years 2004-2009, heliocentric distance 9.0-9.4 AU) and post-equinox (years 2009-2012, heliocentric distance 9.4-9.7 AU) periods. This approach allows us to follow seasonal variations because the subsolar point moves from south latitudes during the pre-equinox to north latitudes during the post-equinox period. VIMS data are further filtered to retrieve all pixels of each satellite matching the following multiple conditions: 1. acquired from distances lower than 300,000 km (corresponding to a VIMS spatial resolution better than 150 km/pixel) and 2. with incidence and emission angles lower than 80◦ (to avoid very oblique and distorted views 3. with 10:00 ≤ LST ≤ 14:00 hr, e.g., around noon time (to limit influence of local time) and of the surface) and 16 Figure 6: Tethys disk-resolved observation V1567089225. Top left panel: I/F(3.6 µm) image mapped in orthographic projection. Top right panel: solar incidence angle (in deg) map projection. Bottom left panel: ISS basemap. Bottom right panel: VIMS-derived temperature projection. 4. taken in the 2004-2009 (pre equinox) and 2009-2012 (post equinox) periods (to follow temper- ature seasonal changes between north and south hemispheres). A summary of the number of cubes and pixels resulting from this selection is summarized in Table 17 Satellite Mimas Enceladus Tethys Dione Rhea Hyperion Iapetus number of observations 703 1917 802 938 1070 523 159 pixels 10581 104787 60626 120838 144747 8087 46860 number of Resolution (km/bin) of a 1◦ × 1◦ bin 3.5 4.4 9.3 9.8 13.3 2.3 12.8 Table 8: Summary of Saturn's icy satellites observations processed in this work. 8 for each satellite. Temperature maps are rendered in cylindrical projection by applying a fixed 1◦ × 1◦ grid in longitude and latitude. As satellites have different radii, the spatial resolution on the maps changes according to the values reported in Table 8. Finally, for each 1◦ × 1◦ bin, the median value of the temperature and the data redundancy (number of VIMS pixels covering a given bin) have been calculated and rendered on the cylindrical maps. A discussion of the results obtained is given in the next paragraphs for each satellite. 6.1. Mimas The principal geological feature of the small Mimas is the presence of the 110 km-wide Herschel crater dominating the center of the leading hemiphere. Many other impact craters are spread across the surface which, in contrast with the outer moons, lack significant tectonic features. An equatorial lens on the leading hemisphere, caused by the bombardment of MeV energy magnetospheric electrons corresponds to the Thermal Anomaly Region (TAR) reported by Howett et al. (2011a). Following the data selection criteria previously exposed, only pre-equinox data are available for Mimas. The temperature map shown in Fig. 7-top left panel covers the leading hemisphere between 95◦ ≤ lon ≤ 180◦, -60◦ ≤ lat ≤ +50◦. The VIMS dataset has up to 13 independent pixels covering a given 1◦ × 1◦ bin (Fig. 7-top right panel). On the equator of the leading hemisphere the TAR has a T=100 K, about 10 K colder than the external regions. The lens extends up to latitudes ±45◦ and is interrupted in a pattern corresponding to the Herschel crater located at the equator at about lon=100◦ where the temperature increases up to T=110 K (dark red pixels). VIMS temperature distribution across the TAR matches with the light blue regions visible on ISS IR-Green-UV color composite map (Fig. 7-bottom left panel) and with CIRS night-time temperature map by Howett et al. (2011a) (Fig. 7-bottom left panel). 6.2. Enceladus Enceladus' surface is shaped by the active processes occurring in the warm "tiger stripes" features located in the south pole region. These features consists of four main linear, almost parallel troughs (Alexandria, Cairo, Baghdad, and Damascus Sulci) about 500 m deep, 2 km wide and 130 km long 18 Top Right panel: VIMS data redundancy map (number of VIMS pixels/bin). Top left panel: Pre-equinox temperature map derived from the 3.6 µm peak Figure 7: Mimas maps. Bottom left position. panel: context image map based on Cassini-ISS IR-Green-UV color composite (by P. Schenk, available from http://photojournal.jpl.nasa.gov/catalog/PIA18437). Thermal Anomaly Region (TAR) extension is indicated by the oval; Herschel crater is identified by the letter H. Bottom right panel: CIRS night-time temperature map (Howett et al., 2011a). (Porco et al., 2006) from which plumes are ejected. According to Waite et al. (2012) the plumes are composed mainly by water vapor and carbon dioxide, carbon monoxide or molecular nitrogen, and methane. Moreover, the plume activity is not constant but shows a periodicity linked to orbital position (Hedman et al., 2013) and tides (Porco et al., 2014). Wide fractured plains, resurfaced by tectonism, are located on both hemispheres while the north hemisphere shows more impact craters (Jaumann et al., 2009). For Enceladus both trailing and leading hemisphere coverage (Fig. 8 - top and center panels) is available during the pre-equinox and post-equinox periods. Pre-equinox temperature map includes the fractured and ridged plains of the trailing hemisphere spanning above Diyar and Sarandib Planitia (marked as DP and SP in Fig. 8 - bottom left panel), between 185◦ ≤ lon ≤ 360◦, -85◦ ≤ lat ≤ +50◦ with a narrow gap at about lon=270◦ (Fig. 8 -top left panel). A small coverage of the leading south hemisphere between 0◦ ≤ lon ≤ 15◦ is available. The maximum temperature measured is T=120 K above Damascus Sulcus (DS) rendered in red color on the temperature map. With the exclusion of 19 the south pole active region, most of entire south hemisphere during the summer season has a T=105 K (yellow pixels) while the regions located northwards of lat=10◦, in winter season, most are colder at T≤ 95 K. Figure 8: Enceladus maps. Top left panel: Pre-equinox temperature map. Top Right panel: VIMS pre-equinox data redundancy map. Center left panel: Post-equinox temperature map. Center Right panel: VIMS pre-equinox data redundancy map. Bottom left panel: context image map based on Cassini-ISS IR-Green-UV color composite (by P. Schenk, available from http://photojournal.jpl.nasa.gov/catalog/PIA18435). Locations of Diyar Planitia (DP), Sarandib Planitia (SP), Alexandria (AS) and Damascus Sulci (DS) are indicated. Bottom right panel: distribution of the plumes particles accumulation on the surface in mm/year by Kempf et al. (2010). The post-equinox temperature map covers the heavy cratered terrains on the leading hemisphere 20 between 0◦ ≤ lon ≤ 220◦, -85◦ ≤ lat ≤ +75◦ with a narrow gap at about lon=110◦ (Fig. 8 - center left panel). A partial coverage of the trailing equatorial hemisphere between 330◦ ≤ lon ≤ 360◦ is available. On this map the maximum temperature measured is T=110 K in the region corresponding to Alexandria Sulcus (rendered in red color, indicated as AS in Fig. 8 - bottom left panel). A wide region of the leading hemisphere, rendered in orange has T=100 K spanning on both north (summer) and south (winter) hemispheres. Towards the trailing hemisphere temperature drops below T<95 K (light blue pixels). Temperature variations appear correlated with surface colors (see the ISS IR-Green-UV color composite map shown in Fig. 8 - bottom left panel). The two yellowish areas centered in the middle of the leading and trailing north-equatorial hemispheres on ISS map appear warmer in VIMS maps than the neutral gray areas on Diyar Planitia placed in the antisaturnian hemisphere. Conversely, the relationship with the distribution of the tiger stripes plumes deposits is less evident (see the deposit model by Kempf et al. (2010) shown in Fig. 8 - bottom right panel, valid for grains smaller than 5 µm). The downwelling fraction of the material released from tiger stripes deposits preferentially in the nearby regions (green areas in Fig. 8 - bottom right panel) and in the south hemisphere along two V-shaped regions centered on lon=45◦ and 225◦ meridians (red areas). According to Kempf et al. (2010) these deposits grow with a rate of 0.5 mm/year in the vicinity of active sources which reduces to 10−5 mm/year at the equator. As a consequence of the accumulation of fresh water ice grains, these two V-shaped regions appear less yellow and more neutral in color at UV-visible wavelengths on the ISS map. VIMS post-equinox temperature map (Fig. 8 - center left panel) doesn't vary but stays at a T=105 K (orange pixels) across a wide sector of the leading hemisphere, where the heavy cratered terrains are located, including the V-shaped plume deposit centered on the lon = 45◦ meridian. A similar temperature T=105 K (yellow pixels) is observed on a large part of the equatorial and mid-southern latitudes areas of the trailing hemisphere on the pre-equinox map (Fig. 8 - top left panel). In this case the plume deposit accumulates across the fractured and ridged plains. Conversely, a small decrease of temperature (T<100 K) between 0◦ < lon < 30◦ in the south hemisphere and towards the lon=180◦ meridian (antisaturnian point) are observed. Since these two areas are immediately outside the plume deposits, a similar behavior implies a change in composition, grain size and thermal inertia or a combination among these factors. Noteworthy, in correspondence of the V-shape deposits, VIMS has measured higher water ice band depth and larger grain sizes (up to 40 µm) with respect to the nearby equatorial regions (Jaumann et al., 2008). A similar result implies the presence of a maturation process, like annealing or sintering, in which the small (< 5 µm) plume particles coalesce in larger grains resulting in a more compact surface layer. As a consequence of these evidences, one should expect high thermal inertia across the two V-shaped plume deposits and low elsewhere. However, VIMS temperature data taken around noon local solar time show a similar trend. Finally, thanks to the large redundancy of the Enceladus dataset, it is possible to investigate In order to follow daytime the daytime temperature variation for a given point of the surface. temperature variations, the VIMS geometry archive has been exploited in order to search the point having the more extended coverage in local solar time. From this search results that the 1◦ × 1◦ bin 21 Figure 9: Enceladus daytime temperature variations. Left panel: VIMS daytime temperature variation at lon=95◦, lat=-3◦. Temperature retrieval is limited to T≥88 K (horizontal dashed line). VIMS point at local time 105◦ is an upper limit. Right panel: CIRS daytime and nocturnal temperature cycle for points at −10◦ ≤ lat ≤ 0◦ (Howett et al., 2010). centered at lon=95◦, lat=-3◦ has the maximum temporal coverage, being observed between 6 hr ≤ LST ≤ 16 hr and therefore covering about all daytime times. The resulting temperature daytime curve (Fig. 9 - left panel) shows the increase of temperature from T≤ 88 K at LST=105◦ to T=89 K at LST=225◦. The maximum T=104 K is observed around noon at LST=165-195◦. This to demonstrate that the method based on the 3.6 µm continuum peak wavelength has the sensitivity to follow daytime variation of the temperature. As mentioned before, temperature values retrieved by VIMS with this method appear higher than similar ones measured by CIRS at longer wavelengths. The daytime temperature curve measured by CIRS on points located at −10◦ ≤ lat ≤ 0◦ are in fact lower (Fig. 9 - right panel, from Howett et al. (2010)). At noon CIRS measures a temperature T=80 K, about 25 K lower than VIMS. The different skindepth sensed by two instruments can be the cause of this effect. 6.3. Tethys Tethys' surface is densely cratered, showing among the largest impact basins in the Saturnian system, like the 400 km-wide Odysseus crater. Cutting across the Saturn-facing meridian is located the 2500 km-long Ithaca Chasma, a 100 km-wide rift running approximately along the north-south direction with multiple parallel fractures. Tethys, like Dione and Rhea, show an evident albedo asymmetry between the leading (bright) and trailing (dark) hemispheres (Verbiscer and Veverka, 1989; Buratti et al., 1990; Schenk et al., 2011). Tethys' equatorial lens on the leading hemisphere equator is the place where the magnetosphere focuses high energy (E>1 MeV) electrons resulting in alteration of surface grains Paranicas et al. (2010a,b). As a consequence of this processes an enhancement of the UV signal and a darkening in the infrared has been observed (Schenk et al., 22 2011). The trailing hemisphere experiences the interaction with the plasma flow particles resulting in an accumulation of non-icy material (Jaumann et al., 2009). A wide coverage of Tethys's surface is available for both pre and post-equinox datasets ( Fig. 10 - top and center right panels). Thank to this is possible to map the temperature changes across the leading hemisphere Thermal Anomaly Region (TAR) feature and to follow seasonal temperature changes occurring on the north and south hemisphere regions. The cold TAR, located in the equato- rial region of the leading hemisphere, appears at T=100 K (light orange pixels) on both pre (Fig. 10 - top left panel) and post-equinox Fig. 10 - center left panel) VIMS datasets. On the post-equinox map the equatorial border of the TAR located on the antisaturnian hemisphere appears less con- trasted respect to the saturnian side. A similar "Pac-Man" feature (Howett et al., 2012) is visible on Cassini-CIRS data (Fig. 10 - bottom right panel). On the pre-equinox map the oval-shaped bound- ary of the feature is noisy as a consequence of the scarcity of high spatial resolution observations available, in particular in the antisaturnian quadrant towards the Odysseus crater (Fig. 10 - top right panel). On Ithaca Chasma (IC), the 100 km-wide rift spanning from about lon=45◦, lat=-75◦ to lon=315◦, lat=75◦ and intersecting the equatorial lens border at about lon=0◦, VIMS observes a low T≤ 88 K (blue pixels). This decrease of temperature is a consequence of the presence of bright material and shadows occurring in the very irregular rift structure. Moreover, VIMS data indicate that the seasonal cycle changes the temperature distribution: the south hemisphere, warmer during the pre-equinox period, becomes colder during the post-equinox while the reverse is happening on the north hemisphere. The TAR remains the coldest region of the leading hemisphere during the entire seasonal cycle. Starting form the TAR, moving in the south summer hemisphere down to lat=-50◦, the tem- perature rises to T=110 K (red pixels). Conversely, moving northward of the TAR in the winter hemisphere the temperature remains below T=105 K (orange pixels). The center of Tethys' trailing hemisphere is redder and darken than the leading, as shown in the ISS map in Fig. 10 - bottom left panel. As a consequence of this low albedo feature, the temperature increases on the trailing hemisphere. Despite limited by the coverage, VIMS map shows that in the south trailing hemisphere, antisaturnian quadrant, between 175◦ ≤ lon ≤ 230◦, 0◦ ≤ lon ≤ -60◦ a wide region is at T=110 K (red pixels). Similar high temperature is observed in the north hemisphere around lon=315◦. Finally, on the saturnian quadrant of the trailing hemisphere at about lon=350◦-360◦ lat=0◦-10◦ is visible the north branch of Itacha Chasma at T≤ 88 K (blue pixels). Lower temperatures are seen around Telemachus crater (lon=340◦, lat=55◦). 6.4. Dione Dione's surface is characterized by geomorphological and albedo differences between the leading and trailing hemispheres. Old crater units are located on the bright leading hemisphere while the dark trailing shows a complex network of chasmata, probably caused by extensional tectonism, resulting in the surfacing of bright wispy lineaments (Jaumann et al., 2009). VIMS coverage on Dione is mainly available across two wide regions centered across lon=0◦ and 180◦ meridians (see Fig. 11 - right top and center panels). The pre-equinox temperature map (Fig. 23 Figure 10: Tethys maps. Top left panel: Pre-equinox temperature map. Top Right panel: VIMS pre-equinox data redundancy map. Center left panel: Post-equinox temperature map. Center Right panel: VIMS pre-equinox data redundancy map. Bottom left panel: context image map based on Cassini-ISS IR-Green-UV color composite (by P. Schenk, available from http://photojournal.jpl.nasa.gov/catalog/PIA18439). Locations of Thermal Anomaly Region (TAR), Odysseus crater (O), Ithaca Chasma (IC) and Telemachus crater (T) are indicated. Bottom right panel: CIRS daytime temperature on ISS IR/UV basemap (courtesy C. Howett, personal communication). 11 - top left panel) shows that the maximum temperature T=140 K is correlated with the dark material units at the center of the trailing hemisphere (see ISS IR-Green-UV color composite map in Fig. 11 - bottom left panel), corresponding to Ilia-Tumus (IT) and Tiburtus-Mezentium (TM) 24 craters. The bright Eurotas Chasmata (EC) and the meridional part of the antisaturnian hemisphere up to lat =85◦ are cooler at T=125 (orange pixels). Figure 11: Dione maps. Top left panel: Pre-equinox temperature map. Top Right panel: VIMS pre-equinox data redundancy map. Center left panel: Post-equinox temperature map. Center Right panel: VIMS pre-equinox data redundancy map. Bottom left panel: context image map based on Cassini-ISS IR-Green-UV color composite (by P. Schenk, available from http://photojournal.jpl.nasa.gov/catalog/PIA18434). Locations of Ilia-Tumus craters (IT), Tiburtus-Mezentium craters (TM), Eurotas Chasmata (EC), Creusa crater (C), Padua Chasmata (PC) and Palinurus crater (P) are indicated. Bottom right panel: CIRS thermal inertia map (Howett et al., 2014). A great part of the leading hemisphere between -60◦ ≤ lat ≤ 40◦ is at T=115 K (yellow pixels). 25 A temperature of 90 K (light blue pixels) is reached in the north regions, in winter season, above lat=40◦. A few pixels are at T≤ 88 K in the bright terrains near the recent Creusa crater (C). The low temperature measured in this area is compatible with the prominent water ice band depth measured by VIMS on the crater and on the bright ejecta rays departing from it (Stephan et al., 2010; Scipioni et al., 2013). The seasonal change is well-evident comparing the pre-equinox with the post-equinox map (Fig. 11 - center left panel): despite the lower coverage on the trailing hemisphere is possible to distinguish how the south hemisphere regions become cooler, down to T ≤ 88 K for lat ≤ 50 K (dark blue pixels). Conversely, the dark units on the trailing hemisphere remain stable at T=140 K (red pixels). In the south part of the maximum temperature spot is visible the meridional segment of Padua Chasmata (PC) where the temperature drops to T=135 K (light red pixels). The equatorial belt of the leading hemisphere is at T=115 K (yellow pixels) for lat = ± 40◦. Mid-latitude regions in both hemispheres are at T=90 K (light blue pixels). Finally, VIMS temperature maps don't show evidence of thermal anomaly on Dione's leading hemisphere nor clear temperature changes in correspondence of the thermal inertia peak (at 16 J m−2 K−1 s−1/2) above Palinurus crater (P) as reported by CIRS (Howett et al., 2014) and shown in Fig. 11 - bottom right panel. 6.5. Rhea Similarly to Dione, also Rhea's surface possesses geomorphological and albedo differences between the leading and trailing hemispheres. The leading hemisphere appear bright, showing a great variety of impact craters, from small ones to 400 km-wide Tirawa and Mamaldi basins. Among the small craters on the leading hemisphere is the remarkable Inktomi, the source of many extended ejecta rays. The trailing hemisphere is darker, with bright highlands crossed by a complex system of chasmata, appearing as wispy filaments, which extends along the north-south direction. A resurfacing event could be the cause of these features (Jaumann et al., 2009). The albedo asymmetry between the two hemispheres is probably caused by the exposure of the leading surface to E ring particles while magnetosphere focuses charged particles on the trailing hemisphere (Schenk et al., 2011) resulting in accumulation of non-icy material. The first process causes the layering of bright water ice grains while the second causes reddening and darkening through the implantation of charged particles. For reference, the ISS color context image with the position of the geomorphological features cited in the text is shown in Fig. 12 - bottom left panel. The VIMS dataset offers a wide coverage of Rhea's surface with missing parts at about lon=90◦ and 270◦ on both pre and post-equinox maps (Fig. 12 - top and center right panels). On the pre-equinox map (Fig. 12 - top left panel) there is a clear temperature difference between the leading hemisphere (which is almost a uniform 125 K, shown by the orange pixels) and the trailing hemisphere (which has a maximum temperature of 150, shown by the red pixels, particularly around the Kumpara-Heller craters (KH) and above Powehiwehi crater (P). Above Galunlati, Avalki, Avaiki Chasmata (or Chasmata Region, CR) running along north-south direction between lon=250◦-300◦ meridians a temperature T=140 K (light red pixels) is retrieved. The two ancient Tirawa (T) and Mamaldi (M) impact basins don't show a temperature change with respect to the surroundings. However, moving beyond the north rim of Tirawa, the temperature 26 drops to T=110 K. Figure 12: Rhea maps. Top left panel: Pre-equinox temperature map. Top Right panel: VIMS pre-equinox data redundancy map. Center left panel: Post-equinox temperature map. Center Right panel: VIMS pre-equinox data redundancy map. Bottom left panel: context image map based on Cassini-ISS IR-Green-UV color composite (by P. Schenk, available from http://photojournal.jpl.nasa.gov/catalog/PIA18438). Locations of Tirawa (T) and Mamaldi (M) impact basins, Kumpara-Heller craters (KH), Chasmata Region (CR), Powehiwehi crater (P), Wakonda crater (W), Inktomi crater (I), Nzame-Karora-Haoso craters (NKH) and Arunaka-Con craters (AC) are indicated. Bottom right panel: CIRS thermal inertia map (Howett et al., 2014). A temperature T=135-140 K (light red pixes) is measured across the low albedo units in the trailing hemisphere, including the ones around the CR. The region around Wakonda (W) crater 27 appears colder at T=125 K (orange pixels). The colder temperatures T< 110 K (light and dark pixels) are measured in the north (winter) hemisphere during the pre-equinox period above lat=20◦ in the antisaturnian hemiphere. A significant temperature anomaly is observed around the Inktomi (I) crater, placed in the south (summer) leading hemisphere at lat=-12.5◦: while surrounding terrains show T=125 K the crater has a T=110 K (light blue pixels). A similar thermal anomaly feature is visible on the thermal inertia map derived by Howett et al. (2014) from Cassini-CIRS data (Fig. 12 - bottom right panel). VIMS and CIRS are in agreement because an increase of thermal anomaly causes a decrease of daytime temperature. Temperature differences between leading and trailing hemisphere continue to be present in the post-equinox map (Fig. 12 - center left panel). The equatorial region of the leading hemisphere is at T=120 K (light orange pixles) with the exclusion of a wide region at T<110 K (light and dark pixels) spanning from equator close to Nzame-Karora-Haoso craters (NKH) to lat=-30◦, where Arunaka-Con craters (AC) are located. Another region with similar low temperature is observed in the antisaturnian meridional quadrant (around lon=150◦, lat=-50◦) and seems to be associated with the bright material surrounding the dark units in the trailing hemisphere (see ISS color context map in Fig. 12 - bottom left panel). Large variations of temperature are seen in the surroundings of CR on the trailing hemisphere. These are caused by the albedo differences between the bright rays departing from the CR features, which are at T=125 K (orange pixels) and the neighbor dark material units between T=150 K (red pixels) and T=135-140 (light red pixels). 6.6. Hyperion Hyperion's irregular and porous body shows two distinct morphological units. A strong water ice signature and high albedo characterize the first unit class. Carbon dioxide and C-N, or possibly adsorbed H2 (Clark et al., 2012), are linked to dark material accumulated on the bottoms of cup-like craters, forming a second unit (Cruikshank et al., 2007). The limited number of available Hyperion observations fulfilling the selection rules exposed at the beginning of this section has forced us to increase the range of local solar times from 10≤LST≤14 hr to 6≤LST≤18 hr. As a consequence of this change, and differently from all the other satellites except Iapetus, for Hyperion we are reporting average daytime temperature maps in Fig. 13 - top and bottom left panels. The satellite's surface temperature is quite uniform on both pre and post- equinox datasets, equal to T=140 K and T=130 K, respectively. This difference between the pre and post-equinox periods is probably caused by the averaging of daytime temperature taken at very different local times. Certainly VIMS data indicate that the Hyperion's surface is almost isothermal. 6.7. Iapetus At a mean distance of 59 Saturn radii from Saturn, Iapetus is the only regular moon in the Solar System orbiting at the largest distance from its planet. Moreover, the moon is orbiting outside Saturn's magnetosphere (which extends up to 20 Saturn's radii, Gombosi et al. (2009)) thus exposing the surface to solar weathering. Iapetus' leading hemisphere surface, corresponding to the Cassini Regio (CR), appears strongly contaminated by a layer of dark organic material and carbon dioxide (Buratti et al., 2005; Filacchione et al., 2007; Cruikshank et al., 2008, 2010; Clark et al., 2012). The 28 Figure 13: Hyperion maps. Top left panel: Pre-equinox temperature map for 6≤LST≤18. Top Right panel: VIMS pre-equinox data redundancy map. Center left panel: Post-equinox temperature map for 6≤LST≤18. Center Right panel: VIMS pre-equinox data redundancy map. trailing hemisphere shows a more pristine water ice-rich surface across the northern Roncevaux Terra (RT) and the southern Saragossa Terra (ST) where the large Engelier crater (E) is located (Denk et al., 2010). The deposition of dark material and the high diurnal temperatures measured on the leading hemisphere have been suggested as possible causes of thermal migration of water ice (Spencer and Denk , 2010). A discussion about the origin and composition of the aromatic and aliphatic hydrocarbons seen on Iapetus' leading hemisphere surface is given by Cruikshank et al. (2014). Different pathways have been investigated to trace the possible sources of the dark material, including the transfer of dust from Phoebe (Burns et al., 1979; Tosi et al., 2010; Tamayo et al., 2011), the dynamics of Phoebe's ring particles (Verbiscer et al., 2009; Hamilton et al., 2015) and the dust resulting from collisions among the outer irregular satellites (Bottke et al., 2010; Tosi et al., 2010; Tamayo et al., 2011). Morevoer, Cassini Radar observations at 2.2 cm wavelength taken above the leading hemisphere have indicated that Iapetus' icy subsurface is more consolidated than the uppermost dark material layers (Le Gall et al., 2014). Only pre-equinox data are available for Iapetus, but they give nearly complete coverage in the 30◦ 29 ≤ lon ≤ 300◦ range (Fig. 14 - top right panel). For Iapetus, as for Hyperion, the interval of local solar times has been increased from 10≤LST≤14 hr to 6≤LST≤18 hr in order to achieve better spatial coverage. Therefore the VIMS temperature map of Iapetus shown in Fig. 14 - top left panel, is built by selecting observations taken during all daytime hours. The maximum retrieved temperature T> 170 K is measured above the dark Cassini Regio (CR) in the leading hemisphere where it appears uniformly distributed and well correlated with the dark material unit. Moving towards the trailing hemisphere, VIMS has observed the Carcassone Montes (CM) transition region between the dark and bright units where temperature gradually drops to T=160 K (orange pixels) and then to T=150 K (yellow pixels). This is the same region where Denk et al. (2010) have observed color differences within the dark terrain, which shows a less reddish color than the uniformly dark CR. The bright icy terrains on the north trailing hemisphere in Roncevaux Terra (RT) and in the south Saragossa Terra (ST) are at T=140 K (light blue). The Hamon crater (H) and other three nearby craters, whose floors are filled by dark material, are remarkably warmer than neighboring terrains: around these craters a temperature peak of T=160 K with an external corona at T=150 K are measured, resulting in a clear contrast to the nearby terrains at T=140 K (light blue). Finally, a thermal contrast is seen between the central peak of the large Engelier crater (E) at T=150 K (yellow pixels) and the crater's floor at T=140 K (light blue). 7. Conclusions Daytime temperature of Saturn's icy satellites are derived from Cassini-VIMS data by fitting the position of the continuum at 3.6 µm on I/F spectra of water ice. This spectral feature is temperature-dependent and has been calibrated by using laboratory data at some specific conditions. However this retrieval is based on the assumption of pure water ice composition. The presence of other endmembers could alter the shape of the continuum at 3.6 µm, implying uncertainties in the temperature derivation process. However, since water ice is the dominant species on many satellites and the 3.6 µm peak is always present in their spectra, even on Iapetus' dark terrain (Filacchione et al., 2007), we have assumed that the method can be exploited within the framework of these limitations. The availability of laboratory reflectance data allows the derivation of temperature in the range 88≤T≤168 K. The error associated with the fitting technique is of the order of 5 K. Further modeling is necessary to disentangle composition, including contaminant fraction, mixing modality and grain size distribution, from thermal inertia. While composition can be inferred from VIMS data, the 3.6 µm reflectance peak method allows us to determine only diurnal temperatures. In absence of night time temperature measurements, thermal inertia cannot be properly inferred. The method has been applied to different VIMS datasets with different spatial resolutions, from disk-integrated observations to regional coverage data suitable to build maps through data filtering, averaging and mapping techniques. Disk-integrated observations allow us to explore large scale thermal response between leading and trailing hemispheres of the satellites, resulting in the following properties: 1. Mimas, Enceladus and Hyperion don't show noteworthy temperature differences between the two hemispheres. 30 Figure 14: Iapetus maps. Top left panel: Pre-equinox temperature map. Top Right panel: VIMS pre-equinox data redundancy map. Bottom left panel: context image map based on Cassini-ISS IR-Green-UV color composite (by P. Schenk, available from http://photojournal.jpl.nasa.gov/catalog/PIA18436). Locations of Cassini Regio (CR), Carcassone Montes (CM), Roncevaux Terra (RT), Saragossa Terra (ST), Hamon crater (H) and Engelier crater (E) are indicated. 2. Satellites possessing an albedo asymmetry between leading and trailing hemispheres, e.g., Tethys, Dione and Rhea, show temperature higher of about 10 K on the dark trailing than on the bright leading hemisphere; 3. Much higher temperature is observed on the dark leading hemisphere of Iapetus than on the bright trailing hemisphere. Since at disk-integrated spatial resolution is not possible to resolve geomorphological classes of in- terest on the surfaces, the temperature is correlated with albedo at hemispherical scale. Moreover, VIMS results clearly show the existence of temperature trend among satellites. In fact, tempera- ture increases with the orbital distance from Saturn: at T≤ 88 K Mimas, Enceladus and Tethys are the colder satellites (see Fig. 15). The minimum position of the 3.6 µm peak is observed on Enceladus. Moving towards larger orbital distances the noon local time temperature progressively increases: on Dione T=98-118 K, on Rhea T=108-128 K, on Hyperion T=118-128 K. On Iapetus' trailing hemisphere T=128-148 K is measured. The maximum temperature T>168 K is observed 31 on Iapetus leading hemisphere. This trend matches the distribution of the albedo and consequently of water ice and chromophores across the Saturn satellites' surfaces. Compositional gradients have been traced by means of the 2.0 µm water ice band depth and the visible spectral slopes derived from VIMS reflectance spectra: while the water ice abundance is almost constant, spectral slopes show a reddening as they move from inner to outer satellites (Filacchione et al., 2013). Figure 15: Temperature distribution as function of satellites orbital distance from Saturn as derived from VIMS disk-integrated observations. The correlation of temperature with geomorphological features becomes evident when VIMS observations taken at regional scale are projected in cylindrical maps. VIMS temperature maps taken around noon time (10≤LST≤14) show the following findings: • Detection of the thermal anomaly across the equatorial lens of Mimas and Tethys; • Measurement of the temperature T>115K close to Enceladus' Damascus and Alexandria sulci in the south pole region; • Verification of seasonal temperature changes: on Tethys, Dione and Rhea higher temperature are measured above the south hemisphere during pre-equinox and above north hemisphere during post-equinox epochs; 32 • The temperature distribution appears correlated with albedo features: the temperature in- creases on low albedo units located on Tethys, Dione and Rhea trailing hemispheres. • The thermal anomaly region on Rhea's Inktomi crater detected by CIRS (Howett et al., 2014) is confirmed by VIMS: this area appears colder with respect to surrounding terrains when observed at the same local solar time. Moreover, the method is able to follow daytime temperature variations: on the equatorial region of Enceladus values of 88≤T≤110 K have been measured between 6≤LST≤16 hr. VIMS daytime temperature maps (6≤LST≤18 hr) are available for Hyperion, which appears remarkably uniform at about T=130-140 K, and for Iapetus, which shows a remarkable tempera- ture asymmetry between the dark Cassini Regio (T> 168 K) and the bright plains on the trailing hemisphere (T=140 K). Considering the high temperatures derived by VIMS, a question arises: how is possible to maintain CO2 on the warm surface units of Iapetus, where T>168 K? These high temperatures put some severe constraints on the occurrence of CO2. VIMS has observed the 4.255 µm absorption band of CO2 adsorpted in water ice above Iapetus' dark terrain units (Buratti et al., 2005; Filacchione et al., 2007; Cruikshank et al., 2010). Since carbon dioxide is not among the pristine volatile species available at the time of formation of the Saturn's satellites, or all of the original CO2 near the surface has been lost, or it must be synthesized by UV and cosmic rays irradiation of water ice and a carbon source (Moore et al., 1983) in the lower subsurface layers. If all the CO2 is complexed with H2O, as discussed in Cruikshank et al. (2010), and is a true clathrate, then the H2O cage that holds the CO2 can maintain it even at relatively high temperatures: Sandford and Allamandola (1990) have observed CO2 bound to H2O ice up to T=150 K. More recent studies have reported that the H2O − CO2 clathrate hydrates undergo to "self-preservation" mode, an anomalous slow decomposition of the cage happening below the melting point of the ice (Oancea et al., 2012). In this regime the CO2 hydrated core is maintained insulated by a very thin water ice crust preventing the sublimation of the CO2 (Falenty and Kuhs, 2009). This phenomenon is commonly observed at 1 bar between T=200-270 K, but it has also been observed under vacuum or at moderate pressure. A similar scenario is therefore plausible to maintain CO2 bound to H2O at T> 168 K on Iapetus dark terrains. Keeping in mind its intrinsic limitations, the 3.6 µm continuum peak wavelength technique can be employed to remotely sense surfaces' temperature of water ice-rich bodies in the outer solar system. In these cases infrared spectroscopy offers the major advantage to allow us to retrieve simultaneously the surface composition, grain size distribution and temperature from the same dataset. Acknowledgments The authors acknowledge the financial support from INAF-IAPS, ASI-Italian Space Agency and NASA through the Cassini project. We gratefully thank the Cassini Icy Satellites Working group, VIMS technical team at Lunar and Planetary Lab (University of Arizona, Tucson, AZ) and the Cassini-Huygens Project at JPL (Pasadena, CA) for observations planning, sequencing and data 33 archiving. Without all them this study would be impossible. This research has made use of NASA's Astrophysics Data System. References References Acton, C. H., 1996. Ancillary data services of NASA's Navigation and Ancillary Information Facility, PSS, 44, Issue 1, 65-70, doi: 10.1016/0032-0633(95)00107-7 Bottke, W. F., Nesvorny, D., Vokrouhlicky, D. Morbidelli, A., 2010. The Irregular Satellites: The Most Collisionally Evolved Populations in the Solar System, AJ, 139, 994-1014, doi: 10.1088/0004- 6256/139/3/994. Brown, R.H., Baines, K.H., Bellucci, G., Bibring, J.-P., Buratti, B.J., Capaccioni, F., Cerroni, P., Clark, R.N., Coradini, A., Cruikshank, D.P., Drossart, P., Formisano, V., Jaumann, R., Langevin, Y., Matson, D.L., McCord, T.B., Mennella, V., Miller, E., Nelson, R.M., Nicholson, P.D., Sicardy, B., Sotin, C., 2004. The Cassini Visual and Infrared Mapping Spectrometer (VIMS) investigation. Space Sci. Rev. 115 (1-4), 111-168. Brown, R. H., Clark, R. N., Buratti, B. J., Cruikshank, D. P., Barnes, J. W., Mastrapa, R. M. E., Bauer, J. Newman, S., Momary, T., Baines, K. H., Bellucci, G., Capaccioni, F., Cerroni, P., Combes, M., Coradini, A., Drossart, P., Formisano, V., Jaumann, R., Langevin, Y., Matson, D. L., McCord, T. B., Nelson, R. M., Nicholson, P. D., Sicardy, B., Sotin, C., 2006. Composition and Physical Properties of Enceladus' Surface, Science, 311, 1425-1428, doi: 10.1126/science.1121031. Buratti, B.J., Mosher, J.A., Johnson, T.V., 1990. Albedo and color maps of the saturnian satellites, Icarus, 87, 339-357. Buratti, B. J., Cruikshank, D. P., Brown, R. H., Clark, R. N., Bauer, J. M., Jaumann, R., McCord, T. B., Simonelli, D. P., Hibbitts, C. A., Hansen, G. B., Owen, T. C., Baines, K. H., Bellucci, G., Bibring, J.-P., Capaccioni, F., Cerroni, P., Coradini, A., Drossart, P., Formisano, V., Langevin, Y., Matson, D. L., Mennella, V., Nelson, R. M., Nicholson, P. D., Sicardy, B., Sotin, C., Roush, T. L., Soderlund, K., Muradyan, A., 2005. Cassini Visual and Infrared Mapping Spectrometer Observations of Iapetus: Detection of CO2, ApJ, 622, L149-L152, doi: 10.1086/429800. Buratti, B. J., Soderlund, K., Bauer, J., Mosher, J. A., Hicks, M. D., Simonelli, D. P., Jaumann, R., Clark, R. N., Brown, R. H., Cruikshank, D. P., Momary, T., 2008. Infrared (0.83-5.1 µ m) photometry of Phoebe from the Cassini Visual Infrared Mapping Spectrometer, Icarus, 193, 309- 322, doi: 10.1016/j.icarus.2007.09.014. Burns, J. A., Lamy, P. L., and Soter, S., 1979. Radiation forces on small particles in the solar system, Icarus, 40, 1-48. 34 Carlson, R., Smythe, W., Baines, K., Barbinis, E., Becker, K., Burns, R., Calcutt, S., Calvin, W., Clark, R., Danielson, G., Davies, A., Drossart, P., Encrenaz, T., Fanale, F., Granahan, J., Hansen, G., Herrera, P., Hibbitts, C., Hui, J., Irwin, P., Johnson, T., Kamp, L., Kieffer, H., Leader, F., Lellouch, E., Lopes-Gautier, R., Matson, D., McCord, T., Mehlman, R., Ocampo, A., Orton, G., Roos-Serote, M., Segura, M., Shirley, J., Soderblom, L., Stevenson, A., Taylor, F., Torson, J., Weir, A., Weissman, P., 1996. Near-Infrared Spectroscopy and Spectral Mapping of Jupiter and the Galilean Satellites: Results from Galileo's Initial Orbit, Science, 274, 385-388, doi: 10.1126/science.274.5286.385 Ciarniello, M., Capaccioni, F., Filacchione, G., Clark, R. N., Cruikshank, D. P., Cerroni, P., Coradini, A., Brown, R. H., Buratti, B. J., Tosi, F., Stephan, K., 2011. Hapke modeling of Rhea surface properties through Cassini-VIMS spectra, Icarus, 214, 541-555, DOI:10.1016/j.icarus.2011.05.010. Ciarniello, M., Capaccioni, F., Filacchione, G., 2014. A test of Hapke's model by means of Monte Carlo ray-tracing, Icarus, 237, 293-305, DOI: 10.1016/j.icarus.2014.04.045. Clark, R. N., Brown, R. H., Jaumann, R., Cruikshank, D. P., Nelson, R. M., Buratti, B. J., Mc- Cord, T. B., Lunine, J., Baines, K. H., Bellucci, G., Bibring, J.-P., Capaccioni, F., Cerroni, P., Coradini, A., Formisano, V., Langevin, Y., Matson, D. L., Mennella, V., Nicholson, P. D., Sicardy, B., Sotin, C., Hoefen, T. M., Curchin, J. M., Hansen, G., Hibbitts, K., Matz, K.-D., 2005. Com- positional maps of Saturn's moon Phoebe from imaging spectroscopy, Nature, 435, 66-69, doi: 10.1038/nature03558. Clark, R. N., Curchin, J. M., Jaumann, R., Cruikshank, D. P., Brown, R. H., Hoefen, T. M., Stephan, K., Moore, J. M., Buratti, B. J., Baines, K. H., Nicholson, P. D., Nelson, R. M., 2008. Compositional mapping of Saturn's satellite Dione with Cassini VIMS and implications of dark material in the Saturn system, Icarus, 193, 372-386, doi: 10.1016/j.icarus.2007.08.035. Clark, R. N., Cruikshank, D. P., Jaumann, R., Brown, R. H., Curchin, J. M., Hoefen, T. D., Stephan, K., Buratti, B. J., Filacchione, G., Baines, K. H., Nicholson, P. D., 2012. The Composition of Iapetus: Mapping Results from Cassini VIMS, 2012. Icarus, 218, 831-860, doi:10.1016/j.icarus.2012.01.008 Coradini, A., Tosi, F., Gavrishin, A. I., Capaccioni, F., Cerroni, P., Filacchione, G., Adriani, A., Brown, R. H., Bellucci, G., Formisano, V., D'Aversa, E., Lunine, J. I., Baines, K. H., Bibring, J.-P., Buratti, B. J., Clark, R. N., Cruikshank, D. P., Combes, M., Drossart, P., Jaumann, R., Langevin, Y., Matson, D. L., McCord, T. B., Mennella, V., Nelson, R. M., Nicholson, P. D., Sicardy, B., Sotin, C., Hedman, M. M., Hansen, G. B., Hibbitts, C. A., Showalter, M., Griffith, C., Strazzulla, G., 2008. Identification of spectral units on Phoebe, Icarus, 193, 233-251, doi: 10.1016/j.icarus.2007.07.023. Coradini, A. and 48 colleagues, 2011. The Surface Composition and Temperature of Asteroid 21 Lutetia As Observed by Rosetta/VIRTIS. Science, 334, 492-494, doi: 10.1126/science.1204062 35 Cruikshank, D. P., Dalton, J. B., Ore, C. M. D., Bauer, J., Stephan, K., Filacchione, G., Hendrix, A. R., Hansen, C. J., Coradini, A., Cerroni, P., Tosi, F., Capaccioni, F., Jaumann, R., Buratti, B. J., Clark, R. N., Brown, R. H., Nelson, R. M., McCord, T. B., Baines, K. H., Nicholson, P. D., Sotin, C., Meyer, A. W., Bellucci, G., Combes, M., Bibring, J.-P., Langevin, Y., Sicardy, B., Matson, D. L., Formisano, V., Drossart, P., Mennella, V., 2007. Surface composition of Hyperion, Nature, 448, 54-56, doi: 10.1038/nature05948. Cruikshank, D. P., Wegryn, E., Dalle Ore, C. M., Brown, R. H., Bibring, J.-P., Buratti, B. J., Clark, R. N., McCord, T. B., Nicholson, P. D., Pendleton, Y. J., Owen, T. C., Filacchione, G., Coradini, A., Cerroni, P., Capaccioni, F., Jaumann, R., Nelson, R. M., Baines, K. H., Sotin, C., Bellucci, G.,Combes, M., Langevin, Y., Sicardy, B., Matson, D. L., Formisano, V., Drossart, P., Mennella, V., 2008. Hydrocarbons on Saturn's satellites Iapetus and Phoebe, Icarus, 193, 334-343, doi: 10.1016/j.icarus.2007.04.036. Cruikshank, D. P., Meyer, A. W., Brown, R. H., Clark, R. N., Jaumann, R., Stephan, K., Hibbitts, C. A., Sandford, S. A., Mastrapa, R. M. E., Filacchione, G., Ore, C. M. D., Nicholson, P. D., Buratti, B. J., McCord, T. B., Nelson, R. M., Dalton, J. B., Baines, K. H., Matson, D. L., 2010. Carbon dioxide on the satellites of Saturn: Results from the Cassini VIMS investigation and revisions to the VIMS wavelength scale, Icarus, 206, 561-572, doi: 10.1016/j.icarus.2009.07.012. Cruikshank, D. P., Dalle Ore, C. M., Clark, R. N., Pendleton, Y. J., 2014. Aromatic and aliphatic organic materials on Iapetus: Analysis of Cassini VIMS data, Icarus, 233, 306-315, doi: 10.1016/j.icarus.2014.02.011. Denk, T., Neukum, G., Roatsch, T., Porco, C. C., Burns, J. A., Galuba, G. G., Schmedemann, N., Helfenstein, P., Thomas, P. C., Wagner, R. J., West, R. A., 2010. Iapetus: Unique Sur- face Properties and a Global Color Dichotomy from Cassini Imaging, Science, 327, 435-439, doi: 10.1126/science.1177088 Falenty, A, Kuhs, W. F., 2009. Self-Preservation of CO2 Gas Hydrates Surface Microstructure and Ice Perfection, J. Phys. Chem. B, 113 (49), 15975-15988, doi: 10.1021/jp906859a. Filacchione, G., Capaccioni, F., McCord, T. B. Coradini, A., Cerroni, P., Bellucci, G., Tosi, F., D'Aversa, E., Formisano, V., Brown, R. H., Baines, K. H., Bibring, J. P., Buratti, B. J., Clark, R. N., Combes, M., Cruikshank, D. P., Drossart, P., Jaumann, R., Langevin, Y., Matson, D. L., Mennella, V., Nelson, R. M., Nicholson, P. D., Sicardy, B., Sotin, C., Hansen, G., Hibbitts, K., Showalter, M., Newman, S., 2007. Saturn's icy satellites investigated by Cassini-VIMS. I. Full- disk properties: 350-5100 nm reflectance spectra and phase curves, Icarus, 186, 259-290, doi: 10.1016/j.icarus.2006.08.001. Filacchione, G., Capaccioni, F., Clark, R. N., Cuzzi, J. N., Cruikshank, D. P., Coradini, A., Cerroni, P., Nicholson, P. D., McCord, T. B., Brown, R. H., Buratti, B. J., Tosi, F., Nelson, R. M., 36 Jaumann, R., Stephan, K., 2010. Saturn's icy satellites investigated by Cassini-VIMS. II. Results at the end of nominal mission, Icarus, 206, 507-523, doi: 10.1016/j.icarus.2009.11.006. Filacchione, G., Capaccioni, F., Ciarniello, M., Clark, R. N., Cuzzi, J. N., Nicholson, P. D., Cruik- shank, D. P., Hedman, M. M., Buratti, B. J., Lunine, J. I., Soderblom, L. A., Tosi, F., Cerroni, P., Brown, R. H., McCord, T. B., Jaumann, R., Stephan, K., Baines, K. H., Flamini, E., 2012. Sat- urn's icy satellites and rings investigated by Cassini-VIMS: III - Radial compositional variability, Icarus, 220, 1064-1096, doi: 10.1016/j.icarus.2012.06.040. Filacchione, G., Capaccioni, F., Clark, R. N., Nicholson, P. D., Cruikshank, D. P., Cuzzi, J. N., Lunine, J. I., Brown, R. H., Cerroni, P., Tosi, F., Ciarniello, M., Buratti, B. J., Hedman, M. M., Flamini, E., 2013. The radial distribution of water ice and chromophores across Saturn's system, ApJ, 766, issue 2, article id. 76, doi: 10.1088/0004-637X/766/2/76. Filacchione, G., Ciarniello, M., Capaccioni, F., Clark, R. N., Nicholson, P. D., Hedman, M. M., Cuzzi, J. N., Cruikshank, D. P., Dalle Ore, C. M., Brown, R. H., Cerroni, P., Altobelli, N., Spilker, L. J., 2014. Cassini-VIMS observations of Saturn's main rings: I. Spectral properties and temperature radial profiles variability with phase angle and elevation, Icarus, 241, 45-65, doi: 10.1016/j.icarus.2014.06.001. Fink, U., Larson, H. P., 1975. Temperature dependence of the water ice spectrum between 1 and 4microns: application to Europa, Ganymede and Saturn's rings. Icarus, 24, 411-420. Flasar, F. M., Kunde, V. G., Abbas, M. M., Achterberg, R. K., Ade, P., Barucci, A., B´ezard, B., Bjoraker, G. L., Brasunas, J. C., Calcutt, S., Carlson, R., C´esarsky, C. J., Conrath, B. J., Coradini, A., Courtin, R., Coustenis, A., Edberg, S., Edgington, S., Ferrari, C., Fouchet, T., Gautier, D., Gierasch, P. J., Grossman, K., Irwin, P., Jennings, D. E., Lellouch, E., Mamoutkine, A. A., Marten, A., Meyer, J. P., Nixon, C. A., Orton, G. S., Owen, T. C., Pearl, J. C., Prang´e, R., Raulin, F., Read, P. L., Romani, P. N., Samuelson, R. E., Segura, M. E., Showalter, M. R., Simon-Miller, A. A., Smith, M. D., Spencer, J. R., Spilker, L. J., Taylor, F. W., 2004. Exploring The Saturn System In The Thermal Infrared: The Composite Infrared Spectrometer, Space Science Reviews,15, 1-4, 169-297. doi: 10.1007/s11214-004-1454-9. Fletcher, N. H., 1970. The Chemical Physics of Ice, Cambridge University Press. Goguen, J. D., Buratti, B. J., Brown, R. H., Clark, R. N., Nicholson, P. D., Hedman, M. M., Howell, R. R., Sotin, C., Cruikshank, D. P., Baines, K. H., Lawrence, K. J., Spencer, J. R., Blackburn, D. G., 2013. The temperature and width of an active fissure on Enceladus measured with Cassini VIMS during the 14 April 2012 South Pole flyover, Icarus, 226, 1128-1137, doi: 10.1016/j.icarus.2013.07.012. 37 Gombosi, T. I., Armstrong, T. P., Arridge, C. S., Khurana, K. K., Krimigis, S. M., Krupp, N., Persoon, A. N., Thomsen, M. F., 2009. Saturn's magnetospheric configuration, Saturn after Cassini/Huygens, 203-256, Springer, doi:10.1007/978-1-4020-9215-2. Grundy, W. M., Buie, M. W., Stansberry, J. A., Spencer, J. R., Schmitt, B., 1999. Near infrared spectra of icy outer solar system surfaces: remote determination of H2O ice temperatures. Icarus 142, 536-549. Grundy, W. M., Schmitt, B., 1998. The temperature dependent near infrared absorption spectrum of hexagonal H2O ice. Journal of Geophysical Research, 103, 25809-25822. Hamilton, D. P., Skrutskie, M. F., Verbiscer, A. J., Masci, F. J. 2015. Small particles dominate Sat- urn?s Phoebe ring to surprisingly large distances, Nature, 522, 185-187, doi: 10.1038/nature14476. Hansen, C. J., Esposito, L., Stewart, A. I. F., Colwell, J., Hendrix, A. Pryor, W., She- mansky, D., West, R., 2006. Enceladus' Water Vapor Plume, Science, 311, 1422-1425, doi:10.1126/science.1121254 Hobbs, P. V., 1974. Ice Physics, Clarendon Press. Oxford. Hedman, M. M., Gosmeyer, C. M., Nicholson, P. D., Sotin, C., Brown, R. H., Clark, R. N., Baines, K. H., Buratti, B. J., Showalter, M. R., 2013. An observed correlation between plume activity and tidal stresses on Enceladus. Nature, 500, 182-184, doi: 10.1038/nature12371. Howett, C. J. A., Spencer, J. R., Pearl, J., Segura, M., 2010. Thermal inertia and bolometric Bond albedo values for Mimas, Enceladus, Tethys, Dione, Rhea and Iapetus as derived from Cassini/CIRS measurements. Icarus, 206, 573-593, doi: 10.1016/j.icarus.2009.07.016. Howett, C. J. A., Spencer, J. R., Schenk, P., Johnson, R. E., Paranicas, C., Hurford, T. A., Verbiscer, A., Segura, M., 2011. A high-amplitude thermal inertia anomaly of probable magnetospheric origin on Saturn's moon Mimas. Icarus, 216, 221-226, doi: 10.1016/j.icarus.2011.09.007. Howett, C. J. A., Spencer, J. R., Pearl, J., Segura, M., 2011. High heat flow from Ence- ladus' south polar region measured using 10-600 cm−1 Cassini/CIRS data, JGR, 116, E03003, doi:10.1029/2010JE003718. Howett, C. J. A., Spencer, J. R., Hurford, T., Verbiscer, A., Segura, M., 2012. Pac- Man returns: An electron-generated thermal anomaly on Tethys, Icarus, 221, 1084-1088, doi:10.1016/j.icarus.2012.10.013. Howett, C. J. A., Spencer, J. R., Hurford, T., Verbiscer, A., Segura, M., 2014. Thermophysical property variations across Dione and Rhea. Icarus, 241, 239-247, doi: 10.1016/j.icarus.2014.05.047. 38 Imanaka, H., Cruikshank, D.P., Khare, B.N., McKay, C.P., 2012. Optical constants of Titan tholins at mid-infrared wavelengths (2.5-25 µm) and the possible chemical nature of Titan's haze particles, Icarus, 218, 247-261. Jaumann, R., Stephan, K., Brown, R. H., Buratti, B. J., Clark, R. N., McCord, T. B., Coradini, A., Capaccioni, F., Filacchione, G., Cerroni, P., Baines, K. H., Bellucci, G., Bibring, J.-P., Combes, M., Cruikshank, D. P., Drossart, P., Formisano, V., Langevin, Y., Matson, D. L., Nelson, R. M., Nicholson, P. D., Sicardy, B., Sotin, C., Soderbloom, L. A., Griffith, C., Matz, K.-D., Roatsch, Th., Scholten, F., Porco, C. C., 2006. High-resolution CASSINI-VIMS mosaics of Titan and the icy Saturnian satellites, Planetary and Space Science, 54, 1146-1155, doi: 10.1016/j.pss.2006.05.034. Jaumann, R., Stephan, K., Hansen, G. B., Clark, R. N., Buratti, B. J., Brown, R. H., Baines, K. H., Newman, S. F., Bellucci, G., Filacchione, G., Coradini, A., Cruikshank, D. P., Griffith, C. A., Hibbitts, C. A., McCord, T. B., Nelson, R. M., Nicholson, P. D., Sotin, C., Wagner, R., 2008. Distribution of icy particles across Enceladus' surface as derived from Cassini-VIMS measurements. Icarus, 193, 407-419, doi: 10.1016/j.icarus.2007.09.013. Jaumann, R., Clark, R. N., Nimmo, F., Hendrix, A. R., Buratti, B. J., Denk, T., Moore, J. M., Schenk, P. M., Ostro, S. J., Srama, R., 2009. Icy Satellites: Geological Evolution and Surface Processes, in Saturn from Cassini-Huygens, edited by M.K. Dougherty, L.W. Esposito, and S.M. Krimigis, Springer, Berlin, 459-509, doi: 10.1007/978-1-4020-9217-6 20. Kempf, S., Beckmann, U, Schmidt, J., (2010). How the Enceladus dust plume feeds Saturn?s E ring, Icarus, 206, 446-457, doi: 10.1016/j.icarus.2009.09.016. Kuiper, G.P., 1957. Infrared observations of planets and satellites. Astronomical Journal, 62, 295. Le Gall, A., Leyrat, C., Janssen, M. A., Keihm, S., Wye, L. C., West, R., Lorenz, R. D., 2014. Iapetus' near surface thermal emission modeled and constrained using Cassini RADAR Radiometer microwave observations, Icarus, 241, 221-238, doi: 10.1016/j.icarus.2014.06.011. Leto, G., Gomis,O., Strazzulla,G., 2005. The reflectance spectrum of water ice: is the 1.65 µm peak a good temperature probe? Memorie della Societ`a Astronomica Italiana Supplement, 6, 57. Mastrapa, R., Bernstein, M., Sandford, S., Roush, T., Cruikshank, D., Dalle Ore, C., 2008. Optical constants of amorphous and crystalline H2O-ice in the near infrared from 1.1 to 2.6 µm. Icarus, 197, 307-320, doi: 10.1016/j.icarus.2008.04.008. Mastrapa, R. M., Sandford, S. A., Roush, T L., Cruikshank, D. P., Dalle Ore, C M., 2009. Optical constants of amorphous and crystalline H2O-ice: 2.5 - 22 µm (4000 - 455 cm−1) optical constants of H2O-ice. Astrophysical Journal, 701, 1347-1356, doi: 10.1088/0004-637X/701/2/1347. Moore, M. H., Donn, B., Khanna, R., A'Hearn, M. F., 1983. Studies of proton-irradiated cometary- type ice mixtures, Icarus, 54, 388-505, doi: 10.1016/0019-1035(83)90236-1. 39 Nimmo, F., Spencer, J. R., Pappalardo, R. T., Mullen, M. E., 2007. Shear heating as the origin of the plumes and heat flux on Enceladus. Nature. 447, 289?291, doi:10.1038/nature05783 Oancea, A., Grasset, O., Le Menn, E., Bollengier, O., Bezacier, L., Le Mouelic, S., Tobie, G., 2012. Laboratory infrared reflection spectrum of carbon dioxide clathrate hydrates for astrophysical remote sensing applications, Icarus, 221, 900-910, doi: 10.1016/j.icarus.2012.09.020. Paranicas, C., Mitchell, D. G., Krimigis, S. M., Carbary, J. F., Brandt, P. C., Turner, F. S., Rous- sos, E., Krupp, N., Kivelson, M. G., Khurana, K. K., Cooper, J. F., Armstrong, T. P., Burton, M., 2010a. Asymmetries in Saturn's radiation belts, JGR, 115, Issue A7, CiteID A07216, doi: 10.1029/2009JA014971. Paranicas, C., Mitchell, D. G., Roussos, E., Kollmann, P., Krupp, N., Muller, A. L., Krimigis, S. M., Turner, F. S., Brandt, P. C., Rymer, A. M., Johnson, R. E., (2010b). Transport of en- ergetic electrons into Saturn's inner magnetosphere, JGR, 115, Issue A9, CiteID A09214, doi: 10.1029/2010JA015853. Pinilla-Alonso, N., Roush, T. L., Marzo, G., Cruikshank, D. P., Dalle Ore, C. M., 2011. Iapetus surface variability revealed from statistical clustering of a VIMS mosaic: the distribution of CO2. Icarus, 215, 75-82, doi: 10.1016/j.icarus.2011.07.004. Pitman, K. M., Buratti, B. J., Mosher, J. A., 2010. Disk-integrated bolometric Bond albedos and rotational light curves of saturnian satellites from Cassini Visual and Infrared Mapping Spectrom- eter. Icarus, 206, 537-560, doi: 10.1016/j.icarus.2009.12.001. Porco, C. C., Helfenstein, P., Thomas, P. C., Ingersoll, A. P., Wisdom, J., West, R., Neukum, G., Denk, T., Wagner, R., Roatsch, T., Kieffer, S., Turtle, E., McEwen, A., Johnson, T. V., Rathbun, J., Veverka, J., Wilson, D., Perry, J., Spitale, J., Brahic, A., Burns, J. A., DelGenio, A. D., Dones, L., Murray, C. D., Squyres, S., 2006. Cassini Observes the Active South Pole of Enceladus, Science, 311, 1393-1401, doi:10.1126/science.1123013 Porco, C., DiNino, D., Nimmo, F., 2014. How the Geysers, Tidal Stresses, and Thermal Emis- sion across the South Polar Terrain of Enceladus are Related, AJ, 148, 45, doi:10.1088/0004- 6256/148/3/45. Postberg, F., Schmidt, J., Hillier, J., Kempf, S., Srama, R., 2011. A salt-water reservoir as the source of a compositionally stratified plume on Enceladus, Nature, 474, 620-622, doi:10.1038/nature10175. Sandford, S. A.; Allamandola, L. J., 1990. The physical and infrared spectral properties of CO2 in astrophysical ice analogs, ApJ, 355, 357-372, doi: 10.1086/168770. Schenk, P., Hamilton, D. P., Johnson, R. E., McKinnon, W. B., Paranicas, C., Schmidt, J., Showalter, M. R., 2011. Plasma, plumes and rings: Saturn system dynamics as recorded in global color patterns on its midsize icy satellites. Icarus, 211, 740-757, doi:10.1016/j.icarus.2010.08.016. 40 Scipioni, F., Tosi, F., Stephan, K., Filacchione, G., Ciarniello, M., Capaccioni, F., Cerroni, P. and the VIMS Team, 2013. Spectroscopic classification of icy satellites of Saturn I: Identification of terrain units on Dione, Icarus, 226, 1331-1349, doi: 10.1016/j.icarus.2013.08.008. Scipioni, F., Tosi, F., Stephan, K., Filacchione, G., Ciarniello, M., Capaccioni, F., Cerroni, P. and the VIMS Team, 2014. Spectroscopic classification of icy satellites of Saturn II: Identification of terrain units on Rhea, Icarus, 234, 1-16, doi: 10.1016/j.icarus.2014.02.010. Spencer, J. R., Pearl, J. C., Segura, M., Flasar, F. M., Mamoutkine, A., Romani, P., Buratti, B. J., Hendrix, A. R., Spilker, L. J., Lopes, R. M. C., 2006. Cassini Encounters Enceladus: Background and the Discovery of a South Polar Hot Spot, Science, 311, 1401-1405, doi:10.1126/science.1121661. Spencer, J. R., Barr, A. C., Esposito, L. W., Helfenstein, P., Ingersoll, A. P., Jaumann, R., McKay, C. P., Nimmo, F., Hunter Waite, J., 2009. Enceladus: an active cryovolcanic satellite, in Saturn from Cassini-Huygens , edited by M.K. Dougherty, L.W. Esposito, and S.M. Krimigis, Springer, Berlin, 683-724. ISBN 978-1-4020-9216-9. Spencer, J. R., Denk, T., 2010. Formation of Iapetus? Extreme Albedo Dichotomy by Exogenically Triggered Thermal Ice Migration, Science, 327, 432. Stephan, K., Jaumann, R., Wagner, R., Clark, R. N., Cruikshank, D. P., Hibbitts, C. A. Roatsch, T., Hoffmann, H., Brown, R. H., Filacchione, G., Buratti, B. J., Hansen, G. B., McCord, T. B., Nicholson, P. D., Baines, K. H., 2010. Dione' s spectral and geological properties, Icarus, 206, 631-652, DOI: 10.1016/j.icarus.2009.07.036. Stephan, K., Jaumann, R., Wagner, R., Clark, R. N., Cruikshank, D. P., Giese, B., Hibbitts, C. A., Roatsch, T., Matz, K.-D., Brown, R. H., Filacchione, G., Capaccioni, F., Scholten, F., Buratti, B. J., Hansen, G. B., Nicholson, P. D., Baines, K. H., Nelson, R. M., Matson, D. L., 2012. The saturnian satellite Rhea as seen by Cassini VIMS. Planetary and Space Science, 61, 142-160, DOI: 10.1016/j.pss.2011.07.019. Taffin, C., Grasset, O., LeMenna, E., Bollengier, O., Giraud, M., LeMouelic, S., 2012. Temper- ature and grain size dependence of near-IR spectral signature of crystalline water ice: From lab experiments to Enceladus' south pole, Planetary and Space Science, 61, 124-134, doi: 10.1016/j.pss.2011.08.015 Tamayo, D., Burns, J. A., Hamilton, D. P., Hedman, M. M., 2011. Finding the trigger to Iapetus' odd global albedo pattern: Dynamics of dust from Saturn's irregular satellites, Icarus, 215, 260-278, doi: 10.1016/j.icarus.2011.06.027. Tosi, F., Turrini, D., Coradini, A., Filacchione, G. and the VIMS Team, 2010. Probing the origin of the dark material on Iapetus, MNRAS, 403, 1113-1130, doi: 10.1111/j.1365-2966.2010.16044.x. 41 Verbiscer, A.J., Veverka, J., 1989. Albedo dichotomy of Rhea: Hapke analysis of Voyager photometry, Icarus, 82, 336-353. Verbiscer, A. J., Skrutskie, M. F., Hamilton, D. P., 2009. Saturn's largest ring, Nature, 461, 1098- 1100, doi: 10.1038/nature08515. Waite, J. H., Combi, M. R., Ip, W.-H., Cravens, T. E., McNutt, R. L., Kasprzak, W., Yelle, R., Luh- mann, J., Niemann, H., Gell, D. Magee, B., Fletcher, G., Lunine, J., Tseng, W.-L., 2006. Cassini Ion and Neutral Mass Spectrometer: Enceladus Plume Composition and Structure, Science, 311, 1419-1422, doi:10.1126/science.1121290. Warren, S. G.,1984. Optical constants of ice from the ultraviolet to the microwave, Applied Optics, 23,1206-1225. Zubko, V. G., Mennella, V., Colangeli, L., Bussoletti, E., 1996. Optical constants of cosmic carbon analogue grains I. Simulation of clustering by a modified continuous distribution of ellipsoids. Mon. Not. R. Astron. Soc., 282, 1321-1329. 42
0902.2063
1
0902
2009-02-12T15:12:51
Detection and Characterization of Planetary Systems with $\mu$as Astrometry
[ "astro-ph.EP" ]
Astrometry as a technique has so far proved of limited utility when employed as either a follow-up tool or to independently search for planetary mass companions orbiting nearby stars. However, this is bound to change during the next decade. In this review, I start by summarizing past and present efforts to detect planets via milli-arcsecond astrometry. Next, I provide an overview of the variety of technical, statistical, and astrophysical challenges that must be met by future ground-based and space-borne efforts in order to achieve the required degree of astrometric measurement precision. Then, I discuss the planet-finding capabilities of future astrometric observatories aiming at micro-arcsecond precision, with a particular focus on their ability to fully describe multiple-component systems. I conclude by putting astrometry in context, illustrating its potential for important contributions to planetary science, as a complement to other indirect and direct methods for the detection and characterization of planetary systems.
astro-ph.EP
astro-ph
Title : will be set by the publisher Editors : will be set by the publisher EAS Publications Series, Vol. ?, 2018 9 0 0 2 b e F 2 1 . ] P E h p - o r t s a [ 1 v 3 6 0 2 . 2 0 9 0 : v i X r a DETECTION AND CHARACTERIZATION OF PLANETARY SYSTEMS WITH µAS ASTROMETRY A. Sozzetti 1 Abstract. Astrometry as a technique has so far proved of limited utility when employed as either a follow-up tool or to independently search for planetary mass companions orbiting nearby stars. However, this is bound to change during the next decade. In this review, I start by summarizing past and present efforts to detect planets via milli- arcsecond astrometry. Next, I provide an overview of the variety of technical, statistical, and astrophysical challenges that must be met by future ground-based and space-borne efforts in order to achieve the required degree of astrometric measurement precision. Then, I dis- cuss the planet-finding capabilities of future astrometric observatories aiming at micro-arcsecond precision, with a particular focus on their ability to fully describe multiple-component systems. I conclude by putting astrometry in context, illustrating its potential for important contributions to planetary science, as a complement to other indirect and direct methods for the detection and characterization of planetary systems. 1 Introduction The present-day catalog of extrasolar planets1 includes a panoply of systems con- taining more than one planetary companion, up to a maximum of five in the case of the 55 Cnc system (Fischer et al. 2007). The astounding diversity of multi- ple systems encompasses some extreme configurations discovered by means of the light-time-travel technique around pulsars (Wolszczan & Frail 1992), post-AGB stars (Silvotti et al. 2007), SdB+M dwarf eclipsing binaries (Lee et al. 2009), or by direct imaging of young stars and their dusty disks (Kalas et al. 2008; Marois et al. 2008). However, decade-long, high-precision (1-3 m s−1) radial-velocity surveys have so far contributed the bulk of the well-characterized multiple-planet 1 INAF - Osservatorio Astronomico di Torino, Strada Osservatorio, 20 - 10025 Pino Torinese Italy 1See for example Jean Schneider's Extrasolar Planet Encyclopedia at http://exoplanet.eu c(cid:13) EDP Sciences 2018 DOI: (will be inserted later) 2 Title : will be set by the publisher systems orbiting normal stars in the solar neighborhood known to-date. Their intriguing properties provide fundamental clues and insights for improved under- standing of the complex processes of planetary systems formation, early orbital evolution, and long-term dynamical interaction. For example, multiple systems appear to be quite common. Recent estimates indicate that ∼ 30% of stars with planets have more than one planetary mass companion (integrating over all spec- tral types and for nearby dwarfs within 200 pc). Planetary systems exhibit great dynamical diversity, with several families identified, which include hierarchical sys- tems, secularly interacting systems, and systems in mean motion resonances (e.g., Go´zdziewski et al. 2008, and references therein). Most recently, the HD 45364 two-planet systems was shown by Correia et al. (2009) to be locked in a 3:2 res- onance, analogous to the one formed by Neptune and Pluto in our Solar System. Planetary systems appear to have different orbital elements distribution functions with respect to those of single-planet systems. In addition, distributions for sys- tems containing only low-mass (Neptune and Super-Earths) planets may also differ from those of systems containing gas giants. Finally, there are hints that planet frequency fp may also be a different function of the host star's properties (M⋆, [Fe/H]) in single- and multiple-planet systems (for comprehensive summaries of the characteristics of multiple-planet systems detected by radial-velocity surveys see for example Wright et al. (2009) and the reviews by J. Wright and S. Udry in this volume). The observational data on multiple systems have important im- plications for the proposed models of formation and early evolution of planetary systems (for reviews see for example Lissauer & Stevenson (2007), Durisen et al. (2007), Nagasawa et al. (2007), and references therein), provide important clues on the relative role of several proposed mechanisms of dynamical interactions be- tween forming planets, gaseous/planetesimals disks, and distant companion stars (for reviews see for example Papaloizou et al. (2007), Levison et al. (2007), Ford & Rasio (2008), and references therein), and allow to measure the likelihood of formation and survival of terrestrial planets in the Habitable Zone2 of the parent star (Menou & Tabachnik 2003; Jones et al. 2005; Hinse et al. 2008, and references therein). Multiple-planet systems are thus clearly excellent laboratories to search for fossil evidence of formation and dynamical evolution mechanisms. However, given the present limitations in our ability to elucidate in a unified manner the various phases of the complex processes of planet formation and evolution, some of the key questions on the physical characterization and architecture of planetary systems (how many dynamical families can be identified? Are their orbits coplanar? What is the origin of their eccentricities? Are the parameters' distribution functions and fp(M⋆, [Fe/H]) actually different for single- and multiple-planet systems?) still await a clear answer. To this end, help from future data, obtained with a variety of techniques, over a wide range of observing wavelengths, both from the ground 2For any given star, the region of habitability is defined as the range of orbital distances at which a potential water reservoir, the primary ingredient for the development of a complex biology, would be found in liquid form (e.g., Kasting et al. 1993) A. Sozzetti: Planetary Systems with µas Astrometry 3 and in space, will prove invaluable. As for the most successful of indirect detection techniques, Doppler surveys are extending their time baseline and/or are achieving higher velocity precision (≤ 1 m s−1, see for example Pepe & Lovis 2008), to continue searching for planets at in- creasingly larger orbital distances (e.g., Wright et al. 2007) and with increasingly smaller masses (e.g., Mayor et al. 2008). During the next decade, RV surveys will get close to answering the hot question of how common are planetary systems with architectures similar to the Solar System. Ultimately, the limiting factor may not be the intrinsic stability of new-generation spectrographs, but rather the primary stars themselves, through astrophysical noise sources such as stellar surface activ- ity, rotation, and acoustic p-modes. These problems are already limiting severely Doppler surveys from investigating the existence of giant planets orbiting stars significantly departing from our Sun in age, mass, and metal content. While ground-based wide-field photometric transit surveys are allowing us to unveil fundamental properties of strongly irradiated giant planets (Charbonneau et al. 2007, and references therein), the Kepler (Borucki et al. 2003) and CoRoT (Baglin et al. 2002) missions are designed to photometrically detect transiting Earth-sized planets in the Habitable Zone of solar-type host stars, providing the first measure of the occurrence of rocky planets and ice-giants. The exquisite photometric precision and the long time baseline of the measurements will allow to look for additional companions in systems where one planet is found to be transiting, thanks to the possible detection of tiny variations in the predicted time of transit center induced by the gravitational perturbation of one or more outer companions, not necessarily transiting (for a review of the transit timing method and its potential see M. Holman, this volume). However, the host stars will reside at typical distances beyond 250 pc, making imaging and spectroscopic follow-up of the planets difficult. The prospects for detailed characterization of giant planets and Super Earths transiting nearby solar-type as well as cool stars are tied to the approval of proposed all-sky surveys in space (e.g., TESS), and to the possible success of upcoming ground-based photometric searches for transiting rocky planets aroud M dwarfs (e.g., MEarth). Gravitational microlensing surveys from the ground within the next decade have the potential to deliver a complete census of the cold planet population down to ∼ 10 M⊕ orbiting low-mass stars at separations a > 1.5 AU. Proposed microlensing observatories in space (e.g., MPF) could extend the census to plan- ets of ∼ 1 M⊕ with separations exceeding 1 AU (e.g., Gaudi 2007, and refer- ences therein). We note however that observations with this technique are non- reproducible and follow-up analyses are virtually impossible (the detected systems typically residing at over 1 kpc from the Sun), thus such findings will mostly have statistical value but will help little toward the physical characterization of planetary systems. During the next twenty years, the prospects are becoming increasingly "bright" for the direct detection of exoplanets and the spectroscopic characterization of their atmospheres using techniques to spatially or temporally separate them from their parent stars. Data from upcoming and proposed observatories (for a review see for 4 Title : will be set by the publisher example Beuzit et al. 2007, and references therein) for visible-light, near- and mid- infrared imaging and spectroscopy and equipped with single- and multiple-aperture telescopes from the ground (e.g., VLT/SPHERE, ELT/EPICS) and in space (e.g., JWST, SPICA, Darwin and TPF-C/I avatars) will completely transform our view of the nature of planetary systems. In this review paper I will focus on what the contribution of astrometry from the ground and in space will be to the astrophysics of planetary systems. I will provide an historical perspective on past efforts to detect planets with astrometry, I will address some of the most relevant challenges to be faced in the transition from milli-arcsecond (mas) to micro-arcsecond (µas) astrometry, and I will conclude with a discussion on future prospects, by putting this technique in perspective with other planet-detection methods. 2 Blunders and Successes of mas Astrometry Similarly to the spectroscopic technique, astrometric measurements can detect the stellar wobble around the system barycenter due to the gravitational perturbation of nearby planets. The main observable (assuming circular orbits) is the 'astro- metric signature', i.e. the apparent semi-major axis of the stellar orbit: α = (cid:18) Mp M⊙(cid:19)(cid:18) M⊙ M⋆ (cid:19)(cid:16) ap 1 AU(cid:17)(cid:16) pc d (cid:17) arcsec (2.1) However, by reconstructing the orbital motion in the plane of the sky, astrometry alone can determine the entire set of seven orbital elements, thus breaking the Mp sin i degeneracy intrinsic to Doppler measurements and allowing to derive an actual mass estimate for the companion. In multiple systems, astrometric mea- surements can determine the mutual inclination angle between pairs of planetary orbits: cos irel = cos iin cos iout + sin iin sin iout cos(Ωout − Ωin), (2.2) where iin and iout, Ωin and Ωout are the inclinations and lines of nodes of the inner and outer planet, respectively. Thus, meaningful estimates can be obtained of the full three-dimensional geometry of any planetary system, without restrictions on the orbital alignment with respect to the line of sight. 2.1 Ground-Based Astrometry Shortly after the end of World War II, Otto Struve (1952) had already summa- rized the merit of searching for planets using precision radial-velocities, transit photometry, and astrometry. In his own words, "one of the burning questions of astronomy deals with the frequency of planet-like bodies in the galaxy which belong to stars other than the Sun". At that time, interest in the problem had been stimulated by the pre-war 'discoveries' of planet-like companions around 61 Cygni, and 70 Ophiuchi by Strand (1943), and Reuyl & Holmberg (1943), who A. Sozzetti: Planetary Systems with µas Astrometry 5 had presented astrometric measurements based on long-term time-series of photo- graphic plates. These announcements had indeed been interpreted as supporting evidence for proposed theories on the origin of the Solar System (Alfv´en 1943) and speculations on the frequency of planetary systems in space (Jeans 1943). The evidence for planetary companions detected by ground-based astrometry around 61 Cyg and 70 Oph has long been proved incorrect (e.g., Heintz 1978). These two early examples are not the only ones. The two most famous decades- long "quarrels" concern Barnard's Star and Lalande 21185. The best-known is the long-term effort to detect planets around Barnard's star by van de Kamp (1963, 1969a, 1969b, 1975, 1982). The claim was not confirmed by Gatewood & Eichhorn (1973), while Hershey (1973), Heintz (1976), and finally Croswell (1988) showed that van de Kamps "detections" were the likely result of instrumental errors. As for the planetary companion to Lalande 21185, the announcement by Lippincott (1960a, 1960b) and Hershey & Lippincott (1982) was initially not confirmed by Gatewood (1974) and Gatewood et al. (1992), while in the most recent contribution to the subject Gatewood (1996) claimed instead one or even two planets could possibly be orbiting the star. In both these cases, the jury is still out. 2.2 Hipparcos Astrometry Prompted by the success of Doppler surveys for giant planets of nearby stars and by the need to find a method to break the Mp − i degeneracy intrinsic to radial-velocity measurements, several authors have re-analyzed in recent years the Hipparcos Intermediate Astrometric Data (IAD), in order to either detect the planet-induced stellar astrometric motion of the bright hosts, most of which had been observed by the satellite, or place upper limits to the magnitude of the perturbation, in the case of no detections. The Hipparcos IAD have been re- processed alone, or in combination with either the spectroscopic information or with additional ground-based astrometric measurements. The Hipparcos IAD have been used to put upper limits on the size of the astro- metric perturbations by Perryman et al. (1996) and by Zucker & Mazeh (2001), who could rule out at the ∼ 2-σ level the hypothesis of low-mass stellar com- panions disguised as planets for over two dozen objects. Preliminary astrometric masses for ∼ 30 Doppler-detected planets were announced about a decade ago by several authors (Mazeh et al. 1999; Zucker & Mazeh 2000; Gatewood et al. 2001; Han et al. 2001). On the one hand, the suprising conclusion of these works is that a significant fraction (∼ 40%) of the planet candidates are instead stars, and the remainder sub-stellar companions are in most cases brown dwarfs rather than planets. The results stem from the derivation of a vast majority of quasi-face-on orbits, i.e. with i on the order of only a few degrees. On the other hand, Pourbaix (2001), Pourbaix & Arenou (2001), and later Zucker & Mazeh (2001) demonstrated that the statistical significance of most of the Hipparcos astrometric orbits is ques- tionable at best, and that the systematically very small inclination angles can arise as an artifact of the fitting procedures utilized to dig out signals below the noise 6 Title : will be set by the publisher level of the Hipparcos data. Interestingly enough, the only system with an Hip- parcos orbit deemed statistically acceptable by Pourbaix & Arenou (2001) which corresponded to an M-dwarf companion, CrB (Gatewood et al. 2001), was then called into question by recent high-resolution infrared spectroscopic measurements (Bender et al. 2005). Successful Hipparcos astrometric orbital solutions for a few Doppler-detected systems containing companions with minimum masses close to and slight above the dividing line between planets and brown dwarfs have recently obtained by Reffert & Quirrenbach (2006) and Sozzetti & Desidera (2009). 2.3 HST/FGS Astrometry A firm (3-5 σ level) upper mass limit of ∼ 30 MJ on the mass of a spectroscopically detected extrasolar planet was placed by McGrath et al. (2002), who failed to re- veal astrometric motion of the Mp sin i = 0.88 MJ object on a 14.65-day orbit in the 1 Cnc multiple-planet system using narrow-field relative HST/FGS astrometry. Finally, the first undisputed value of the actual mass of a Doppler-detected planet was obtained by Benedict et al. (2002) who derived the perturbation size, inclina- tion angle, and mass of the outer companion in the multiple-planet system GJ 876 from a combined fit to HST/FGS astrometry and high-precision radial-velocities. In recent years, HST/FGS astrometry has helped determining the actual mass of the Neptune-mass planet in the 1 Cnc system, under the assumption of copla- narity (McArthur et al. 2004), has allowed to measure the mass of the long-period planet in orbit around ε Eri (Benedict et al. 2006), and has permitted to reveal the nature of the companion to HD 33636 as an M dwarf rather than a massive planet (Bean et al. 2007). Until the advent of µas-level precision astrometric facilities from the ground and in space, mas astrometry with HST/FGS will continue to deliver important results. Data have already been gathered and are being analyzed in order to determine the actual masses of the Doppler-detected planets HD 47536b, HD 136118b, HD 168443c, HD 145675b, and HD 38529c (Benedict et al. 2008). Furthermore, HST/FGS observations are also being collected for a handful of multiple systems, including HD 128311, HD 2020206, µ Ara, γ Cep and υ And, with the aim of measuring directly the degree of coplanarity among detectable components. Most recently, the first coplanarity test has been carried out by Bean & Seifahrt (2009) in the case of the two outer planets in the GJ 876 systems using a combination of Doppler measurements and HST/FGS astrometry, albeit with the crucial help of a priori dynamical considerations. 3 The Challenges of µas Astrometry The state-of-the-art astrometric precision is nowadays set to ∼ 1 mas by Hipparcos and HST/FGS (see § 3.5). By looking at Eq. 2.1, one realizes how the magnitude of the perturbation induced by a 1 Jupiter-mass planet in orbit at 5 AU around a 1-M⊙ star at 10 pc from the Sun is α ≃ 500 µas. For the same distance and primary mass, a 'hot Jupiter' with ap = 0.01 AU induces α = 1 µas, and an A. Sozzetti: Planetary Systems with µas Astrometry 7 Earth-like planet (ap = 1 AU) causes a perturbation α = 0.33 µas. One then understands why astrometric measurements with mas precision have so far proved of limited utility when employed as either a follow-up tool or to independently search for planetary mass companions orbiting nearby stars. Indeed, µas astrome- try is almost coming of age, provided the demanding technological and calibration requirements to achieve the required level of measurement precision are met. Sozzetti (2005) has provided a review of methods and instrumentation to detect and characterize planetary systems with astrometry. I summarize here the main points, focusing in particular on the challenges inherent to correctly modeling astrometric measurements in which multiple planetary signals are present. 3.1 Correcting for Astrophysical Effects With the goal of achieving µas-level precision, astrometric observations may have to be corrected first for a variety of effects that modify the apparent position of the target. These can be classical in nature or intrinsically relativistic, and can be due to a) the motion of the observer (e.g., aberration), b) secular variations in the target space motion with respect to the observer (e.g., perspective acceleration), or c) the gravitational fields of massive bodies in the vicinity of the observer (light deflection). Taking into account such effects is particularly important for global astrometric measurements such as those that will be carried out by Gaia (less so for differential measurements). Indeed, several attempts have been made in the past years (Brumberg 1991; Klioner & Kopeikin 1992; de Felice et al. 1998, 2001, 2004, 2006; de Felice & Preti 2006, 2008; Vecchiato et al. 2003; Klioner 2003, 2004; Crosta et al. 2003) to develop schemes for the reduction of astrometric observations at the µas precision level directly within the framework of General Relativity, either employing non-perturbative approaches or the PPN formalism (Will 1993). A model of relativistic astrometry based on the Klioner (2003) PPN formulation is considered the baseline for the reduction of Gaia data by the Gaia Data Processing and Analysis Consortium. 3.2 Noise Sources The astrometric measurement process is carried out in the presence of observa- tional error sources (both random and systematic) which depend on the mode of operation (wide-, narrow-angle, or global astrometry), operational wavelength (visible or near-infrared) and the instrument (monolithic or diluted configuration) used to carry out the measurements. As for instrumental errors, for single-dish architectures the most technologically challenging to deal with will be the ability to achieve location accuracies of ∼ 1/1000 of a pixel for CCD detectors, and the capability to minimize geometric distortions of optical systems. The diffraction-limited image quality afforded by adaptive optics systems modifies the relative importance of these error terms to some degree (Cameron et al. 2009). For interferometers, both accuracies of tens of picometers on the position of the delay lines and positional stabilities of of 8 Title : will be set by the publisher ∼ 10 nm on internal optical pathlengths, in order to ensure maintenance of the fringe visibility, will have to be achieved (see Sozzetti 2005, and references therein). For both architectures, non-uniform system throughput due to the time evolution of the optical parameters (Gai & Cancelliere 2008) could also induce significant degradation in the achievable astrometric precision. For ground-based instrumentation, the atmosphere constitutes an additional source of noise through both its turbulent layers (a random component) and due to the differential chromatic refraction (DCR) effect (a systematic component). The limitations due to the former effect can be overcome to a significant degree utilizing diluted rather than monolithic configurations (Shao & Colavita 1992). The latter has been proven by several studies to be often the predominant limitation to precision astrometry from the ground (e.g., Monet al. 1992; Lazorenko 2006). For space-borne observatories, additional random and systematic error sources must be taken into account, which are introduced by the satellite operations and environment. The most relevant class of uncertainties can be broadly defined in terms of attitude errors (induced by solar wind, micrometeorites, particle radi- ation, radiation pressure, thermal drifts and spacecraft jitter), for which ad hoc calibration procedures must be implemented case by case. In § 3.5 I describe in some detail how these problems have been approached in the past, and how they are being addressed for future programs. Finally, for µas-level astrometric precision several 'astrophysical' noise sources (due to the environment or intrinsic to the target) begin to play a significant role. These include the dynamical effect of previously unknown stellar compan- ions, astrometric 'jitter' induced by stellar surface activity (spots, flares), and by the presence of a circumstellar protoplanetary disk (the motion of the disk center of mass provoked by the excitation of spiral density waves by an embedded planet, dynamical perturbations due to the disk self-gravity if it's marginally unstable, and time-variable, asymmetric starlight scattering). Such effects have been stud- ied in detail by several authors (Sozzetti 2005, and references therein; Eriksson & Lindegren 2007). The general picture is that, unlike the radial-velocity case, µas-level astrometry is significantly less affected by the above astrophysical noise sources. 3.3 Modeling Planetary Systems There are many difficulties inherent to the problem of astrometric orbit fitting for planetary systems. Orbital fits require highly non-linear fitting procedures, with a large number of model parameters: For a system with np planetary com- panions, the model parameters of the orbital fit will be 5 + 7 × np, i.e. the five classic astrometric parameters + seven orbital elements per object, not includ- ing additional solutions for the space motion of reference stars. Particularly for np > 1 , several complex problems must be overcome in order to successfully fit multiple Keplerian orbits. These include, for example, a) the trade-off between accuracy in the determination of the mutual inclination angles between pairs of planetary orbits, single-measurement precision and redundancy in the number of A. Sozzetti: Planetary Systems with µas Astrometry 9 observations with respect to the number of estimated model parameters (typically nobs >> np), b) the merging of radial velocity + astrometric datasets in combined solutions to improve the quality of orbit reconstruction and mass determination, c) the careful assessment of the relative robustness and reliability of different pro- cedures for multiple-planet orbital fits, and d) the challenge to correctly identify signals with amplitude close to the measurement uncertainties (those typically pro- duced by terrestrial planets), particularly in the presence of larger signals induced by other companions and/or sources of astrophysical noise (due to, e.g., stellar surface structure or variably illuminated disks) of comparable magnitude. All the above issues can have a significant impact in any attempt to gauge the actual capabilities of an astrometric instrument aiming at detecting and characterizing planetary systems, and terrestrial planets in particular. Recent studies (Casertano et al. 2008; Traub et al., this volume; Wright et al., in preparation; Sozzetti et al., in preparation) have begun investigating in detail some of the abovementioned critical aspects. First of all, in these works several independent algorithms for single- and multiple-component orbital fits have been implemented and utilized. These robust, global least-squares fitting procedures adopt, for example, partial linearizations of the multi-body Kepler problem, dif- ferent minimization techniques to optimally search the orbital parameter space, such as Bayesian inference and Markov-Chain Monte Carlo analysis (e.g., Ford & Gregory 2007), or frequency decomposition (e.g., Konacki et al. 2002), and are used to carry out single- and multiple-planet orbital solutions on simulated astro- metric data alone as well as on combined high-precision astrometric+RV datasets. For example, within the context of a double-blind tests campaign for planet de- teciton with SIM-Lite (see § 4.2 and Traub et al. in this volume), the figure of merit utilized in the (iterative) minimization process is the sum of the separate χ2 values: χ2 tot = X(cid:18) xM − x⋆ σx (cid:19)2 σy (cid:19)2 +X(cid:18) yM − y⋆ +X(cid:18) RV M − RV ⋆ σRV (cid:19)2 , (3.1) where x⋆, y⋆ are the position differences (in rectangular coordinates) between a target and a reference object as measured by SIM-Lite and RV ⋆ the radial velocity of the primary, respectively, σx, σy, and σRV are the associated uncertainties, and xM , yM , and RV M are the predicted values based on the orbital elements at each iteration, and the sum is carried out over all observations of each kind. A particularly valuable aspect of this approach is the ability to carry out combined (three-dimensional) solutions even when one type of observation (astrometry or spectroscopy, and sometimes both) has insufficient coverage for an independent fit. Simultaneous orbital fits can particularly strenghten the determination of orbital elements and masses of the companions because they fully exploit the redundancy constraints available from both types of data. Second, in the above studies a number of statistical indicators (periodogram FAPs, F-tests, MLR tests) have been used, and the criteria for regulating relative agreement among them established, for a most thorough, robust assessment of the quality and reliability of the orbital solutions obtained. In addition, detailed 10 Title : will be set by the publisher understanding of the statistical properties of the uncertainties associated with the model parameters have been obtained, through improved understanding of the relative merit of error analyses based on e.g., covariance matrices, χ2 surface mapping, and bootstrapping procedures. This approach is necessary because of the large parameter space to be investigated and due to the fact that for highly non-linear fitting procedures the statistical properties of the solutions are not at all trivial (and significantly differ from those of linear models). Doppler surveys are already facing the challenges of fitting multiple-component orbits, and it is not uncommon to find strong disagreement between solutions (and sometimes number of planets detected!) presented by different teams (e.g., Butler et al. 2006; Gregory 2007). Finally, for multiple-component orbital fits, future work will also focus on the inclusion of N-body integrators, in order to account for possibly significant dy- namical interactions. In multi-planet systems, planet-planet interactions can sig- nificantly alter the RV and/or astrometric signature of the system. In such cases where interactions are important (as for the GJ 876 planetary system), a full dy- namical (Newtonian) fit involving an n-body code must be used to properly model the data and to ensure the short- and long-term stability of the solution. 3.4 Achieving µas Astrometry: Ground-Based Experiments The possibility of achieving µas-level astrometric precision from the ground with monopupil telescopes at visible wavelengths has been tested by numerous past experiments (e.g., Gatewood 1987; Han 1989; Monet et al. 1992), essentially con- firming the theoretical limits imposed by atmospheric noise (e.g., Lindegren 1980) that hamper the ability to significantly push below uncertainties on the order of ∼ 1 mas, particularly in the long term. The best short-term precision achieved with the 5-m Palomar telescope (Pravdo & Shaklan 1996) motivated the Stellar Planet Survey (STEPS), an astrometric survey for very low-mass companions to nearby M-dwarf stars. The long-term noise floor is ∼ 1 mas (Pravdo et al. 2005). Recently theoretical studies which include adaptive optics observations and sym- metrization of the reference frame to remove low-frequency components of the image motion spectrum and improve image centroid at the data-reduction stage predict improved performances for 10-m class telescopes (Lazorenko & Lazorenko 2004). The theoretical predictions have recently begun to be put under experi- mental tests. Lazorenko (2006), Lazorenko et al. (2007), Roll et al. (2008) and Cameron et al. (2009) have employed relatively narrow-field imagers on the Palo- mar and VLT telescopes to demonstrate short-term 100 − 300 µas precision (see also the papers by Roll et al. and Helminiak in this volume). Pioneering stud- ies of coronagraphic astrometry (Digby et al. 2006), which encompass the use of an occulting mask in conjunction with adaptive optics devices, have investigated several methods (centroiding, instrument feedback, analysis of point-spread func- tion symmetry) to carry out precision astrometry of an occulted star. Preliminary results show performances not below mas-level precision. The promise of long-baseline optical/infrared interferometry for high-precision A. Sozzetti: Planetary Systems with µas Astrometry 11 Fig. 1. Dependence of astrometric accuracy of PRIMA with two ATs on stellar brightness (K-band magnitude) for 10′′ separation between primary (PS) and secondary (SeS) star and 30 min integration time. Note that this figure is the result of an error budget simulation, based on PRIMA subsystem specifications, but without knowing the actual throughput and performance of all components. The actual sensitivity of PRIMA will be verified on real stars during the commissioning of the instrument in 2009. Credits: Launhardt et al. 2008. astrometry (Shao & Colavita 1992) has been tested by a number of experiments in the past (e.g., Colavita et al. 1994) The best performances have been achieved by the Palomar Testbed Interferometer in phase-referencing mode, with ∼ 100 µas short-term accuracy for ∼ 30′′ binaries (Lane et al. 2000) and 20 − 50 µas for sub- arcsec binaries (Lane & Muterspaugh 2004) within the context of the Palomar High-precision Astrometric Search for Exoplanet Systems (PHASES) program. The PHASES observations have been able to exclude tertiary companions with masses as small as a few Jupiter masses with a < 2 AU in several binary systems (Muterspaugh et al. 2006). The predicted astrometric performances of large aperture, ground-based inter- ferometers equipped with adaptive optics systems, such as Keck-I (Ragland et al. 2008) and the VLTI (Glindemann et al. 2000), hold promise to approach the actual limits of this technique from the ground. The ASTrometric and phase-Referenced Astronomy (ASTRA) project upgrade (Pott et al. 2008) of Keck-I will add dual- star capability for high sensitivity observations and dual-star astrometry. The new facility, to be ready within 2 years' time, has quoted limiting magnitudes 10 mag and 15 mag in K-band for narrow-angle astrometry at the ∼ 100 µas level between 12 Title : will be set by the publisher pairs of objects separated by < 20 − 30 arcsec. PRIMA, the instrument for Phase- Referenced Imaging and Micro-arcsecond Astrometry at the VLTI (Delplancke 2008), is currently being integrated and tested at Paranal, with science operations presently scheduled for early 2010. Similarly to Keck-I, the PRIMA/VLTI facility is designed to perform narrow-angle interferometric astrometry with two ATs of a target and one reference star separated by up to 1 arcmin. Launhardt et al. (2008) have shown that, using error budget simulations based on PRIMA subsys- tem specifications, in order to achieve the astrometric precision of 10 µas in 30 min integration time, the primary target and the reference (of limiting K-band brightness 8 mag and 14 mag, respectively) should not be separated by more than ∼ 20′′ (see Figure 1). The Exoplanet Search with PRIMA (ESPRI) Consortium (Launhardt et al. 2008) will carry out a three-fold observing program focused on the astrometric characterization of known radial velocity planets within ≤ 200 pc from the Sun, the astrometric detection of low-mass planets around nearby stars of any spectral type within ≈ 15 pc from the Sun, and the search for massive planets orbiting young stars with ages in the range 5 − 300 Myr within ∼ 100 pc from the Sun. The initial target list includes ∼ 900 stars, but the number of good references will likely reduce the batch by almost an order of magnitude, resulting in a tentative target list of ∼ 100 stars to be observed with PRIMA during the 5-yr duration of the survey (Launhardt et al. 2009, in preparation). An extensive program of preparatory observations to characterize suitable references and weed out from the target list excessively active stars and short-period binaries is underway, which includes high dynamic range near-infrared photometric imaging and spectroscopic observations. 3.5 Achieving µas Astrometry: Space-Borne Observatories Relative, narrow-angle astrometry from space has been performed so far with the Fine Guidance Sensors aboard HST, while global astrometric measurements have been carried out for the first time by Hipparcos. For HST/FGS astrometry with respect to a set of reference objects near the target (within the 5 × 5 arcsec instantaneous field of view of FGS), pre-launch error budget estimates (e.g., Bahcall & O'Dell 1980) predicted 1-2 mas single- measurement precision down to mv ≈ 16. This performance level has been demonstrated by e.g., Benedict et al. (1994, 1999) who, in the data reduction of the two-dimensional interferometric measurements, devised ad hoc calibration and data reduction procedures to remove a variety of random and systematic error sources from the astrometric reference frame (intra-observation spacecraft jitter, temperature variations and temperature-induced changes in the secondary mirror position, constant and time-dependent optical field angle distortions, intra-orbit drifts, lateral color corrections). The limiting factor is the spacecraft jitter. A single-measurement precision below 0.5-1 mas is out of reach for HST/FGS. For the Hipparcos all-sky survey, the achievable precision (Lindegren 1989) in the along-scan direction had received pre-launch estimates of ∼ 1 mas on bright A. Sozzetti: Planetary Systems with µas Astrometry 13 objects, with degradation of a factor of several at the magnitude limit of the sur- vey, depending on location on the sky (given the pre-determined scanning law adopted for the satellite). Actual sky-averaged uncertainties confirmed the pre- dictions, with typical uncertainties degrading from ∼ 1.0 mas for mv < 7 to ∼ 4.5 mas for mv ≥ 11 (e.g., Kovalevsky 2002), achieved with a calibration and iterative reduction scheme in several steps (Lindegren & Kovalevsky 1989) that allowed to successfully derive values of positions, proper motions, and parallaxes simulta- neously for ∼ 120, 000 stars by bridging one-dimensional angular measurements along the satellite's instantaneous scanning direction into a global astrometric so- lution over the whole celestial sphere. Without the presence of the atmosphere, and similarly to HST/FGS, the best-achievable single-measurement precision is limited by the uncertainties in the determination of the along-scan attitude. The ability to suppress systematics by at least two orders of magnitude for a space-borne instrument is a major technological goal. Both SIM-Lite and Gaia promise to achieve this level of astrometric precision. The newly redesigned SIM- Lite has a shorter baseline (6 m) with respect to the old SIM configuration (10 m), and it replaces a guide interferometer with a telescope star tracker. It will deliver better than 1 µas narrow-angle astrometry in 1.5 hr integration time (Goul- lioud et al. 2008) on bright targets (mv ≤ 7) and moderately fainter references (mv ≃ 9 − 10). For this purpose, an accuracy on the position of the delay lines of a few tens of pm with a 6-m baseline must be achieved (Zhai et al. 2008). Further- more, a positional stability of internal optical pathlengths of ∼ 10 nm is required, in order to ensure maintenance of the fringe visibility (Goullioud et al. 2008). For Gaia, the success in meeting the goal of ≈ 10 µas single-measurement astrometric precision to hunt for planets around bright stars (mv < 13) will depend on a) the ability to attain CCD centroiding errors not greater than 1/1000 of a pixel in the along-scan direction (Gai et al. 2001) and b) the capability to limit in- strumental uncertainties (thermo-mechanical stability of telescope and focal plane assembly, metrology errors in the monitoring of the basic angle) and calibration errors (chromaticity, charge transfer inefficiency, satellite attitude, focal plane-to- field coordinates transformation) down to the few µas level (e.g., Perryman et al. 2001). 4 The Potential of µas Astrometry A number of authors have tackled the problem of evaluating the sensitivity of the astrometric technique required to detect extrasolar planets and reliably measure their orbital elements and masses (Sozzetti 2005, and references therein). Those works mostly relied on simplying assumptions with regard to a) the error models to be applied to the data (e.g., simple gaussian distributions, perfect knowledge of the instruments) and b) the analysis procedures to be adopted for orbit re- construction (mostly ignoring the problem of identifying adequate configurations of starting values from scratch). The two most recent exercises on this subject (Casertano et al. 2008; Traub et al., this volume) have revisited earlier findings using a more realistic double-blind protocol. In this particular case, several teams 14 Title : will be set by the publisher Fig. 2. Left: Number of giant planets that could be detected and measured by Gaia, as a function of increasing distance. Starcounts are obtained using the Besancon model of stellar population synthesis (Bienaym´e et al. 1987), while the Tabachnik & Tremaine (2002) model for estimating planet frequency as a function of mass and orbital period is utilized. Right: Number of multiple-planet systems that Gaia could potentially detect, measure, and for which coplanarity tests could be carried out successfully. Credits: Casertano et al. 2008. of "solvers" were handled simulated datasets of stars with and without planets and independently defined detection tests, with levels of statistical significance of their own choice, and orbital fitting algorithms, using any local, global, or hybrid solution method that they devised was best. The solvers were provided no a priori information on the actual presence of planets around a given target. 4.1 Gaia DBT In the large-scale, double-blind test (DBT) campaign carried out to estimate the potential of Gaia for detecting and measuring planetary systems, Casertano et al. (2008) showed that a) planets with α ≃ 6σ (where σ is the single-measurement error) and orbital periods shorter than the nominal 5-yr mission lifetime could be accurately modeled, and b) for favorable configurations of two-planet systems with well separated periods (both planets with P ≤ 4 yr and α/σ ≥ 10, redundancy over a factor of 2 in the number of observations) it would be possible to carry out meaningful coplanarity tests, with typical uncertainties on the mutual inclination angle of ≤ 10 deg. Both subtle differences as well as significant discrepancies were found in the orbital solutions carried out by different solvers. This constitutes further evidence that the convergence of non-linear fitting procedures and the quality of orbital solutions (particularly for multiple systems and for systems with small astrometric signals) can both be significantly affected by the choice of the starting guesses for the parameters in the orbital fits, by the adoption of different statistical indicators of the quality of a solution, and varied levels of significance of the latter. Overall, the authors concluded that Gaia could discover and measure massive giant planets (Mp ≥ 2 − 3 MJ) with 1 < a < 4 AU orbiting solar-type stars as far as the nearest star-forming regions, as well as explore the domain of Saturn- mass planets with similar orbital semi-major axes around late-type stars within 30-40 pc. These results can be turned into a number of planets of given mass A. Sozzetti: Planetary Systems with µas Astrometry 15 Fig. 3. Left: reliability of detection (number of detected planets divided by the number of detected planets + the number of false alarms) vs. planet type, for various teams contributing to the SIM-Lite double-blind tests campaign. Right: completeness (number of detected planets divided by number of detectable planets) vs. planet type, for the same teams. Credits: Wes Traub. and orbital separation that can be detected and measured by Gaia, using Galaxy models and the current knowledge of exoplanet frequencies. By inspection of the tables in Figure 2, one then finds that Gaia's main strength will be its ability to measure accurately orbits and masses for thousands of giant planets, and to per- form coplanarity measurements for a few hundred multiple systems with favorable configurations. 4.2 SIM-Lite DBT The second study carried out in double-blind mode set out to answer a number of questions relating to the detectability of Earth-like planets (terrestrial masses and Habitable-Zone orbits) in multi-planet systems, using a combination of SIM-Lite astrometry and ground-based radial-velocity observations. As discussed in detail by Traub et al. (this volume), the theoretical predictions of required signal-to-noise ratio to detect and measure planetary systems were proven correct (similarly to the Casertano et al. (2008) analysis). Both reliability of detection (i.e., the probability that a detection based on a periodogram analysis is true and not a false alarm) and completeness (i.e., the probability that a planet will be detected) were gauged, and found in very good agreement with the expectations. Similarly to Casertano et al. (2008), this study also highlighted how different detection and orbit fitting algorithms, given the same datasets, can perform in measurably different ways (see Figure 3). Overall, it was demonstrated that other planets do not significantly interfere with the detection of terrestrial planets in the Habitable Zone, and to reach the sensitivity needed to detect Earth-like planets ∼ 40% of a 5-yr SIM-Lite mission with σ = 0.82 µas is required, with the additional help of 15-yr baseline RV data. A second phase of the SIM-Lite double-blind excercise is underway, in which the study will be extended to real stars in the SIM-Lite target list, all detections 16 Title : will be set by the publisher will be subjected to additional statistical tests and long-term stability analyses in order to more robustly assess the reliability of any detection and the confidence on multi-component orbital solutions, and requirements will be placed on the accuracy on masses and orbital parameters necessary to usefully inform direct imaging surveys about the epoch and location of maximum brightness, in order to estimate optimal visibility. The latter issue is of particular relevance in light of concerns recently raised by Brown (2009) about the ultimate requirements on astrometry to support the planning of direct observations of terrestrial habitable planets, indicating µas-level precision may not suffice. 5 Summary An improvement of 2-3 orders of magnitude in achievable measurement precision, down to the µas level, would allow this technique to achieve in perspective the same successes of the Doppler method, for which the improvement from the km s−1 to the m s−1 precision opened the doors for ground-breaking results in ex- oplanetary science. Indeed, µas astrometry is almost coming of age. Provided the demanding technological and calibration requirements to achieve the required level of measurement precision are met, future observatories at visible and near- infrared wavelengths, using both monolithic as well as diluted architectures from the ground (VLTI/PRIMA, Keck-I) and in space (Gaia, SIM-Lite) hold promise for crucial contributions to many aspects of planetary systems astrophysics (formation theories, dynamical evolution, internal structure, detection of Earth-like planets), in combination with data collected with other indirect and direct techniques. Figure 4 shows the Mp-a diagram with the plotted present-day and achievable sensitivities of transit photometry and radial-velocity, and with the expected SIM- Lite and Gaia detection thresholds at 10 pc, 25 pc, and 200 pc, respectively. The presently known planets detected by the various methods are also shown, along with the predicted distribution of recent models (Ida & Lin 2008). At first glance, one could get the impression that the impact of astrometric measurements (except for those obtained by SIM-Lite around the nearest stars) may not bear great potential. However, the relative importance of different planet detection techniques should not be gauged by looking at their discovery potential per se, but rather in connection to outstanding questions to be addressed and answered in the science of planetary systems. Some of the most important issues for which µas astrometry will play a key role in the next decade include: a) a significant refinement of our understand- ing of the statistical properties of extrasolar planets: for example, the predicted Gaia database of several thousand extrasolar planets with well-measured proper- ties will allow to test the fine structure of giant planet parameters distributions and frequencies, and to investigate their possible changes as a function of stellar mass, metal content, and age with unprecedented resolution; b) crucial tests of theoretical models of gas giant planet formation and migration: for example, spe- cific predictions on formation time-scales and the role of varying metal content in the protoplanetary disk will be probed with unprecedented statistics thanks to the A. Sozzetti: Planetary Systems with µas Astrometry 17 Fig. 4. Exoplanets discovery space for the astrometric, Doppler, and transit techniques. Detectability curves are defined on the basis of a 3-σ criterion for signal detection. The upper and middle blue curves are for Gaia astrometry with σA = 10 µas, assuming a 1-M⊙ G dwarf primary at 200 pc and a 0.4-M⊙ M dwarf at 25 pc, respectively, while the lower blue curve is for SIM-Lite astrometry of a 1-M⊙ star at 10 pc with σA = 0.8 µas. For both Gaia and SIM-Lite, survey duration is set to 5 yr. The radial velocity curves (red lines) assume σRV = 3 m s−1 (upper curve) and σRV = 1 m s−1 (lower curve), M⋆ = 1M⊙, and 10-yr survey duration. For visible-light transit photometry (green curves), the assumption are σV = 5 × 10−3 mag (upper curve) and σV = 1 × 10−5 mag (lower curve), S/N = 9, M⋆ = 1 M⊙, R⋆ = 1 R⊙, uniform and dense (>> 1000 datapoints) sampling. Pink dots indicate the inventory of Doppler-detected exoplanets as of December 2008. Transiting systems are shown as light-blue filled diamonds, while the red hexagons are planets detected by microlensing. Solar System planets are also shown as green pentagons. The yellow small dots represent a theoretical distribution of masses and final orbital semi-major axes from Ida & Lin (2008). thousands of metal-poor stars and hundreds of young stars screened for giant plan- ets out to a few AUs by Gaia, VLTI/PRIMA, and SIM-Lite; c) key improvements in our comprehension of importants aspects of the astrophysics of multiple-planet systems: for example, coplanarity tests for hundreds of multiple-planet systems will be carried out with Gaia, SIM-Lite, and VLTI/PRIMA, and this, in combi- 18 Title : will be set by the publisher nation with data available from Doppler measurements and transit timing, could allow to discriminate between various proposed mechanisms for eccentricity exci- tation, thus significantly improving our comprehension of the role of dynamical interactions in the early as well as long-term evolution of planetary systems; d) an important contribution to the understanding of direct detections of giant extraso- lar planets: for example, accurate knowledge of all orbital parameters and actual mass are essential for understanding the thermophysical conditions on a planet and for determining its visibility. Actual mass estimates and full orbital geome- try determination for suitable systems (with typical separations > 0.1′′), obtained by means of high-precision astrometric measurements (with Gaia, SIM-Lite, and VLTI/PRIMA), will inform direct imaging surveys about the epoch and location of maximum brightness, in order to estimate optimal visibility, and will help in the modeling and interpretation of the phase functions and light-curves of giant planets (the first prediction about where and when to look for the planet ǫ Eridani b was recently made by Benedict et al. (2006) using HST/FGS astrometry); e) the collection of essential supplementary data for the optimization of the target lists of future observatories (e.g., Beichman et al. 2007) aiming at the direct detection and spectroscopic characterization of terrestrial, habitable planets in the vicinity of the Sun. For example, astrometry of all nearby stars within 25 pc of the Sun with 1-10 µas precision (with SIM-Lite and Gaia in space, and VLTI/PRIMA from the ground) will provide a comprehensive database of F-G-K-M stars screened for Jupiter-, Saturn-, and Neptune-mass companions out to several AUs. These ob- servations would help probing the long-term dynamical stability of their habitable zones, where terrestrial planets may have formed, and maybe found, complement- ing on-going efforts of Doppler surveys and studies of exo-zodiacal cloud emission (with ground-based facilities such as Keck-I, VLTI, and LBTI). 6 Acknowledgements I am especially grateful to the Conference organizers for giving me the opportu- nity to write this review. I am indebted to Mario Lattanzi, Stefano Casertano, and Alessandro Spagna for guiding my first steps in the realm of high-precision astrometry. I warmly thank Fritz Benedict, Wes Traub, Jason Wright, and Ralf Launhardt for very useful discussions and for providing material ahead of publi- cation. Financial support from the Italian Space Agency through ASI contract I/037/08/0 (Gaia Mission - The Italian Participation to DPAC) is gratefully ac- knowledged. References Alfv´en, H. 1943, Nat., 152, 721 Baglin, A., et al. 2002, in Radial and Nonradial Pulsations as Probes of Stellar Physics, eds. C. Aerts, T. R. Bedding, & J. Christensen-Dalsgaard, ASP Conf. Ser., 259, 626 Bahcall, J. N., & O'dell, C. R. 1980, in Scientific Research with the Space Telescope, Proc. IAU Colloq. 54, NASA CP-2111, 5 A. Sozzetti: Planetary Systems with µas Astrometry 19 Bean, J.L., et al. 2007, AJ, 134, 749 Bean, J.L., & Seifahrt, A. 2009, A&A, in press (arXiv:0901.3144) Bean, J. L., et al. 2007, AJ, 134, 749 Beichman, C.A., Fridlund, M., Traub, W.A., Stapelfeldt, K.R., Quirrenbach, A., & Sea- ger, S. 2007, in Protostars and Planets V, B. Reipurth, D. Jewitt, and K. Keil (eds.), University of Arizona Press, Tucson, 915 Bender, C., Simon, M., Prato, L., Mazeh, T., & Zucker, S. 2005, AJ, 129, 402 Benedict, G. F., et al. 1994, PASP, 106, 327 Benedict, G. F., et al. 1999, AJ, 118, 1086 Benedict, G. F., et al. 2002, ApJl, 581, L115 Benedict, G. F., et al. 2006, AJ, 132, 2206 Benedict, G. F., McArthur, B.E., & Bean, J.L. 2008, Proc. IAU Symp. 248, 23 Beuzit, J.-L., Mouillet, D., Oppenheimer, B.R., & Monnier, J.D. 2007, in Protostars and Planets V, B. Reipurth, D. Jewitt, and K. Keil (eds.), University of Arizona Press, Tucson, 717 Bienaym´e, O., Robin, A. C., & Cr´ez´e, M. 1987, A&A, 180, 94 Borucki, W. J., et al. 2003, in Future EUV/UV and Visible Space Astrophysics Missions and Instrumentation, eds. J. C. Blades & O. H. W. Siegmund, Proc. SPIE, 4854, 129 Brown, R.A. 2009, ApJ submitted (arXiv:0901.4897) Brumberg, V. A. 1991, Essential Relativistic Celestial Mechanics (Bristol: Hilger) Butler, R.P., et al. 2006, ApJ, 646, 505 Cameron, P.B., Britton, M.C., & Kulkarni, S.R. 2009, AJ, 137, 83 Charbonneau, D., Brown, T.M., Burrows, A., & Laughlin, G. 2007, in Protostars and Planets V, B. Reipurth, D. Jewitt, and K. Keil (eds.), University of Arizona Press, Tucson, 701 Correia, A.C.M., et al. 2009, A&A, in press (arXiv:0902.0597) Crosta, M.T., Vecchiato, A., de Felice, F., Lattanzi, M.G., & Bucciarelli, B. 2003, Celes- tial Mechanics and Dynamical Astronomy, 87, 209 Croswell, K. 1988, Astronomy, 16, 6 de Felice, F., Lattanzi, M. G., Vecchiato, A., & Bernacca, P. L. 1998, A&A, 332, 1133 de Felice, F., Bucciarelli, B., Lattanzi, M. G., & Vecchiato, A. 2001, A&A, 373, 336 de Felice, F., Crosta, M. T., Vecchiato, A., Lattanzi, M. G., & Bucciarelli, B. 2004, ApJ, 607, 580 de Felice, F., Vecchiato, A., Crosta, M.T., Bucciarelli, B., & Lattanzi, M.G. 2006, ApJ, 653, 1552 de Felice, F., & Preti, G. 2006, Classical and Quantum Gravity, 23, 5467 de Felice, F., & Preti, G. 2008, Classical and Quantum Gravity, 25, 165015 Delplancke, F. 2008, New Astronomy Reviews, 52, 199 Digby, A.P., et al. 2006, ApJ, 650, 484 Durisen, R. H., Boss, A. P., Mayer, L., Nelson, A. F., Quinn, T., & Rice, W. K. M. 2007, in Protostars and Planets V, B. Reipurth, D. Jewitt, and K. Keil (eds.), University of Arizona Press, Tucson, 607 Eriksson, U., & Lindegren, L. 2007, A&A, 476, 1389 Fischer, D.A., et al. 2008, ApJ, 675, 790 20 Title : will be set by the publisher Ford, E.B., & Gregory, P.C. 2007, ASP Conf. Ser., 379, 189 Ford, E.B., & Rasio, F.A. 2008, ApJ, 686, 621 Gai, M., Carollo, D., Delb`o, M., Lattanzi, M. G., Massone, G., Bertinetto, F., Mana, G., & Cesare, S. 2001, A&A, 367, 362 Gai, M., & Cancelliere, R. 2008, MNRAS, 391, 1451 Gatewood, G. D., & Eichhorn, H. 1973, AJ, 78, 769 Gatewood, G. D. 1974, AJ, 79, 52 Gatewood, G. D. 1987, AJ, 94, 213 Gatewood, G. D., Stein, J., de Jonge, J. K., Persinger, T., Reiland, T., & Stephenson, B. 1992, AJ, 104, 1237 Gatewood, G. D. 1996, BAAS, 28, 885 Gatewood, G. D., Han, I., & Black, D. C. 2001, ApJl, 548, L61 Gaudi, B.S. 2007, ASP Conf. Ser., 398, 479 Glindemann, A., et al. 2000, Proc. SPIE, 4006, 2 Goullioud, R., Catanzarite, J.H., Dekens, F.G., Shao, M., & Marr, J.C., IV 2008, Proc. SPIE, 7013, 151 Go´zdziewski, K., Migaszewski, C., & Musieli´nski, A. 2008, Proc. IAU Symp. 249, 447 Gregory, P.C. 2007, MNRAS, 381, 1607 Han, I. 1989, AJ, 97, 607 Han, I., Black, D. C., & Gatewood, G. D. 2001, ApJl, 548, L57 Heintz, W. D. 1976, MNRAS, 175, 533 Heintz, W.D., 1978, ApJ, 220, 931 Hershey, J. L. 1973, 78, 421 Hershey, J. L., & Lippincott, S. L. 1982, AJ, 87, 840 Hinse, T.C., Michelsen, R., Jørgensen, U.G., Go´zdziewski, K., & Mikkola, S. 2008, A&A, 488, 1133 Ida, S., & Lin, D.N.C. 2008, ApJ, 685, 584 Jeans, J.H. 1943, Nat., 152, 721 Jones, B. W., Underwood, D. R., & Sleep, P. N. 2005, ApJ, 622, 1091 Kalas, P., et al. 2008, Science, 322, 1345 Klioner, S. A., & Kopeikin, S. M. 1992, AJ, 104, 897 Klioner, S. A. 2003, AJ, 125, 1580 Klioner, S. A. 2004, Phys. Rev. D, 69, 124001 Konacki, M., Maciejewski, A.J., & Wolszczan, A. 2002, ApJ, 567, 566 Kovalevsky, J. 2002, Modern Astrometry, 2nd ed. (Berlin: Springer-Verlag) Lane, B. F., Colavita, M. M., Boden, A. F., & Lawson, P. R. 2000, Proc. SPIE, 4006, 452 Lane, B. F., & Muterspaugh, M. W. 2004, ApJ, 601, 1129 Launhardt, R., et al. 2008, Proc. IAU Symp. 248, 417 Lazorenko, P.F. 2002, A&A, 396, 353 Lazorenko, P.F., & Lazorenko, G.A. 2004, A&A, 427, 1127 Lazorenko, P.F. 2006, A&A, 449, 1271 A. Sozzetti: Planetary Systems with µas Astrometry 21 Lazorenko, P.F., Mayor, M., Dominik, M., Pepe, F., Segransan, D., & Udry, S. 2007, A&A, 471, 1057 Lee, J.W., et al. 2009, AJ, 137, 3181 Levison, H.F., Morbidelli, A., Gomes, R., & Backman, D. 2007, in Protostars and Planets V, B. Reipurth, D. Jewitt, and K. Keil (eds.), University of Arizona Press, Tucson, 669 Lindegren, L. 1980, A&A, 89, 41 Lindegren, L. 1989, in The Hipparcos Mission: Pre-Launch Status, ESA SP-1111, Vol. III, 311 Lindegren, L., & Kovalevsky, J. 1989, in The Hipparcos Mission: Pre-Launch Status, ESA SP-1111, Vol. III, 1 Lippincott, S. L. 1960, AJ, 65, 349 Lippincott, S. L. 1960, AJ, 65, 445 Lissauer, J. J., & Stevenson, D. J. 2007, in Protostars and Planets V, B. Reipurth, D. Jewitt, and K. Keil (eds.), University of Arizona Press, Tucson, 591 Marois, C., Macintosh, B., Barman, T., Zuckerman, B., Song, I., Patience, J., Lafreniere, D., & Doyon, R. 2008, Science, in press (arXiv:0811.2606) Mayor, M., et al. 2008, A&A, 493, 639 Mazeh, T., Zucker, S., dalla Torre, A., & van Leeuwen, F. 1999, ApJl, 522, L149 McArthur, B. E., et al. 2004, ApJl, 614, L81 McGrath, M. A., et al. 2002, ApJl, 564, L27 Menou, C., & Tabachnik, S. 2003, ApJ, 583, 473 Monet, D. G., Dahn, C. C., Vrba, F. J., Harris, H. C., Pier, J. R., Luginbuhl, C. B., & Ables, H. D. 1992, AJ, 103, 638 Muterspaugh, M.W., et al. 2006, ApJ, 653, 1469 Nagasawa, M., Thommes, E.W., Kenyon, S.J., Bromley, B.C., & Lin, D.N.C. 2007, in Protostars and Planets V, B. Reipurth, D. Jewitt, and K. Keil (eds.), University of Arizona Press, Tucson, 639 Papaloizou, J.C.B., Nelson, R.P., Kley, W., Masset, F.S., & Artymowicz, P. 2007, in Protostars and Planets V, B. Reipurth, D. Jewitt, and K. Keil (eds.), University of Arizona Press, Tucson, 655 Pepe, F.A., & Lovis, C. 2008, Physica Scripta, 130, 014007 Perryman, M. A. C., et al. 1996, A&A, 310, L21 Perryman, M. A. C., et al. 2001, A&A, 369, 339 Pott, J.-U., et al. 2008, Proc. summerschool "Astrometry and Imaging with the Very Large Telescope Interferometer" (arXiv:0811.2264) Pourbaix, D. 2001, A&A, 369, L22 Pourbaix, D., & Arenou, F. 2001, A&A, 372, 935 Pravdo, S. H., & Shaklan, S. B. 1996, ApJ, 465, 264 Pravdo, S.H., Shaklan, S.B., & Lloyd, J. 2005, ApJ, 630, 528 Ragland, S., et al. 2008, Proc. SPIE, 7013, 10 Reffert, S., & Quirrenbach, A. 2006, A&A, 449, 699 Reuyl, D., & Holmberg, E. 1943, ApJ, 97, 41 Roll, T., Seifahrt, A., & Neuhauser, R. 2008, Proc. IAU Symp. 248, 48 22 Title : will be set by the publisher Shao, M., & Colavita, M. M. 1992, A&A, 262, 353 Silvotti, R., et al. 2007, Nat., 449, 189 Sozzetti, A. 2005, PASP, 117, 1021 Sozzetti, A., & Desidera, S. 2009, A&A, submitted Strand, K.A. 1943, PASP, 55, 29 Struve, O. 1952, The Observatory, 72, 199 Tabachnik, S., & Tremaine, S. 2002, MNRAS, 335, 151 van de Kamp, P. 1963, AJ, 68, 515 van de Kamp, P. 1969, AJ, 74, 238 van de Kamp, P. 1969, AJ, 74, 757 van de Kamp, P. 1975, AJ, 80, 658 van de Kamp, P. 1982, Vistas in Astronomy, 26, 141 Vecchiato, A., Lattanzi, M.G., Bucciarelli, B., Crosta, M., de Felice, F., & Gai, M. 2003, A&A, 399, 337 Will, C. M. 1993, Theory and Experiment in Gravitational Physics (rev. ed.; Cambridge: Cambridge Univ. Press) Wolszczan, A., & Frail, D. A. 1992, Nat., 355, 145 Wright, J.T., et al. 2007, ApJ, 657, 533 Wright, J.T., Upadhyay, S., Marcy, G.W., Fischer, D.A., Ford, E.B., & Asher, J.J. 2009, ApJ, in press (arXiv:0812.1582) Zhai, C., et al. 2008, Proc. SPIE, 7013, 157 Zucker, S., & Mazeh, T. 2000, ApJl, 531, L67 Zucker, S., & Mazeh, T. 2001, ApJ, 562, 549
1602.01368
1
1602
2016-02-03T17:11:09
Directly observing continuum emission from self-gravitating spiral waves
[ "astro-ph.EP" ]
We use a simple, self-consistent, self-gravitating semi-analytic disc model to conduct an examination of the parameter space in which self-gravitating discs may exist. We then use Monte-Carlo radiative transfer to generate synthetic ALMA images of these self-gravitating discs to determine the subset of this parameter space in which they generate non-axisymmetric structure that is potentially detectable by ALMA. Recently, several transition discs have been observed to have non-axisymmetric structure that extends out to large radii. It has been suggested that one possible origin of these asymmetries could be spiral density waves induced by disc self-gravity. We use our simple model to see if these discs exist in the region of parameter space where self-gravity could feasibly explain these spiral features. We find that for self-gravity to play a role in these systems typically requires a disc mass around an order of magnitude higher than the observed disc masses for the systems. The spiral amplitudes produced by self-gravity in the local approximation are relatively weak when compared to amplitudes produced by tidal interactions, or spirals launched at Lindblad resonances due to embedded planets in the disc. As such, we ultimately caution against diagnosing spiral features as being due to self-gravity, unless the disc exists in the very narrow region of parameter space where the spiral wave amplitudes are large enough to produce detectable features, but not so large as to cause the disc to fragment.
astro-ph.EP
astro-ph
Mon. Not. R. Astron. Soc. 000, 1 -- 13 (2015) Printed 17 September 2018 (MN LATEX style file v2.2) Directly observing continuum emission from self-gravitating spiral waves Cassandra Hall1(cid:63), Duncan Forgan2, Ken Rice1,Tim J. Harries3, Pamela D. Klaassen4 and Beth Biller1 1SUPA†, Institute for Astronomy, University of Edinburgh, Blackford Hill, Edinburgh, EH9 3HJ, UK 2SUPA†, School of Physics & Astronomy, University of St Andrews, North Haugh, St Andrews, KY16 9SS, UK 3Department of Physics and Astronomy, University of Exeter, Stocker Road, Exeter EX4 4QL, UK 4UK Astronomy Technology Center, Royal Observatory Edinburgh, Blackford Hill, Edinburgh, EH9 3HJ, UK January 2016 ABSTRACT We use a simple, self-consistent, self-gravitating semi-analytic disc model to conduct an examination of the parameter space in which self-gravitating discs may exist. We then use Monte-Carlo radiative transfer to generate synthetic ALMA images of these self-gravitating discs to determine the subset of this parameter space in which they generate non-axisymmetric structure that is potentially detectable by ALMA. Recently, several transition discs have been observed to have non-axisymmetric structure that extends out to large radii. It has been sug- gested that one possible origin of these asymmetries could be spiral density waves induced by disc self-gravity. We use our simple model to see if these discs exist in the region of pa- rameter space where self-gravity could feasibly explain these spiral features. We find that for self-gravity to play a role in these systems typically requires a disc mass around an order of magnitude higher than the observed disc masses for the systems. The spiral amplitudes produced by self-gravity in the local approximation are relatively weak when compared to amplitudes produced by tidal interactions, or spirals launched at Lindblad resonances due to embedded planets in the disc. As such, we ultimately caution against diagnosing spiral fea- tures as being due to self-gravity, unless the disc exists in the very narrow region of parameter space where the spiral wave amplitudes are large enough to produce detectable features, but not so large as to cause the disc to fragment. Key words: planets and satellites:formation -- Solar system: formation -- stars:pre-main- sequence -- planetary systems: formation -- planetary formation: protoplanetary discs. Monte Carlo radiative transfer synthetic images MWC758 SAO206462 HD142527 1 INTRODUCTION It is widely accepted that low-mass stars form through the collapse of cold, dense molecular cloud cores (Terebey, Shu & Cassen 1984; McKee & Ostriker 2007). These cores will typically contain some amount of angular momentum, meaning that all the mass cannot fall directly onto the central protostar; some must first pass through a protostellar accretion disc. In these discs, molecular viscosity alone does not exert large enough torques to redistribute angular momentum out to large radii, allowing mass to accrete onto the central protostar. However, instabilities that develop into the turbu- lent regime can produce considerable torques that can then drive mass transport. If these discs are sufficiently massive then self-gravity could (cid:63) Email: [email protected] † Scottish Universities Physics Alliance c(cid:13) 2015 RAS be significant, and the gravitational instability (GI) could be the main angular momentum transport mechanism (Toomre 1964; Laughlin & Bodenheimer 1994) during these early times. If GI is significant in these discs, then we would expect there to be non- axisymmetric structures, typically spiral density waves. Discs around very young stars are, however, heavily embed- ded in their cloud cores, making them difficult to observe at optical wavelengths (Dunham et al. 2014). High resolution interferometric observations, in radio or sub-mm, are therefore required to resolve the disc. Currently, however, observations of this wavelength with a high enough resolution to resolve spiral arms are rare. Here we examine the parameter space of self-gravitating pro- tostellar discs to determine the range of accretion rates, disc masses and outer radii in which extended spiral features could be detected by ALMA. Previous studies which describe simulated ALMA ob- servations of protostellar discs have tended to focus on reproducing the specific morphology of discs, using numerical methods such as 2 Cassandra Hall, Duncan Forgan, Ken Rice, Tim J. Harries, Pamela D. Klaassen and Beth Biller Smoothed Particle Hydrodynamics (SPH), rather than an examina- tion of the parameter space in which they are detectable (Cossins, Lodato & Testi 2010; Ruge et al. 2013; Douglas et al. 2013). Dipierro et al. (2014) and Dipierro et al. (2015) have shown, using simulated observations from SPH simulations, that non- axisymmetric structure in self-gravitating protostellar discs is de- tectable at a wide range of wavelengths using ALMA. Our ap- proach differs to that used in Dipierro et al. (2014) and Dipierro et al. (2015), in that rather than deriving the physical disc struc- ture from numerical simulations, where an artificial cooling law has been imposed, we use a self-consistent, analytic geometry coupled with 3D Monte Carlo radiative transfer (MCRT) to generate emis- sion maps at typical ALMA frequencies. This geometry, unlike in SPH simulations by Dipierro et al. (2014), Dipierro et al. (2015) and Lodato & Rice (2004), is intended to be more realistic. We wish to note, however, that we do not include in our MCRT simula- tions the effect of dust trapping in the spiral arms noted by Dipierro et al. (2015). We suspect that including this would increase the de- tectability of the spiral structure, and we leave further investigation of this to future work. Once we have our MCRT images, these are then used as in- put sky models to the ALMA simulator from the CASA software package (ver 4.3.0) to generate synthetic ALMA images. We stress that it is not our aim to match specific morphology, but rather to ex- amine the conditions under which the over-density in spiral arms is sufficient so as to be detected. We then use this to investigate the re- gion of parameter space (characterised by disc mass accretion rate and outer radius) in which GI-driven spiral density waves may be detectable by ALMA. Currently, there are few, if any, observations which are strictly comparable with what we consider here. However, several transi- tion discs have recently been observed to have non-axisymmetric structure that extends out to large radii. Bearing in mind the diffi- culty of finding strictly comparable samples, we take three of these transition discs as test cases, and apply our simple geometry to them. We aim to determine if these discs exist in the parameter space where self-gravity could be a feasible explanation for their spiral structure. We do not, however, generate synthetic images of these cases, since these test cases are imaged in NIR and scattered light, which is quite different from continuum mm emission. The paper is laid out as follows: The model is described in Section 2, with the 2D and 3D structure given in Sections 2.2 and 2.3 respectively. Section 2.4 describes our Monte Carlo radiative transfer method, and Section 2.5 describes the generation of syn- thetic observations. Our results are given in Section 3. We give the surface density profiles achieved from our model in Section 3.1, and compare our models to scenarios where a fixed cooling law is imposed in Section 3.2. We discuss our general results from our synthetic imaging in Section 3.3, and apply our analytical geometry to test cases in Sections 3.4.1, 3.4.2 and 3.4.3. We draw conclusions in Section 4. 2 MODEL 2.1 Outline Here we use an existing 1D model to examine the parameter space of self-gravitating discs initially developed by Clarke (2009) (see also Rice & Armitage 2009; Forgan & Rice 2013b). We develop it to include 2D and 3D structure, fitting spirals of the shape typ- ically found in Smoothed Particle Hydrodynamics (SPH) simula- tions (given in Figure 1). Figure 1. This is a simulation image of a self-gravitating disc that has reached a state of quasi-equilibrium, with parameterised cooling such that β = 9. Two spirals have been plotted. The blue line is a logarithmic spi- ral, the white line is an Archimedal spiral. The logarithmic spiral appears to fit the spiral arms of the disc the best, although it is fairly arbitrary since Equation (17) has two free parameters. 2.2 Radial Geometry We expect a self-gravitating protostellar system to settle into a state of quasi-steady thermal equilibrium (Paczynski 1978; Gam- mie 2001; Rice & Armitage 2009), with a constant accretion rate M given by M = sΣ 3παc2 Ω , (1) where cs is the sound speed, Σ is the surface density, Ω is the an- gular frequency (since we are considering Keplerian rotation, the epicyclic frequency is simply Ω) and α is the dimensionless shear stress, composed of both hydrodynamic (i.e. Reynolds stress) and gravitational parts. Remarkably, this can be expressed both simply and analytically as (Gammie 2001) αgrav = 4 9 1 γ(γ − 1)β , (2) where γ is the ratio of specific heats and β is a dimensionless con- stant which parameterises cooling, and is given by β = tcΩ, where tc is the cooling time. The scale height H is given by H = cs Ω and the midplane density is given by ρ0 = Σ 2H . (3) (4) (5) For a chosen M and outer radius R, we iterate our code until the surface density Σ produces the accretion rate we are attempting to match. A disc is susceptible to gravitational instability if the Toomre (1964) parameter Q ≈ 1.5 − 1.7 (Durisen et al. 2007). For our purposes, we impose the following condition at all radii in the disc, Q = csΩ πGΣ = 2, (6) c(cid:13) 2015 RAS, MNRAS 000, 1 -- 13 Directly observing continuum emission from self-gravitating spiral waves 3 where G is the gravitational constant. With the chosen accre- tion rate and outer radius fixed, we then find Σ by guessing an M in Equation overly large value and iterating downwards until (1) matches our chosen M. For a given value of Σ, we obtain the local sound speed by rearranging Equation (6). This allows the cal- culation of the local scale height, and hence the midplane density. An equation of state table is then used to determine the opacity κ from this density and sound speed, and then the optical depth τ is estimated as τ = κΣ using Rosseland mean opacities from Bell & Lin (1994). The cooling rate is then (Hubeny 1990): Λ = 8σT 4 3τ , (7) where T is the midplane temperature and σ is the Stefan- Boltzmann constant. If some source of external irradiation is present, this is modified to Λ = 8σ(T 4 − T 4 irr) 3τ , (8) where Tirr is the temperature set by the external irradiation. The cooling time tc in both cases is then the thermal energy per unit area divided by the cooling rate: tc = 1 Λ c2 sΣ γ(γ − 1) . (9) Since in a quasi-steady state, energy dissipation in a disc is dom- inated by self-gravity (Lodato & Rice 2004), and assuming local thermodynamic equilibrium, the rate per unit time, per unit area at which the kinetic energy of rotation of the disc is dissipated into heat by viscosity (the dissipation rate) is equal to the cooling rate, given by Equation 7. This dissipation rate is R 2 D(R) = 1 2 νvisΣ ∂Ω ∂R , (10) where νvis is kinematic viscosity, and is expressed as (Shakura & Sunyaev 1973) νvis = αc2 s Ω . (11) This allows α to absorb the uncertainties of the pseudo-viscous properties of the disc. Since in a steady state, the cooling rate matches the dissipation rate, we can equate Equations (7), or (9) and (10), and then use Equation (11) to directly determine α. Fi- nally, since the accretion rate in a quasi-steady disc is given by Equation (1), we can check that our calculated M is within some tolerance of our imposed M. If not, we reduce Σ and repeat this M is within some tolerance of our chosen M. Therefore, the until surface density is determined iteratively at every radius and only needs to be integrated to determine the enclosed mass of the disc. The parameter space of our model is shown in Figure 2. For every accretion rate, and radius at that accretion rate, there is a corre- sponding disc mass. 2.3 3D Structure The above method only examines the radial parameter space. To examine structures such as spiral arms, the model must be devel- oped to manage azimuthal asymmetry. Additionally, since we later compute global 3D radiative transfer calculations, we also include a vertical density profile. We begin by defining a Cartesian grid onto which surface density as a function of radius is mapped. This c(cid:13) 2015 RAS, MNRAS 000, 1 -- 13 Figure 2. 2D contour lines of actual disc mass (around a 1 M(cid:12) star) as a function of accretion rate and radius for self-gravitating discs with no external irradiation. produces a 2D disc which is azimuthally uniform at each radius. It has been shown there is a relationship between the amplitude of the density perturbations and the strength of the cooling such that (Cossins, Lodato & Clarke 2009) (cid:32) (cid:33)2 δΣ Σ = 2 β 1 γ(γ − 1) (cid:32) (cid:33) 1 M(cid:102)M , (12) where  is a dimensionless proportionality factor known as the heat- ing factor, and the radial phase and Doppler-shifted phase Mach numbers are M = vp/cs and (cid:102)M = vp/cs. Cossins, Lodato & Clarke (2009) find an empirical relationship such that (cid:104)δΣ(cid:105) (cid:104)Σ(cid:105) ≈ β − 1 2 . (13) Since there is a relationship between α and β given by Equation (2), we impose a spiral perturbation dictated by our αgrav such that (Rice et al. 2011) (cid:104)δΣ(cid:105) (cid:104)Σ(cid:105) ≈ α1/2 grav, (14) (15) where δΣ/Σ is the fractional over-density. This is imposed sinu- soidally so that δΣ(φ) = −(cid:104)δΣ(cid:105) cos(mφ), where m is the azimuthal wavenumber selected by the user. For self-gravitating discs with mass ratios similar to what we'll be con- sidering, this is typically high, so we select 10. We have assumed that the maximum over-density is equal to the average over-density, with φ the phase difference between the azimuthal location of the spiral θspiral and the azimuthal position θx,y in the disc: φ = θspiral − θ(x,y). The maximum over-density then occurs when the (x,y) position is coincident with the position of the spiral. The negative sign in Equation (14) simply forces the maxima to occur coincident with the position of the spiral. A logarithmic spiral of the form (16)  r a , θspiral = 1 b log (17) 4 Cassandra Hall, Duncan Forgan, Ken Rice, Tim J. Harries, Pamela D. Klaassen and Beth Biller was used, where a and b are constants (in this case a = 13.5 and b = 0.38), as this most closely matches the shape of spirals seen in self-gravitating discs from simulation data (see Figure 1). This is somewhat arbitrary, since tweaking parameters can give a close fit with an Archimedean spiral. However, it is not our intention to exactly match morphology. To create accurate skymodels in terms of brightness to put into the ALMA simulator, the vertical density profile must be carefully considered, since the total intensity at the surface of the disc is de- pendent upon the amount of emission and absorption which has occurred between the surface and the midplane. We calculate the density as a function of z using the expression for density in a self- gravitating disc (see e.g. Spitzer (1942) for a full derivation)  ρ(z) = ρ0 , (cid:17) cosh2(cid:16) z 1 Hsg (18) (19) where the self gravitating scale height Hsg is given by Hsg = c2 s πGΣ . It is worth noting that Hsg is approximately equal to H in Equation (4) since Hsg H = csΩ πGΣ = Q = 2. (20) 2.4 Radiative and Molecular Line Transfer Code: TORUS The TORUS radiation transport code (Harries et al. 2004; Kuro- sawa et al. 2004; Haworth et al. 2015) determines radiative equilib- rium in a dusty medium using the Monte Carlo (MC) photon packet method first described by Lucy (1999). Temperatures, densities and dust properties are stored on a three-dimensional adaptive mesh, re- fined in such a way that large gradients in opacity are well resolved. Here we use a cylindrical adaptive mesh, in which the cells are sec- tors of hollow cylinders. When additional resolution is required a cell may be subdivided into four (or eight) children by splitting the radial extent and height of the cell into two, and (optionally) by splitting the azimuthal extent of the sector into two. We use a simple thermal model for the disc geometry, and then calculate a radiative-equilibrium model using TORUS. The radia- tion equilibrium calculation is iterative and full details are given in the above references. Briefly, the radiation field of the protostar is modelled using N photon packets which are allowed to propa- gate through the grid, undergoing a random walk of scatterings, or absorptions and re-emissions, until they escape the computational domain, at which point estimates are made of the absorption rate in each cell. The dust temperatures in the grid are then calculated on a cell-by-cell basis by assuming radiative equilibrium, and the next iteration of the photon loop is performed (using the updated dust temperatures). Once the temperatures have converged it is possible to calculate spectral energy distributions and continuum images for arbitrary viewing angles using the MC method. The TORUS code has been extensively benchmarked and shows good agreement with other independently developed radiative transfer codes (Pinte et al. 2009). For our main radiative transfer results (in Section 3.3) we assume typical values for a pre-main-sequence star, with central source mass of M∗ = M(cid:12), R∗ = 2.325 R(cid:12) and Teff = 4350 K. The dust in our model consists of Draine & Lee (1984) silicates, with a grain size distribution of −q for amin < a < amax, n(a) ∝ a where amin and amax are the minimum and maximum size of the dust grains (0.1 µm and 1.0 µm respectively), and the power-law exponent q is taken to be qism = 3.5 (Mathis, Rumpl & Nordsieck 1977). The dust density is 1% of the gas density. (21) 2.5 The ALMA Simulator The emission maps generated by TORUS are used as inputs to the ALMA simulator built into CASA (ver 4.3) (McMullin et al. 2007). Disc sizes and fluxes are scaled to a distance of ∼ 140 pc (i.e. in Taurus), and we show a set of comparison images at ∼ 50 pc (i.e. in TW Hydrae). Multiple simulations were conducted varying the array size, and therefore imaging resolution and sensitivity, to ensure the op- timum balance between resolution and sensitivity. Built-in noise sources such as atmospheric transmissions were included in the simulations run at varying precipitable water vapour (PWV) lev- els. Typical PWV levels appropriate for observing in the differ- ent ALMA bands were used, specifically: 2.784, 1.262, 1.262 and 0.472 mm for simulated observations at 220, 345, 460 and 680 GHz, respectively. These are consistent with those used, for in- stance, in Dipierro et al. (2014). 3 RESULTS Our results section is broken into four parts. We begin with the basic results from our analytic geometry (Section 3.1), showing that for a given central star mass and disc outer radius, there are regions of parameter space in which the disc cannot exist in a quasi-steady, self-gravitating state. We then show that such discs may exist in these regions of parameter space if they are irradiated with some external source, but a moderate amount of irradiation will remove spiral features. In Section 3.2, we compare the synthetic ALMA image re- sults of our analytical model to image results which used SPH ge- ometries. We show that the assumptions made in SPH simulations probably cause larger contrast in the inner regions of discs. We do, however, reproduce the basic results of Dipierro et al. (2014) using a constant β approach. In Section 3.3, we present results showing the conditions re- quired to directly observe self-gravitating spiral structures with ALMA. We entered a range of accretion rates, and found that even at the maximum accretion rate a 100 AU disc can sustain with- out fragmenting, non-axisymmetry due to disc self-gravity is just detectable at 680 GHz at a distance of 140 pc (i.e. in Taurus). At 50 pc (i.e. in TW Hydrae), it is significantly easier to detect spiral structure, and the features are discernible at 220 GHz. Finally, we make only parameter space comparisons with three observed systems in Section 3.4. We show that for all three systems, it seems unlikely that spiral features which have recently been im- aged in the disc are due to disc self-gravity, unless the disc mass has been significantly underestimated. 3.1 Basic Model We find that for a given radius and host star mass, with no external irradiation considered, there is a maximum accretion rate that any self-gravitating, quasi-steady disc can sustain. Above this accretion c(cid:13) 2015 RAS, MNRAS 000, 1 -- 13 Directly observing continuum emission from self-gravitating spiral waves 5 M = 2.8 × 10−7 M(cid:12) yr−1 M = 3.2 × 10−7 M(cid:12) yr−1 M = 3.6 × 10−7 M(cid:12) yr−1 n o i t a i d a r r I o N K 0 1 t a n o i t a i d a r r I K 0 3 t a n o i t a i d a r r I Figure 3. Logarithmic surface density maps of discs with accretion rate increasing from left to right. Top row: For a 1M(cid:12) central star and no external M = 2.8 × 10−7M(cid:12) yr−1. Any higher, and irradiation, the maximum accretion rate a 100 AU disc can sustain without the outer regions fragmenting is the disc truncates in the outer regions as it becomes unstable to collapse and begins to fragment. Middle row: same three accretion rates, but with external irradiation at 10K added. Although the discs look similar, they vary very slightly in mass. The irradiation is just enough to prevent truncation, without removing spiral features. Bottom row: same discs but with irradiation at 30K added. These discs are also prevented from fragmentation in the outer regions, however the irradiation is also sufficient to remove spiral features completely. rate, the disc begins to truncate as the outer parts become suscep- tible to fragmentation (Rafikov 2005; Forgan & Rice 2011). This is illustrated in the top panel of Figure 3, which shows logarithmic surface density plots of three discs with increasing accretion rate from left to right. Note that to decrease the disc radius to a quar- ter of its original size only requires an increase in accretion rate of around 30%. When more mass is added to the disc, the sound speed (and subsequently temperature from our equation of state) increases since in our model it is set by Equation (6). In the cool outer parts of the disc, the Rosseland mean opacity is related to temperature by κ ∝ T 2 (Whitworth et al. 2010). Therefore our cooling rate now has a dependence Λ ∝ T 2. This increased local cooling rate causes a decrease in local cooling time. In order to maintain the disc in c(cid:13) 2015 RAS, MNRAS 000, 1 -- 13 a quasi-steady state, this local radiative cooling must be balanced by viscous shock heating from the spiral arms, therefore the local αgrav increases to redress the balance in the disc. However, this quasi-steady, self-gravitating torque saturates at around α ∼ 0.1 (Gammie 2001; Rice, Lodato & Armitage 2005) and we expect the disc to fragment, producing bound objects. Since this region of parameter space is not what we are interested in, we set the surface density here to 0. However, a small amount of irradiation, say at 10K, can change the surface density profile of a disc. How this changes de- pends upon the accretion rate of the disc, and therefore how close M was initially to the fragmentation/truncation boundary in the ab- sence of irradiation. The middle panels of Figure 3 show discs with an accretion 6 Cassandra Hall, Duncan Forgan, Ken Rice, Tim J. Harries, Pamela D. Klaassen and Beth Biller plitudes decrease. In our calculations, the α is composed of two parts, α = αgrav + αvisc, (22) 1 grav, 2 where αvisc is the viscous component of the stress due to magne- torotational instability (MRI), which we set to αvisc = 0.01 (see e.g. Kratter, Matzner & Krumholz 2008). Since the perturbation strength of the spiral amplitudes is given by (cid:104)δΣ(cid:105) (23) (cid:104)Σ(cid:105) = α when α → 0 then (cid:104)δΣ(cid:105) (cid:104)Σ(cid:105) → 0. This happens when the midplane temperature T in Equation (8) tends to T 4 irr. In this case, cool- ing rate Λ → 0, so cooling time tc → ∞ and β → ∞. This means αgrav → 0 by Equation (2), and (cid:104)δΣ(cid:105) (cid:104)Σ(cid:105) → 0. In this limit, cooling is balanced by irradiation and αvisc, so αgrav is no longer needed for thermal equilibrium. Essentially, increasing temperature provides extra pressure support against gravitational collapse. In our geometry, however, we do not consider infalling mass from a nascent cloud. This could potentially reverse our result of small amounts of irradiation halting fragmentation, as this infalling mass causes a positive rate of change of local mass, decreasing the Jeans mass and potentially encouraging fragmentation, provided that Q remains constant. Lowering the accretion rate (and therefore disc mass), changes the effect of a given amount of external irradiation. Figure 4 illus- M = 2.0 × 10−7 trates this. It is a disc with an accretion rate of M(cid:12) yr−1, approximately a third lower than in Figure 3. In this case (Figure 4) external irradiation has added additional mass to the outer parts of the disc, as the equilibrium disc structure is now more massive for a given accretion rate and radius, although in this case it does not meet the criteria required for it to fragment (which requires that the local α (cid:38) 0.1). In this disc, the external irradia- tion is maintaining Q ∼ 2 with little dissipation of the gravitational instability. In reality, we should expect Q to increase beyond the marginal limit of self-gravity in the outer parts of such discs. We should bear in mind that this is a simple model in which no infall is considered. Since, in general, adding mass to the outer regions of a self-gravitating disc encourages fragmentation (Kratter et al. 2010; Vorobyov & Basu 2010; Kratter & Murray-Clay 2011; Forgan & Rice 2012), it seems that the Jeans criterion for fragmen- tation may be satisfied at relatively high accretion rates (of order ∼ 10−7 M(cid:12) yr−1) in the presence of sufficiently small (∼ 10 K) irradiation. Figure 5 is a contour plot of the amount of external irradia- tion required to prevent fragmentation of the disc as a function of accretion rate and radius. Below ∼ 3.2 × 10−7 M(cid:12) yr−1, in our local viscous approximation, any disc will be able to regulate it- self against collapse. Above this, differing amounts of irradiation can either prevent fragmentation and totally remove spiral features from the disc, or prevent fragmentation whilst allowing spiral fea- tures to exist. As accretion rate increases, higher temperatures are required to prevent the disc from fragmenting; additionally, beyond 60 AU the determining factor in whether spirals are present or not is temperature, rather than a combination of temperature and accre- tion rate. 3.2 Comparison with imposed constant β In our model, unlike in SPH simulations by Dipierro et al. (2014) and Lodato & Rice (2004), we do not artificially impose a constant c(cid:13) 2015 RAS, MNRAS 000, 1 -- 13 Figure 4. Disc with M = 2.0 × 10−7 M(cid:12) yr−1, 33% lower than the maximum it can sustain without truncating. It is irradiated at 10 K, and this small amount of external irradiation has altered the surface density structure of the outer parts of the disc, as the equilibrium disc structure is now more massive for a given accretion rate and radius. The "spikey" features on the boundary between the outer and inner parts of the disc are a numerical arte- fact due to a small change in surface density from the resolution limit of the grid, and is more pronounced in log space. rate that was initially at the fragmentation boundary before irradi- ation at 10K was added. In this case, the irradiation prevents trun- cation whilst preserving spiral structure. The disc mass varies very slightly, and this is consistent with previous examinations of disc mass under irradiation (Forgan & Rice 2013a) since the accretion rate is only changing slightly, rather than by an order of magnitude. We therefore find that a small amount of external irradiation can prevent fragmentation, whilst preserving the spiral features in the disc. On the other hand, we do find that spiral features in a self- gravitating disc of any accretion rate are erased by a moderate amount of external irradiation, say at 30K. This is demonstrated by the bottom panels in Figure 3, which shows that for a disc with a radius of 100 AU, irradiation at a temperature of 30K can erase non-axisymmetric structure from the disc. It is unsurprising that applying sufficient external irradiation to our systems wipes out spiral structure, as external irradiation has been found to have a stabilising effect on marginally unstable discs (Kratter & Murray- Clay 2011; Rice et al. 2011; Forgan & Rice 2013a). When external irradiation of temperature Tirr is present, the local cooling rate in the disc is reduced and is given by Equation (8). This decrease in cooling rate causes the cooling time tc = β/Ω to increase, which also causes local β to increase. If the disc is in thermal equilib- rium, we then parameterise the viscous shock heating required to balance this cooling as αgrav given by Equation (2), where αgrav can be thought of as an effective gravitational stress. It is easy to see that for an increased β, a smaller α is required to redress the balance. The α is then able to stay below the torque saturation limit αcrit ∼ 0.1 (Gammie 2001; Rice, Lodato & Armitage 2005), and the disc is able to stay in a quasi-steady self-gravitating state out to larger radii. Although in this manner, the disc is able to stay in a quasi-steady, self-gravitating state, the strength of the spiral am- Directly observing continuum emission from self-gravitating spiral waves 7 Realistic α Constant β z H G 0 2 2 z H G 0 8 6 Figure 5. Contour plot showing the minimum temperature in (K) of external irradiation required to halt fragmentation as a function of accretion rate and radius. Shaded region with dashed contours shows where both fragmenta- tion is suppressed and spiral features are erased. The unshaded region with dotted contours shows where added irradiation prevents fragmentation but preserves spiral structure. gravitational αgrav by imposing a constant β. Doing so implies that the spiral amplitude strength (cid:104)δΣ(cid:105)/(cid:104)Σ(cid:105) ≈ α1/2 grav is constant at all radii. In our model we take a more realistic approach and allow α to vary, requiring that our disc remain marginally unstable to non- axisymmetric perturbations (i.e. Q = 2), and that the cooling rate at each radius depends on the local conditions. In such a quasi-steady, self-gravitating disc we expect α to increase with increasing radius (Rice & Armitage 2009). The basic consequence of this is that the assumption of a con- stant β means that there will clearly be regions where the α values will be quite different to what would be the case if more realistic assumptions were used. In particular, the α values would proba- bly be significantly larger in the inner regions, which will cause a greater contrast than would be the case were more realistic assump- tions used. Figure 6 compares two systems, both with the same total mass, but one determining α realistically (left hand panel) and the other assuming that β (and hence α) is constant, fixed at β = 6. The reason we choose the same disc mass, rather than the same accretion rate, is that the total disc flux FD is related to total disc mass MD by FD = 2κ(ν)kBν2TdustMdisc D2c2 , (24) where κ(ν) is opacity, kB is the Boltzmann constant, c is the speed of light and D is the distance to the object. The disc mass is the same in both cases (q = 0.2), but holding β fixed alters the ge- ometry so the equilibrium accretion rate for a given disc mass differs. Therefore M = 5 × 10−8 M(cid:12) yr−1 for realistic α and M = 2.8× 10−7 M(cid:12) yr−1 for constant β yield the same total disc mass. In the fixed β (and therefore fixed α) scenario, the strength of the spiral amplitudes are larger throughout the disc and so they are easily detectable to 3σ confidence level even at ν = 220 GHz. At ν = 680 GHz, the interarm regions are a low enough temperature c(cid:13) 2015 RAS, MNRAS 000, 1 -- 13 Figure 6. Top 4 panels: Synthesised ALMA images for two R = 100 AU discs with different geometries. Both discs have a central star mass of 1 M(cid:12) and central star radius 1 R(cid:12), with Tsource = 6000K, imaged at a distance of 50 pc for clarity. Beam size which gives the best compromise resolution and sensitivity is selected for each disc, details are given in Table 1. Left column has realistic α, whereas right column has in imposed constant β (and therefore α) value. Contours are at 3, 4, 5, 6, 9, 12, 15 & 18 × the RMS in each image. Each image has been scaled to best show the spiral features M = 5 × 10−8 (if present). Disc mass is the same for both (q = 0.21), M(cid:12) yr−1 for realistic α and M = 2.8 × 10−7 M(cid:12) yr−1 for constant β - and, hence, constant α. In both cases, the spiral arms are more clear for β = 6. Bottom panel: Spiral amplitude strength vs radius for the two discs. Red dots are "realistic" α and blue crosses are fixed α(β). to broach the Wien limit, thus increasing the contrast ratio between the spiral arms and the inter-arm regions. The first thing that we should stress is that using our semi- analytic model with fixed beta, we are able to reproduce the basic results of Dipierro et al. (2014). We find that extended structure is detectable to 3 × the RMS noise, and that the fluxes agree to the same order of magnitude in both the realistic α case and imposed constant β (hence α) case, giving us some confidence that our gen- eral method is reasonable. We have, however, assumed our dust grains are dynamically well coupled to the gas such that ρdust is linearly proportional to 8 Cassandra Hall, Duncan Forgan, Ken Rice, Tim J. Harries, Pamela D. Klaassen and Beth Biller ρgas. Dipierro et al. (2015) showed that dust trapping occurs in the spiral arms in self-gravitating discs, however this predominantly occurs for particles with sizes of a centimetre or more (Rice et al. 2004). Therefore, these overdensities are best probed at frequencies that are not considered here (∼ 10 GHz), and so should not change our results. However, what Figure 6 also shows is that in the fixed β sce- nario, the strength of the spiral amplitudes are larger throughout the disc than in the realistic α case. At ν = 680 GHz, this translates to more clearly discernible spiral structure throughout the disc. We could, of course, choose a larger β value, or increase the mass in the realistic alpha case, so that there were at least regions where the amplitudes were comparable. However, it is clear that even in such cases, the constant β assumption would produce unrealistic amplitudes in the inner parts of these discs. For the rest of this paper we allow the effective gravitational stress αgrav to vary as it would do in a "realistic" disc. 3.3 ALMA Images In this section we present results illustrating the conditions under which we may be able to directly observe self-gravitating spiral density structures using ALMA. We assume that all discs have outer radii of 100 AU and follow the procedure described in Sec- tion 2 to determine the amplitude of the spiral density waves, the continuum emission from the disc, and what we would then expect to be observed by ALMA. Figure 7 shows six synthetic ALMA images of three 100AU discs around a 1.0M(cid:12), 2.325 R(cid:12) star, with a surface temperature of T = 4350 K, and with three different accretion rates. All discs are imaged at a distance of 140 pc. Each disc is observed at 220 GHz (top row) and 680 GHz (bottom row). Observing parameters are given in Table 1, and physical parameters are given in Table 2. From left to right, the accretion rates are M = 2.8 × 10−7 M = 1.0 × 10−7 M(cid:12) yr−1 and M = 5.0 × 10−8 M(cid:12) M(cid:12) yr−1, yr−1. Figures 7a and 7d depict a disc with the maximum accretion rate M = 2.8 × 10−7 M(cid:12) yr−1 a disc of outer radius R = 100 AU, with no external irradiation, can sustain without fragmenting. Non-axisymmetric structure is only visible at the higher frequency of 680 GHz. We wish to stress at this point that this is the absolute maximum accretion rate that this disc, with this particular set of parameters and no external irradiation, can sustain. For a sufficiently high accretion rate, and therefore disc-to-star mass ratio q, we reproduce the results of Dipierro et al. (2014) of increasing contrast with increasing frequency, as the Planck law in the interarm regions falls into the Wien limit due to the low tem- perature. Additionally, Figure 7 shows that as accretion rate (and there- fore disc mass) is decreased, the central part of the disc remains de- tectable to at least the 3σ level, but the spiral structure is simply not detectable with ALMA at 220 GHz (or longer wavelengths). Non- axisymmetry is, however, noticeable at higher frequency (shorter wavelengths) since the lower temperatures in the inter-arm regions means the Planck law is in the Wien limit for these frequencies, reducing the intensity of emission. The images in Figure 7 show the difficulty faced when deter- mining the presence of GI in a protostellar disc. Since a distance of 140 pc corresponds to a typical star forming region (Taurus), these are not the most optimistic results. Figure 8 shows the same disc, M = 2.8 × 10−7 M(cid:12) yr−1, imaged at with an accretion rate of 140 pc (220 GHz and 680 GHz), and 50 pc (at 220 GHz and 680 GHz). At a closer distance, it is significantly easier to detect spiral 140 pc 50 pc z H G 0 2 2 z H G 0 8 6 Figure 8. All discs are an outer radius of 100 AU, imaged at 220 GHz and 680 GHz, at a distance of 50pc and 140 pc. M∗ = 1M(cid:12), R∗ = 2.325 R(cid:12) and T∗ = 4350 K. Right column shows the disc imaged at a distance of 50 pc, as in the TW Hydrae region. Left column shows discs imaged at a distance of 140 pc, as in the Taurus region. Contours are at intervals of the RMS, given in Table 1, beam sizes are also in Table 1. structure in the disc, as shown in the images at 50 pc. At a distance of 50 pc, the spiral structure in the outer part of the disc is also detectable at 220 GHz, whereas it is not detectable at 140 pc. It becomes more difficult to detect spiral structure with de- creasing accretion rate and disc mass. Since the strength of our spi- ral amplitudes are determined by δΣ/Σ ≈ α1/2, when the accre- tion rate is lowered, so are the spiral amplitudes. The conclusion we can draw from this examination of parameter space is that in quasi- steady, self-gravitating discs, for any given disc radius and host star mass, there exists a narrow range of accretion rates in which the outer part of the disc does not begin to fragment, but for which the spiral structure is detectable. Additionally, even if the disc is within this parameter space, the distance to the object and the frequency of the observations may also determine the likelihood of spiral struc- ture being detected. With this in mind, it is prudent to caution against diagnos- ing directly imaged non-axisymmetric structure as due to disc self- gravity, unless the disc in question is sufficiently massive. That our discs of lower disc mass/accretion rate fail to produce detectable spiral features appears to conflict with the result found by Dipierro et al. (2014). However, as previously mentioned, our model allows both the cooling time and α to vary locally, so the relative strengths of our perturbations are much less in the outer part of the disc than they would be in an SPH simulation where β is fixed at some rela- tively low value. 3.4 Comparison With Observed Systems Currently, there aren't any observations that are directly compa- rable to what we present here. There are, however, some systems with spiral features typically observed in NIR scattered light. We consider the properties of such systems and compare them to what c(cid:13) 2015 RAS, MNRAS 000, 1 -- 13 Directly observing continuum emission from self-gravitating spiral waves 9 Figure ν (GHz) Distance (pc) Beam size (asec) RMS (µJy beam−1) Contours (× RMS) Integration time (s) PWV (mm) 6 7a 7b 7c 7d 7e 7f 8 220 220 680 680 220 220 220 680 680 680 220 220 680 680 50 50 50 50 140 140 140 140 140 140 140 50 140 50 0.23 × 0.17 0.13 × 0.10 0.13 × 0.11 0.13 × 0.10 0.0706 × 0.0601 0.0706 × 0.0601 0.0706 × 0.0601 0.0435 × 0.0331 0.0435 × 0.0331 0.0435 × 0.0331 0.0706 × 0.0601 0.135 × 0.102 0.0435 × 0.0331 0.0732 × 0.0539 141 82 1674 1732 20 22 20 196 269 253 20 60 196 720 3,4,5,6,9,12,18 3,4,5,6,9,12,18 3,4,5,6,9,12,18 3,4,5,6,9,12,18 3,5,7,9,12,15,18 3,5,7,9,12,15,18 3,5,7,9,12,15,18 4,7,10,13,15,20 3,5,7,9,12,15,18 3,5,7,9,12,15,18 3,5,7,9,12,15,18 4,7,10,13,15,18 4,7,10,13,15,20 4,7,10,13,15,20 1800 1800 7200 1800 7200 7200 7200 7200 7200 7200 7200 7200 7200 7200 2.784 2.784 0.472 0.472 2.784 2.784 2.784 0.472 0.472 0.472 2.784 2.784 0.472 0.472 Table 1. Image details for figures shown in this work. We detail frequency of the synthesised observation, whether a realistic or constant α was used in the image, the accretion rate, disc mass, the size of the beam, the noise of the image, the integration time used and precipitable water vapour (PWV) value. Figure α type M∗ (M(cid:12)) M (M(cid:12) yr−1) q (MD/M∗) 6 7a 7b 7c 7d 7e 7f Realistic Constant Realistic Constant Realistic " " " " " 1 1 1 1 1 1 1 1 1 1 N/A MWC 758 SAO 206462 HD 142527 Realistic " " 2 1.7 2 5 × 10−8 2.8 × 10−7 5 × 10−8 2.8 × 10−7 2.8 × 10−7 1.0 × 10−7 5.0 × 10−8 2.8 × 10−7 1.0 × 10−7 5.0 × 10−8 2.0 × 10−7 5.37 × 10−9 6.92 × 10−8 0.21 0.21 0.21 0.21 0.34 0.25 0.21 0.34 0.25 0.21 0.25 0.1 0.75 Table 2. Physical parameters of the discs used to create synthesis images in this work. Columns are figure number, whether a realistic or constant α was used, host star mass, accretion rate and disc-to-star mass ratio we suggest would be required if self-gravity is to be the source of the spiral features. 3.4.1 MWC 758 The transition disc around Herbig A5 star MWC 758 is located in the edge of the Taurus star forming region at a distance of 279+94−48 pc (van Leeuwen 2007). It is 3.5±2 Myr old (Meeus et al. 2012)and has a stellar mass of 2 M(cid:12) (Isella et al. 2010). The disc mass and radial extent are approximated from sub-millimetre observations as 10−2 M(cid:12) and ∼ 100 AU respectively (Andrews et al. 2011). The accretion rate is estimated as somewhere between 2 × 10−7 M(cid:12) yr−1 (Isella et al. 2008) and 1 × 10−8 M(cid:12) yr−1 (Andrews et al. 2011). The first near-IR (NIR) scattered light images, clearly showing the discovery of spiral arms, were given in Grady et al. (2013), ob- tained using Subaru and 1.1 µm Hubble Space Telescope/NICMOS data. The parameterised fit of the spiral arms was performed by Grady et al. (2013) following Muto et al. (2012). It is possible that such spirals are launched by a perturbing body, if so it would re- c(cid:13) 2015 RAS, MNRAS 000, 1 -- 13 quire a mass of ∼ 5 MJ, which is consistent with continued accre- tion onto the central star. Marino et al. (2015) combine VLA Ka and ALMA maps to show that the disc is clearly non-axisymmetric. The disc is fit with a steady state vortex solution to explain the spiral arms. The authors suggest that the compact emission in VLA Ka data is consistent with an accreting companion object such as a forming planet, which could also be responsible for the spiral arms imaged in scattered light. The companion planet scenario is consistent with simulations conducted by Dong et al. (2015b). Similarly, MWC 758 was imaged in scattered light by Benisty et al. (2015), using VLT/SPHERE to achieve a higher resolution than previously achieved. The spirals arms were again modelled using density wave theory, with two planetary companions launch- ing the spiral arms. Although the spirals are interpreted as being due to planetary companions, other mechanisms, such as GI, can launch spiral waves with low m modes that are capable of match- ing these observed features, as shown by Dong et al. (2015a). The measured disc mass (10−2 M(cid:12)) of MWC 758 is probably too low to trigger gravitational instabilities (see e.g. Gammie 2001), how- 10 Cassandra Hall, Duncan Forgan, Ken Rice, Tim J. Harries, Pamela D. Klaassen and Beth Biller (a) 2.8 × 10−7M(cid:12) yr−1, RMS = 20 µJy beam−1. (b) 1 × 10−7M(cid:12) yr−1, RMS = 22 µJy beam−1. (c) 5 × 10−8M(cid:12) yr−1, RMS = 20 µJy beam−1. z H G 0 2 2 z H G 0 8 6 (d) 2.8 × 10−7M(cid:12) yr−1, RMS = 196 µJy beam−1. (e) 1 × 10−7M(cid:12) yr−1, RMS = 269 µJy beam−1. (f) 5 × 10−8M(cid:12) yr−1, RMS = 253 µJy beam−1. Figure 7. Synthesised ALMA images for R = 100 AU discs with accretion rate decreasing from left to right. Top row is at 220 GHz, bottom row is at 680 M = 2.8 × 10−7M(cid:12) yr−1. Contours GHz. All discs are imaged at 140 pc. Left column is the maximum accretion rate it can sustain without fragmenting, are at multiples of the RMS given in Table 1. Beam size is in bottom left corner, beam details are given in Table 1. Geometry details are given in Table 2. Below M = 5.0 × 10−8M(cid:12) yr−1, at 220 GHz flux from the disc is low enough that thermal noise dominates, so the central region (inner 20 AU) is detectable, but not the extended non-axisymmetric structure. At 680 GHz, asymmetry is noticeable, but the spiral arms are not clearly defined. ever, as discussed in our introduction, there are large uncertainties in the ratio of dust to gas and there is evidence that T-Tauri disc masses have been systematically under-estimated. Observations have revealed a complex morphology of the disk of MWC 758. To understand the origin of these spiral features, both modelling and high resolution images in the sub-mm with ALMA is needed. Scattered light traces the surface variations in a disc (a valid assumption for vertical isothermal hydrostatic equilibrium), whilst to probe structures near the midplane it is preferable to use longer wavelengths with high spatial resolution. In this work, we model MWC 758 as if it is self-gravity that is responsible for these spirals, to see if it is indeed the likely origin of these features. We simply assess whether it exists in the parameter space required for self-gravity to exist. We enter into our model a host star with mass 2 M(cid:12), a disc outer radius of 100 AU and an accretion rate of 2 × 10−7 M(cid:12) yr−1. In order for the disc to be in a quasi-steady, self-gravitating state for these specific parameters requires a disc-to-star mass ratio of q ∼ 0.25. This gives a total disc mass that is over an order of magnitude larger than that given by Andrews et al. (2011). This means that either: (i) The spirals are due to self-gravity, and the mass of the disc sur- rounding MWC 758 has been underestimated by a factor of 50. Even if this is the case, Dong et al. (2015a) have recently shown that self-gravitating spiral arms obey m ∼ 1/q, suggesting that the expected dominant m-mode would be 4, if the spirals are due to self-gravity. However, for m = 2 spiral modes to dominate typically requires q (cid:38) 0.5, and that the accretion rate be high, of order ∼ M ≈ 10−6 M(cid:12) yr−1 (Dong et al. 2015a). Such a disc would have non-local angular momentum transport (Forgan et al. 2011), and as such would not be well-described by the viscous approximation in our analytical model. (ii) The disc is self-gravitating, but the accretion rate is much lower than any of the measured values given by Andrews et al. (2011) for MWC 758, and the measured disc mass is correct. Figure 9 shows that for a disc around a host star of 2 M(cid:12) to have a mass of 10−2 M(cid:12) (or equivalently q ∼ 0.005) requires that the accre- tion rate be of order ∼ 10−10 M(cid:12) yr−1. If this is the case, it is highly unlikely that spiral structure would be detectable since the αgrav, and therefore perturbation strength of the spiral, would be incredibly low. (iii) Both the disc mass of MWC758 and the accretion rate are accu- rate. The spiral structure visible is due to some other mechanism, perhaps planet - disc interaction as discussed in Benisty et al. (2015), and not due to self-gravity. Accretion is therefore driven by something other than GI, such as MRI. c(cid:13) 2015 RAS, MNRAS 000, 1 -- 13 Directly observing continuum emission from self-gravitating spiral waves 11 scription of angular momentum transport, but this would also re- quire the system to have very different properties to those observed. 3.4.2 SAO 206462 SAO 206462 is an isolated 1.7 M(cid:12) Herbig Ae/Be star at a distance of 142 pc in the constellation Lupus (Müller et al. 2011). It has a ∼ 10−3 M(cid:12) disc (Thi et al. 2001) and an accretion rate of 5.37×10−9 M(cid:12) yr−1 (Garcia Lopez et al. 2006) and an outer radius of 140 AU. Scattered light observations in NIR have revealed spiral structure in the outer disc (Muto et al. 2012; Garufi et al. 2013), and sub- mm ALMA observations have revealed large-scale asymmetries in the dust continuum (Pérez et al. 2014). These asymmetries have been fit using a vortex prescription, following Regà aly et al. (2012) by Pérez et al. (2014), however, those authors concluded that the vortex prescription did not reproduce every observed feature, and significant residuals remained which coincided with the spiral arms seen in H-band scattered light. Although the disc mass is probably too low to trigger grav- itational instabilities, as is the case with many T-Tauri stars there is evidence for systematic underestimation of the disc mass. We model SAO 206462 as if disc self-gravity is responsible for the spiral features present in the disc. Using our model, to match the accretion rate of SAO 206462 requires a disc-to-star mass ratio of q ∼ 0.1 in order for the disc to be in a quasi-steady, self gravitating state. Spiral arms which have GI as their origin make additional de- mands on a system that in this case do not seem to be fulfilled. The spirals must be compact (on scales less than ∼ 100 AU), the disc must be massive (q (cid:38) 0.25) and the accretion rate must be high (Dong et al. 2015a). This leaves us with the following available conclusions: (i) The disc mass has been underestimated by several orders of magnitude, and the disc for SAO 206462 is actually well within the self-gravitating regime. However, even if this is the case, such a disc would not produce clear spiral structure due to the low αgrav and therefore spiral amplitude. (ii) The accretion rate is much lower than that measured for SAO 206462, and the disc mass measured is correct. This would fur- ther decrease the amount of flux from the disc and again cause difficulty observing it. (iii) Both the disc mass and accretion rate are accurate, and the spiral features are not due to disc self-gravity. 3.4.3 HD 142527 The transition disc HD142527 has been observed in the near-IR, and has been revealed to have a unique morphology, appearing to consist of two bright arcs facing each other and one spiral arm (Fuk- agawa et al. 2006). The central star's mass and age are respectively estimated at 1.9 − 2.2M(cid:12) and 1 − 12Myr (Fukagawa et al. 2006; Verhoeff et al. 2011). It has an accretion rate of 6.92 × 10−8M(cid:12) yr−1 (Garcia Lopez et al. 2006) and the estimated flow rate of gas in the gap in the disc is between 7 × 10−9 and 2 × 10−7M(cid:12) yr−1 (Casassus et al. 2013). The total disc mass has been measured from gas-to-dust ratios as ∼ 0.1 M(cid:12) (Verhoeff et al. 2011). Spiral arms have been imaged in 12CO J = 2 − 1 and J = 3 − 2 using ALMA by Christiaens et al. (2014), who placed lower limits on the mass of each spiral arm at ∼ 10−6M(cid:12). These features were interpreted as an acoustic wave launched by a planet Figure 9. Contour plot of disc-to-star mass ratio for accretion rate and radius, plotted in the range required for a host star of 2 M(cid:12) to have a quasi-steady, self gravitating disc of mass ratio q = 0.005, or disc mass MD = 0.01 M(cid:12). The position of MWC 758 is marked by a red cross. In order for the system to be in a self-gravitating, quasi-steady state with a measured disc mass of MD = 0.01 M(cid:12) requires an accretion rate of ∼ 10−10 M(cid:12) yr−1. Gravitational instabilities are certainly capable of producing structures which match the morphologies of observed low m-mode systems. However, it also makes demands on the system that in the case do not appear to be met, i.e. that disc-to-star mass ratio and accretion rate are very high. Something else to bear in mind is that our analytic models make assumptions that are likely no longer valid in high mass (q (cid:38) 0.5) discs, in which global (m ∼ 2) spiral modes dominate. When global torques are induced, the angular momentum transport is no longer local, and the semi-analytic models using a local viscous approximation are no longer justified. Additionally, this semi-analytic model uses the midplane cool- ing time to determine the effective gravitational α. For massive discs, this will be largely underestimated compared to the actual α value in a global, radiatively cooling disc (Forgan et al. 2011). Given that the spirals in MWC 758 appear global in nature might suggest that we can't use our semi-analytic model in this compari- son. However, producing such global spirals via GI would require disc properties even more discrepant than our model suggests, and so the basic conclusion would remain unchanged. This should serve as a word of caution to the analysis of future observations of discs with non-axisymmetric structure. Modelling non-local discs in the local approximation will return discs with spiral amplitudes far lower than would realistically be present. On the other hand, such discs would be extremely massive and have high accretion rates. Not only is it unlikely that they would be con- fused for lower mass discs, they will also remain in this phase for a very short time. The local approximation is therefore probably reasonable for anything that is likely to be observed by ALMA. Adding irradiation to MWC 758 would allow the system to maintain a larger disc mass. However, as shown in Section 3, this removes spiral features from the disc. Adding just enough irradia- tion to the disc so that spiral features are still observable does not change the result we get for a system with parameters matching those of MWC 758. If the spirals present in MWC 758 are due to disc self-gravity, then they cannot be modelled using a local pre- c(cid:13) 2015 RAS, MNRAS 000, 1 -- 13 12 Cassandra Hall, Duncan Forgan, Ken Rice, Tim J. Harries, Pamela D. Klaassen and Beth Biller (see e.g. Muto et al. 2012), however since it is now thought proba- ble that HD 142527 has a low-mass stellar companion (Biller et al. 2012), the spiral structures could certainly be tidally induced. GI is an alternative scenario which is able to replicate this grand design spiral structure, and since both Christiaens et al. (2014) and Fuka- gawa et al. (2013) find Q ∼ 2.0, there is evidence for gravitational instability being responsible for the spiral structure present. We assume a 2 M(cid:12) central star, and an accretion rate of 6.92× 10−8M(cid:12) yr−1. Since there is evidence that the disc may extend as far out as 600 AU in radius (Christiaens et al. 2014), we extend our disc out to 600 AU. To match the observed accretion rate of HD 142527 and ex- tend out to 600 AU, such a disc would require an incredibly high disc-to-star mass ratio of q ∼ 0.75. Such a disc would certainly have incredibly high global torques, and in reality would probably not survive in this quasi-steady, self-gravitating state. Therefore, the most likely explanation for the spiral structure observed in HD 142527 is not self-gravity. An alternative explanation for the spiral structure is tidal in- teraction due to its low mass companion (Mcompanion ∼ 0.1 M(cid:12)), which is potentially on an eccentric orbit around HD 142527 (Fuk- agawa et al. 2006; Baines et al. 2006; Biller et al. 2012; Close et al. 2014). 3.4.4 Conclusions from Observed Systems We examined the parameter space of three transition discs to deter- mine if the non-axisymmetric structure which has been imaged in those discs could feasibly be due to disc self-gravity. For all three systems, it seems unlikely, unless the disc mass has been signif- icantly underestimated. Even if the disc mass has been underes- timated, and the disc is self-gravitating, we may expect to see a different number of m-modes dominant in the disc. Self-gravity imposes additional requirements on a system which do not seem to be consistent with the parameters of these systems. 4 DISCUSSION AND CONCLUSION We performed an examination of the parameter space in which self- gravitating discs can exist, using a semi-analytical approach. We generated synthetic observations with the intention of investigating the range of accretion rates, disc masses and disc radii that would allow non-axisymmetric structure to be detected by ALMA. Our intention was not to reproduce the exact morphology of the ob- servations, but rather to understand the strength of a perturbation required to generate an observable spiral arm. Analytical models using a viscous prescription that assumes local angular momentum transport poorly describe systems which are dominant in the low m spiral modes. Modelling non-local discs in the local approximation returns spiral amplitudes far lower than would be present in reality. If a quasi-steady, self-gravitating disc can be described analytically using local transport, then there ex- ists a small range of accretion rates for a given radius where the gravitational stress is high enough to generate observable spirals, but not so high as to cause the outer regions of the disc to fragment. However, non-local transport only becomes significant in discs with masses above half the mass of the central star, and such discs will probably have very short lives (Lodato & Rice 2005; Rice, Mayo & Armitage 2010), so our analysis here is probably reasonable for anything that would be detected by ALMA. Another important factor is external irradiation. If the accre- tion rate is close to the fragmentation limit, a small amount of exter- nal irradiation (∼ 10 K) may prevent fragmentation with increasing accretion rate. If the accretion rate is well below the fragmentation limit, a small amount of irradiation (∼ 10 K) causes the surface density profile of the outer part of the disc to be restructured, as the equilibrium disc structure is now more massive for a given accre- tion rate and radius. If infall from a natal cloud is occurring, this could well be a trigger for fragmentation, as it is likely that in these regions the Jeans criterion would be satisfied. A moderate amount of irradiation (∼ 30 K) can suppress fragmentation up to higher accretion rates, but at the cost of non-axisymmetric structure. that Ultimately, our results suggest there is a relatively small range of parameter space in which a disc could be self- gravitating, not undergo fragmentation, and have spiral amplitudes large enough to be observable by ALMA. Broadly speaking, we would expect the disc mass to exceed 0.1 M(cid:12), the accretion rate to satisfy 10−7 (cid:46) M (cid:46) 10−6 M(cid:12) yr−1 and the outer radius to be not much more than 100 AU. Additionally, the observing frequency and distance to the source also plays a role. We are more likely to observe spiral waves at 680 GHz than at 220 GHz, and it becomes increasingly difficult as the source distance increases. Although self gravitating discs can certainly match the mor- phology of observed systems, they also impose strict additional conditions which may not be met. In essence, our analysis sug- gests that there is a relatively small region of parameter space in which self-gravity may produce observable spiral features. We would therefore caution against interpreting such features as be- ing due to disc self-gravity unless the disc is likely to fall into this region of parameter space. 5 ACKNOWLEDGEMENTS We are very grateful to the anonymous referee, who's comments greatly improved the final version of this paper. CH warmly thanks Giovanni Dipierro and Guillaume Laibe for their insightful and elucidating discussion with her concerning dust grains in this model. KR and BB gratefully acknowledge support from STFC grant ST/M001229/1. DF gratefully acknowledges support from the ECOGAL ERC advanced grant. Some calculations for this pa- per were performed on the University of Exeter Supercomputer, a DiRAC Facility jointly funded by STFC, the Large Facilities Capi- tal fund of BIS, and the University of Exeter, and on the Complex- ity DiRAC Facility jointly funded by STFC and the Large Facili- ties Capital Fund of BIS. TJH acknowledges funding from Exeter's STFC Consolidated Grant (ST/J001627/1). REFERENCES Andrews S. M., Wilner D. J., Espaillat C., Hughes A. M., Dulle- mond C. P., McClure M. K., Qi C., Brown J. M., 2011, ApJ, 732, 42 Baines D., Oudmaijer R. D., Porter J. M., Pozzo M., 2006, MN- RAS, 367, 737 Bell K. R., Lin D. N. C., 1994, ApJ, 427, 987 Benisty M. et al., 2015, A&A, 578, L6 Biller B. et al., 2012, ApJL, 753, L38 Casassus S. et al., 2013, Nature, 493, 191 Christiaens V., Casassus S., Perez S., van der Plas G., Ménard F., 2014, ApJL, 785, L12 c(cid:13) 2015 RAS, MNRAS 000, 1 -- 13 Directly observing continuum emission from self-gravitating spiral waves 13 Paczynski B., 1978, Acta Astronomica, 28, 91 Pérez L. M., Isella A., Carpenter J. M., Chandler C. J., 2014, ApJL, 783, L13 Pinte C., Harries T. J., Min M., Watson A. M., Dullemond C. P., Woitke P., Ménard F., Durán-Rojas M. C., 2009, A&A, 498, 967 Rafikov R. R., 2005, ApJL, 621, L69 Regà aly Z., Juhà asz A., Sà andor Z., Dullemond C. P., 2012, Monthly Notices of the Royal Astronomical Society, 419, 1701 Rice W. K. M., Armitage P. J., 2009, MNRAS, 396, 2228 Rice W. K. M., Armitage P. J., Mamatsashvili G. R., Lodato G., Clarke C. J., 2011, MNRAS, 418, 1356 Rice W. K. M., Lodato G., Armitage P. J., 2005, MNRAS, 364, L56 Rice W. K. M., Lodato G., Pringle J. E., Armitage P. J., Bonnell I. A., 2004, MNRAS, 355, 543 Rice W. K. M., Mayo J. H., Armitage P. J., 2010, MNRAS, 402, 1740 Ruge J. P., Wolf S., Uribe A. L., Klahr H. H., 2013, A&A, 549, A97 Shakura N. I., Sunyaev R. A., 1973, A&A, 24, 337 Spitzer, Jr. L., 1942, ApJ, 95, 329 Terebey S., Shu F. H., Cassen P., 1984, ApJ, 286, 529 Thi W. F. et al., 2001, ApJ, 561, 1074 Toomre A., 1964, ApJ, 139, 1217 van Leeuwen F., 2007, A&A, 474, 653 Verhoeff A. P. et al., 2011, A&A, 528, A91 Vorobyov E. I., Basu S., 2010, ApJL, 714, L133 Whitworth A., Stamatellos D., Walch S., Kaplan M., Goodwin S., Hubber D., Parker R., 2010, in IAU Symposium, Vol. 266, IAU Symposium, de Grijs R., Lépine J. R. D., eds., pp. 264 -- 271 Clarke C. J., 2009, MNRAS, 396, 1066 Close L. M. et al., 2014, ApJL, 781, L30 Cossins P., Lodato G., Clarke C. J., 2009, MNRAS, 393, 1157 Cossins P., Lodato G., Testi L., 2010, MNRAS, 407, 181 Dipierro G., Lodato G., Testi L., de Gregorio Monsalvo I., 2014, MNRAS, 444, 1919 Dipierro G., Pinilla P., Lodato G., Testi L., 2015, MNRAS, 451, 974 Dong R., Hall C., Rice K., Chiang E., 2015a, ArXiv e-prints Dong R., Zhu Z., Rafikov R. R., Stone J. M., 2015b, ApJL, 809, L5 Douglas T. A., Caselli P., Ilee J. D., Boley A. C., Hartquist T. W., Durisen R. H., Rawlings J. M. C., 2013, MNRAS, 433, 2064 Draine B. T., Lee H. M., 1984, ApJ, 285, 89 Dunham M. M. et al., 2014, ArXiv e-prints Durisen R. H., Boss A. P., Mayer L., Nelson A. F., Quinn T., Rice W. K. M., 2007, Protostars and Planets V, 607 Forgan D., Rice K., 2011, MNRAS, 417, 1928 Forgan D., Rice K., 2012, MNRAS, 420, 299 Forgan D., Rice K., 2013a, MNRAS, 430, 2082 Forgan D., Rice K., 2013b, MNRAS, 433, 1796 Forgan D., Rice K., Cossins P., Lodato G., 2011, MNRAS, 410, 994 Fukagawa M., Tamura M., Itoh Y., Kudo T., Imaeda Y., Oasa Y., Hayashi S. S., Hayashi M., 2006, ApJL, 636, L153 Fukagawa M. et al., 2013, PASJ, 65, L14 Gammie C. F., 2001, ApJ, 553, 174 Garcia Lopez R., Natta A., Testi L., Habart E., 2006, A&A, 459, 837 Garufi A. et al., 2013, A&A, 560, A105 Grady C. A. et al., 2013, ApJ, 762, 48 Harries T. J., Monnier J. D., Symington N. H., Kurosawa R., 2004, MNRAS, 350, 565 Haworth T. J., Harries T. J., Acreman D. M., Bisbas T. G., 2015, MNRAS, 453, 2277 Hubeny I., 1990, ApJ, 351, 632 Isella A., Natta A., Wilner D., Carpenter J. M., Testi L., 2010, ApJ, 725, 1735 Isella A., Tatulli E., Natta A., Testi L., 2008, A&A, 483, L13 Kratter K. M., Matzner C. D., Krumholz M. R., 2008, ApJ, 681, 375 Kratter K. M., Matzner C. D., Krumholz M. R., Klein R. I., 2010, ApJ, 708, 1585 Kratter K. M., Murray-Clay R. A., 2011, ApJ, 740, 1 Kurosawa R., Harries T. J., Bate M. R., Symington N. H., 2004, MNRAS, 351, 1134 Laughlin G., Bodenheimer P., 1994, ApJ, 436, 335 Lodato G., Rice W. K. M., 2004, MNRAS, 351, 630 Lodato G., Rice W. K. M., 2005, MNRAS, 358, 1489 Lucy L. B., 1999, A&A, 344, 282 Marino S., Casassus S., Perez S., Lyra W., Roman P. E., Avenhaus H., Wright C. M., Maddison S. T., 2015, ArXiv e-prints Mathis J. S., Rumpl W., Nordsieck K. H., 1977, ApJ, 217, 425 McKee C. F., Ostriker E. C., 2007, ARA&A, 45, 565 McMullin J. P., Waters B., Schiebel D., Young W., Golap K., 2007, in Astronomical Society of the Pacific Conference Se- ries, Vol. 376, Astronomical Data Analysis Software and Sys- tems XVI, Shaw R. A., Hill F., Bell D. J., eds., p. 127 Meeus G. et al., 2012, A&A, 544, A78 Müller A., van den Ancker M. E., Launhardt R., Pott J. U., Fedele D., Henning T., 2011, A&A, 530, A85 Muto T. et al., 2012, ApJL, 748, L22 c(cid:13) 2015 RAS, MNRAS 000, 1 -- 13
1507.07578
3
1507
2015-09-14T20:39:25
High Precision Photometry for K2 Campaign 1
[ "astro-ph.EP" ]
The two reaction wheel K2 mission promises and has delivered new discoveries in the stellar and exoplanet fields. However, due to the loss of accurate pointing, it also brings new challenges for the data reduction processes. In this paper, we describe a new reduction pipeline for extracting high precision photometry from the K2 dataset, and present public light curves for the K2 Campaign 1 target pixel dataset. Key to our reduction is the derivation of global astrometric solutions from the target stamps, from which accurate centroids are passed on for high precision photometry extraction. We extract target light curves for sources from a combined UCAC4 and EPIC catalogue -- this includes not only primary targets of the K2 campaign 1, but also any other stars that happen to fall on the pixel stamps. We provide the raw light curves, and the products of various detrending processes aimed at removing different types of systematics. Our astrometric solutions achieve a median residual of ~ 0.13". For bright stars, our best 6.5 hour precision for raw light curves is ~20 parts per million (ppm). For our detrended light curves, the best 6.5 hour precisions achieved is ~15 ppm. We show that our detrended light curves have fewer systematic effects (or trends, or red-noise) than light curves produced by other groups from the same observations. Example light curves of transiting planets and a Cepheid variable candidate, are also presented. We make all light curves public, including the raw and de-trended photometry, at http://k2.hatsurveys.org.
astro-ph.EP
astro-ph
Draft version 6 July 31, 2018 Preprint typeset using LATEX style emulateapj v. 2/16/10 5 1 0 2 p e S 4 1 . ] P E h p - o r t s a [ 3 v 8 7 5 7 0 . 7 0 5 1 : v i X r a C. X. Huang1, K. Penev1, J. D. Hartman1, G. ´A. Bakos1,2,3, W. Bhatti1, I. Domsa1, M. de Val-Borro1 HIGH PRECISION PHOTOMETRY FOR K2 CAMPAIGN 1 Draft version 6 July 31, 2018 ABSTRACT The two reaction wheel K2 mission promises and has delivered new discoveries in the stellar and exoplanet fields. However, due to the loss of accurate pointing, it also brings new challenges for the data reduction processes. In this paper, we describe a new reduction pipeline for extracting high precision photometry from the K2 dataset, and present public light curves for the K2 Campaign 1 target pixel dataset. Key to our reduction is the derivation of global astrometric solutions from the target stamps, from which accurate centroids are passed on for high precision photometry extraction. We extract target light curves for sources from a combined UCAC4 and EPIC catalogue -- this includes not only primary targets of the K2 campaign 1, but also any other stars that happen to fall on the pixel stamps. We provide the raw light curves, and the products of various detrending processes aimed at removing different types of systematics. Our astrometric solutions achieve a median residual of ∼ 0.127′′. For bright stars, our best 6.5 hour precision for raw light curves is ∼ 20 parts per million (ppm). For our detrended light curves, the best 6.5 hour precision achieved is ∼ 15 ppm. We show that our detrended light curves have fewer systematic effects (or trends, or red-noise) than light curves produced by other groups from the same observations. Example light curves of transiting planets and a Cepheid variable candidate, are also presented. We make all light curves public, including the raw and de-trended photometry, at http://k2.hatsurveys.org. Subject headings: K2, astrometry, photometry 1. INTRODUCTION The Kepler spacecraft ended its primary mission after the failure of two reaction wheels. The K2 mission uses the Kepler spacecraft to perform 80-day observations of selected fields in the ecliptic plane. This brings new op- portunities to study transiting planets around different stellar populations compared to the original Kepler field, such as clusters of young and pre-main sequence stars (Howell et al. 2014). K2 uses the remaining two reaction wheels, and solar radiation pressure, to maintain close to constant pointing of the spacecraft over the 80-day per-field observations. Currently, observations are performed with 21 modules, each module consisting of 4 CCD channels, yielding 76 channels (2 modules failed). Due to the limited band- width, only postage stamps containing proposed targets are downloaded. These postage stamps are typically 25×25 pixels in size (depending on the brightness of the targets and campaigns). These make up only less than 10% percent of the entire field of view (FOV). The ma- jority of stamps are observed at ∼ 30 minutes cadence. Typically, two Full Field Images (FFIs) are downloaded for the beginning and the end of campaign. However, the two reaction wheel mode also brings in new challenges for the data reduction processes. The spacecraft pointing is less stable compared to the primary mission, leading to a potential decrease in the photomet- ric precision. Although the disturbance from the solar pressure is mostly controlled by the two reaction wheels and the thruster firing (every 2 days), there is still a 1 Department of Astrophysical Sciences, Princeton University, Princeton, NJ 08544; email: [email protected] 2 Sloan Fellow 3 Packard Fellow 4 Hungarian Astronomical Association, Budapest, Hungary low frequency motion remaining, resulting in the targets drifting across the field of view. The extracted aperture photometry light curves are dominated by the systemat- ics induced by this drift pattern. Vanderburg & Johnson (2014) (hereafter VA14) minimized this drift system- atic by decorrelating the light curves with the motion of the spacecraft. They achieved a photometric preci- sion that is within a factor of two of the original Kepler photometry. Various other teams also developed their own tools to reduce the K2 data. Aigrain et al. (2015) used aperture photometry and a semi-parametric Gaus- sian process model to extract photometry from the K2 engineering data. Lund et al. (2015) presented K2P, a pipeline specifically designed for astrometric analyses. Foreman-Mackey et al. (2015) and Angus et al. (2015) proposed a method to analysis the K2 data without a general detrending process. There is, however, room for further improvements. VA14 reduction achieved the highest precision among all the past works, but only derived photometry for the pro- posed Kepler targets (not all targets falling on silicon), and are also known to have remaining systematic varia- tions affected by the spacecraft roll (Angus et al. 2015). Aigrain et al. (2015) and Lund et al. (2015) derive pho- tometry for all of the targets on silicon, but achieved slightly lower precision than VA14, especially for the bright stars. Here we present a new reduction of the K2 data drawing on techniques used in analysing data from ground-based surveys (Bakos et al. 2010, e.g.). We approach the K2 pixel file reduction with the fol- lowing steps: 1) improved astrometry for source centroid- ing and flux extraction; 2) photometric extraction for all the stars observed on the K2 postage stamps; 3) removal of first order systematics via a modified External Pa- rameter Decorrelation (EPD) procedure (broadly similar 2 to VA14); 4) further reduction of the shared systematic trends via an implementation of the Trend Filtering Al- gorithm (TFA) and semi-periodic stellar oscillations via cosine-filtering. The global astrometry step is key to this process -- it minimizes the effect of spacecraft drift on the aperture photometry, and allows us to accurately model the spacecraft motion for further detrending. In this paper, we describe our K2 photometry pipeline and the high precision light curves from the reduction of K2 Campaign 1. We introduce our effort of deriving ac- curate astrometry for the K2 observations, making use of the K2 FFIs, and present a revised K2 Campaign 1 tar- get list in §2. In §3, we present our aperture photometry method. In §4, we revisit our detrending techniques and present our light curves at different detrending stages. In §5, we compare our photometry with that of other studies. 2. ASTROMETRY 2.1. Background The first step of our reduction is to derive an accurate astrometric solution of the K2 data. Despite its large pixel scale (∼ 4′′) and PSF FWHM (5 − 6 ′′), the origi- nal Kepler mission turned out to be a great tool for accu- rate astrometry itself because of its extremely high SNR photometry and stable pointing. Monet et al. (2010) re- ported a preliminary astrometric solution precision, from the first few months of Kepler data, to be 0.001 pixel, nominal 4 mas. This high astrometric precision, and high stability of the centroid position, enabled the high photometric precision of the Kepler primary mission. Unlike the original Kepler Mission, the K2 stars typi- cally drift across the CCD plane at a speed of 1-3% of a pixel every 30 min. Since the 30 min K2 frame is com- posed of 270 short exposures of 6 s each, the final PSF is inevitably distorted, and neighbouring stars tend to become blended. Therefore, it is difficult to determine accurate centroids from source extraction alone. Thus, we use an external catalogue, namely the fourth United States Naval Observatory (USNO) CCD Astrograph Cat- alogue, UCAC4 (Zacharias et al. 2013), which has an as- trometric precision of 15-100 mas, to derive good astro- metric solutions for the K2 frames. A good astrometric solution does not only benefit the photometric precision, but also enable us to make max- imal use of the K2 observations. The K2 campaigns ob- serve targets proposed by the community, and each tar- get was then assigned a stamp of size 20-50 pixels across. This stamp size is much bigger than the original Kepler stamp size. In addition to the target, many other sources are observed in a typical K2 stamp. We provide position information and reduced light curves for all of the stars observed in the K2 stamps. We anticipate an improved planet yield from this approach, due to the larger num- ber of sources available, and the availability of the light curves of neighbouring stars, useful in blend analyses. We first derive a general astrometric solution using the Kepler FFIs and our custom developed astrometry soft- ware used for HATNet. This general solution is then used as an initial guess for the remaining stamp observations. We stitch all the K2 stamps together into a "Sparse FFI" (SFFI), with the unobserved regions masked. We fit for an astrometric solution to the SFFI, which is assumed to be a low order polynomial distortion from the FFI astrometric solution. 2.2. Astrometry Standard Catalogue We use the UCAC4 catalogue (Zacharias et al. 2013) as our astrometry standard for deriving the astromet- ric solution. It contains over 113 million objects, and is complete down to magnitude R = 16. The precision of coordinates provided by UCAC4 is ∼ 15 − 100 mas. UCAC4 catalogue also contains proper motion of ∼ 105 million stars, with errors around 1 to 10 mas/yr. Both the coordinates and proper motions are measured on the International Celestial Reference Sys- tem (ICRS) at a mean epoch of 2000. We linearly cor- rected the coordinates based on the proper motions to epoch 2014. UCAC4 also contains Two Micron All-Sky Survey ((2MASS; Skrutskie et al. 2006)) photometry for around 110 million stars, and AAVSO Photometric All- Sky Survey (APASS) five-band (BVgri) photometry for over 51 million stars. For the stars in 2MASS but with- out APASS photometry, their gri band photometry are estimated using the 2MASS magnitudes. BV band mag- nitudes are adopted from the Tycho-2 catalogue where available, otherwise also estimated from 2MASS. We use the B and V magnitude to estimate the magnitude of stars in the Kepler band when needed. 2.3. Astrometry on the Full Frame Image The Kepler FFIs are divided into subimages by read- out channels. There are 84 subimages for each FFI. Two of the CCD modules (8 channels altogether) failed dur- ing the Kepler main mission. The remaining 76 subim- ages were used to create the images from the 38 working CCDs (following the Kepler Instrument Handbook). We use fistar (P´al 2012) for source extraction. The un- certainties of the source extractor is about 0.07 pixels. This is estimated by comparing the extracted source po- sitions on the two different FFIs taken from the Cam- paign 1. The relative shift and rotation between the two FFIs were taken into account by fitting a low order polynomial to the two extracted source lists. We also experimented with other source extractors such as sex- tractor (Bertin & Arnouts 1996), and all gave similar uncertainties. The astrometric solution is provided by anmatch, a software routinely produces arcsecond precision astro- metric solutions for the HATNet/HATSouth observa- tions. anmatch first uses the engine of astrometry.net (Lang et al. 2010) for a low order solution (3rd order Simple Imaging Polynomial [SIP] tweak) to obtain an initial guess, then fits a 3rd order polynomial on a big- ger matched list between the extracted source and the catalogue source to obtain the final solution. The histogram of the residuals from the astrometric solution, for all 38 CCDs in the K2 Campaign 1 FFIs, is plotted in black in Figure 1. The astrometric so- lution residual is defined as the distance between the projected pixel coordinates of catalogue sources and the corresponding coordinates for the same stars from our source extractor. The median of the astrometric solu- tion residual for these raw frames is ∼ 0.032 pixel for K2 Campaign 1. 5 Given the Kepler CCD has a plate scale 5 We expect a factor of two difference between the estimated 3 what we achieved on the FFIs. 2.5. A revised K2 target catalogue We projected the UCAC4 catalogue on the K2 Cam- paign 1 SFFIs using the astrometric solution we obtained in §2.4. Stars with centroids within 3 pixels from the stamp edges were excluded. We also included those stars in the original Ecliptic Plane Input Catalogue (EPIC, Huber & Bryson (2015)) but not in the UCAC4 cata- logue. The EPIC catalogue is a combination of the Hip- parcos catalogue (van Leeuwen 2007), Tycho-2 catalogue (Høg et al. 2000), UCAC4 catalogue, 2MASS, and SDSS DR9 (Ahn et al. 2012) for the selected K2 target stars. There may be systematic offsets between the coordinates from the above catalogue, but they are relatively small (∼ 10 mas), and can be ignored when combining these catalogues. We estimated the B and V magnitudes of stars not in the EPIC catalogue as per the Kepler In- strument Handbook. Altogether, we found 14778 stars from the UCAC4 catalogue, and an additional 7939 stars from the EPIC catalogue only, in K2 Campaign 1. This combined set of K2 target catalogue (22717 stars in total) is larger by 5% than the total in the original EPIC cat- alogue (21647 stars in total). This increase will be more pronounced for other, more crowded K2 fields. Part of the final K2 Campaign 1 target list catalogue is shown in Table 1. We provide the centroid positions on the corresponding postage stamp, for each target. 2.6. The refined motion for each module VA14 pointed out that the K2 photometry is strongly correlated with the centroid positions of the stars. They also found that the centroid position of individual stars, as determined by their weighted light centres, are often not good enough. As such, they chose the centroid mo- tion of a bright star to represent all the stars observed in the same campaign. Taking advantage of our derived as- trometric solution, we find that by combining many stars observed on the same module, we can achieve even better constrained X,Y motion tracks. We define the X, Y mo- tion derived for the centre pixel position of SFFI modules as the refined motion for each module. As an example, we show in Figure 3 the relative X centroid drift of module 4, a module in the corner of the focal plane. We did not derive a rolling motion for the entire spacecraft to avoid correcting for additional rotations between modules. We notice that although the spacecraft attempted to correct its roll drift every 12 hours, the drifting segments can last longer. The drifting segments are defined as a time series of smooth X, Y motion without significant out- liers. We identify each drifting segment, and the outliers in between segments, by applying a 1-d edge detection method (Sobel operator) on the X motion of each mod- ule. A example of the edge detection is show for module 4, in Figure 3, with the red dashed lines separating each segment. 3. PHOTOMETRY For each target, we use fiphot (P´al 2012) to extract photometry in 36 circular apertures around the derived centroids. The flux from the sources are estimated by summing up all the pixels within an aperture and weight- ing edge pixels by the fraction of which lie within the Fig. 1. -- The residual of the astrometric solution for all chan- nels of the K2 Campaign 1 field FFIs (black) and SFFIs (red and hatched). This is computed by comparing the distance between the projected catalogue coordinates and the detected source coor- dinates on the CCDs. of 3.98′′, our astrometric solution residual corresponds to 0.127′′. 2.4. Astrometry for All the Stamps The SFFIs are very sparse. For a single channel, typ- ically more than 95% of the pixels are not downloaded. It is impossible to solve for the astrometric solution of these SFFIs via a direct catalogue matching. We use the FFI astrometric solution as an initial guess to overcome this problem. We first use fistar to extract the sources from the SFFIs. We then project the UCAC4 catalogue on to the SFFIs with the astrometric solutions obtained in §2.3. We use an iterative point matching algorithm allowing a field centre shift from the FFI to SFFI to match the extracted sources and the projected coordinates of cat- alogue stars. We solve for the distortion between these matched pairs using a second order polynomial to obtain the final solution. In Figure 2, we show the correspond- ing region of FFI and SFFI from the same CCD channel (K2 Campaign 1, module 13, channel 41). This region consists of three stamps in the SFFI observations. We marked out the detected source by cyan circles, and the projected sources from catalogue by red circles. The orig- inal K2 targets are marked out in the black circles. Some stamps consist of multiple stars. We also show that in the top rightmost stamp in Figure 2, the projected cata- logue indicates that there are additional sources blended in the primary source's PSF, which was originally missed by the source extractor, and the light of which would be measured together with that of the primary source. We compute our astrometric residuals as per §2.3. The astrometric residuals on the SFFIs for K2 Campaign 1 are shown in red in Figure 1. The median astrometric residual is around 0.034 pixels (0.135′′), comparable with uncertainties from the source extractor, and our astrometric resid- ual, given the different methods by which these two uncertainties are calculated. We estimated the uncertainty of the source extrac- tor by compare the .rms. difference of the source positions on two frames (allowing a spacial transformation), while the uncertainty of the astrometric residual is estimated by the median of residual between the extracted position and the projected solution. 4 Fig. 2. -- An 85 × 56 pixel2 region of the FFI frame (left) and its corresponding SFFI frame (right). The detected sources are marked by cyan circles, and the projected sources from UCAC4 catalogue by red circles. The original K2 targets are marked by black big circles centred on centroids as determined from our astrometric solution. The white region in the SFFI image were not observed, and are masked out. 1 ) l e x i p ( t f i h s 0.5 0 d i o r t n e c x -0.5 -1 1000 1100 1200 cadence 1300 1400 1500 Fig. 3. -- A segment of relative X centroid drift from module 4, between cadence 1000 and 1500. The dashed red lines separate the drifting segments we identified via edge detection. circular aperture. The background flux is measured by taking the median, with iterative outlier rejection, in an annulus of pixels around the aperture, then multiplying it by the area of the aperture, and subtracting it from the flux. The aperture sizes range from 2.5 -- 5 pixels, and are chosen so as to optimize the photometric precision for a wide range of magnitudes. For apertures with sizes smaller (larger) than 4.5 pixels, the background annulus has inner radii of 5 (6) pixels and outer radii of 11 (12) pixels. They are designed to optimize the photometric precision for stars in different magnitude bins. We note that for saturated stars (KepMag < 10), our photomet- ric method cannot capture all the leaked electrons in the stamps, therefore leading to degraded photometry. For the saturated stars, the fixed aperture approach taken by VA14 remains the best way to extract optimized pho- tometry for now. 4. LIGHT CURVES AND DETRENDING We present the raw aperture photometry light curves (described in section §3), and apply a three-step detrend- ing process on our light curves. The detrending method- ology is adapted from the HATNet pipeline, as well as the Kepler light curve detrending pipeline described in Huang et al. (2013). Each step of detrending is aimed to correct different aspects of the noise in the light curves. Users are able to query light curves detrended up to an intermediate step to suit their own purpose 6. In this section, we will first describe the properties of our de- trending methods, and then demonstrate the light curve products from each detrending step. At the end of this section, we will compare our light curves with those from other works. 4.1. Detrending We refer to our light curves from the aperture pho- tometry as RAWLC. Our detrending pipeline applied to these RAWLC can be divided into the following three steps: (1) External Parameter Decorrelation (EPD); (2) Trend filtering (TFA); (3) Cosine filtering (COS). To correct for the photometric variations due to the motion of the spacecraft, we performed EPD on the RAWLC (Bakos et al. 2010). We follow a similar methodology as described in VA14 to deal with the thrust fire events of the spacecraft. Instead of correcting for the drift effect due to a 6 hour roll, we make use of the drift- ing pattern of each module we identified from §2.6. We first reject the data points that fall in between any drift- ing segments. We then divide the data into 6 segments before detrending, as defined in Table 2. These segments are designed to separate large amplitude flux offset in the data, and allowing each segments to be represented by low order smooth functions. For each segment, we itera- tively fit a 3rd order B-spline through the median mag- nitude of each drifting segment with 3-σ outlier rejection until the fit converges. This long term trend represented by the B-spline is then removed. We then fit for the variation due to spacecraft drift as per the following: f (m) = c0 + c1 sin(2π X) + c2 cos(2π X) + c3 sin(2π Y ) + c4 cos(2π Y ) + c5 sin(4π X) + c6 cos(4π X) + c7 sin(4π Y ) + c8 cos(4π Y ), 6 http://k2.hatsurveys.org/ in which, X, Y represent the relative X, Y drift of the module on which the target sits. The fitted X, Y trend is then removed from the original RAWLC, and the B- spline long term trend added back in. This preserves the long-term trend while minimizing the effects of short- term spacecraft motion. The light curves at this stage is called EPDLC. The shared systematics between the stars are then cor- rected using an adaptation of the Trend Filtering Algo- rithm (TFA) designed for Kepler. The idea of TFA is to select a set of template light curves, that is representa- tive of all the systematic variations present in the data. Each target light curve is then corrected based on a lin- ear filter that identifies the shared trends between the target and the template light curves. We found that us- ing only template stars observed in the same channel as the target provided the best results. Since the number of stars observed in each channel in K2 Campaign 1 is quite small, we use all the stars but the target as templates in the TFA procedure. The TFA filtered light curves are denoted as TFALC. The last step is to filter all the low frequency vari- abilities (mostly due to intrinsic stellar variability) us- ing a set of cosine and sine functions. This method was implemented by Huang et al. (2013) for the inde- pendent search of planetary candidates in the original Kepler data. We aim to keep all periodicities at or below the protected timescale of the transit undisturbed, while minimizing any other variations following Kipping et al. (2013). The cosine function detrending is applied to the EPDLC light curves, the resulting light curves are called the COSLC. The cosine function detrending process is independent of the TFA process above. Due to its pur- pose, astrophysical signals such as stellar pulsations are no longer preserved in the COSLC. 4.2. Light Curve Products We provide two types of measurements about the pre- cision of our light curves. We use the point to point median scatter around the median (MAD) to represent the overall variability in the light curves which is used as an estimation of noise level in our transit search al- gorithm. We also report the 6.5 hour precision as per VA14, which characterizes the noise of light curves at a time scale relevant to the transit duration of an earth analog. Figure 4 shows the MAD of our light curves. The best precision of RAWLC, EPDLC, COSLC and TFALC for the bright stars are 1.2 × 10−4 (120 ppm), ∼ 6×10−5 (60 ppm), ∼ 5×10−5 (50 ppm) and ∼ 5×10−5 (50 ppm), respectively. Figure 5 shows the estimated 6.5 hour precision of our light curves. The best precision of RAWLC for the bright stars are ∼ 2 × 10−5 (20 ppm), and ∼ 1.5 × 10−5 (15 ppm) for all the other three types of light curves. We also overlaid the estimation of the bottom envelop of the original Kepler 6.5 Hour precision based on Jenkins et al. (2010). We show that the EPD process always improves both the short time scale (6.5 hours) and the long time scale (whole campaign) precision compared to the RAWLC. We find a greater improvement for the bright stars com- pared to the faint stars. The COS filtering process im- proves the precision in both time scales compared to EPDLC. For most stars, the TFALC has a similar or 5 marginally worse precision compared with the COSLC, but the TFA process tends to preserve the intrinsic vari- abilities of the stars. We note that for a small fraction of the stars, the COSLC and TFALC can have a worse 6.5 hour precision compared to the EPDLC. This is because the cosine filter and TFA algorithm both use linear least square method aiming to minimize the overall point-to- point scattering in the light curves, which is sometime achieved at the cost of increasing noise at specific time scale. We compare the noise properties of the light curves in Figure 6 by showing the ratio of per point root mean square (RMS) and MAD versus magnitude. If the noise is composited with pure white noise, this ratio should be √2. The TFALC have the most white noise compare to other detrending stages. 5. COMPARISON WITH OTHER WORKS Many teams have developed methods to improve K2 photometry. We summarize the different approaches ac- cording to their photometry and detrending methods. 5.1. Photometry Methods of previous works All the teams use aperture photometry method to ex- tract the light curves from K2 data. However, they differ in the details of aperture choice and centroid measure- ments: • Fixed Mask method: The fixed mask method is such that the flux of the target is summed up over pixels within a fixed predetermined mask, while the pixels are ac- counted in a binary way. VA14 used a combina- tion of approximate circular aperture and fitted apertures using the Kepler Pixel Responding Func- tion (Bryson et al. 2010). Foreman-Mackey et al. (2015) and Angus et al. (2015) used approximate circular but binary apertures (do not include pix- els partially) and present the photometry from the best apertures. Lund et al. (2015) used the density-based spatial clustering of applications with noise routine (DBSCAN, Ester et al. (1996)) to chose their pixel mask (aperture). • Moving Circular Aperture method: Aigrain et al. (2015) used 6 circular apertures to extract photometry. The apertures are soft-edged, in the sense that pixels straddling the edge of the aperture contribute partially to the flux. • Centroids from astrometric solution: Aigrain et al. (2015) used centroids derived from their own astrometric solution with the 2MASS all- sky point-source catalogue. • Centroids from WCS header: Foreman-Mackey et al. (2015) and Angus et al. (2015) used centroids from the WCS header of the K2 target pixel files. Only one WCS solution is given for the entire time series of each star. • Centroids from weighted centre of flux: VA14 and Lund et al. (2015) used the weighted centre of flux as the centroids of the stars. We 6 Fig. 4. -- The point to point median standard deviation around the median (MAD) versus Kepler magnitude of all the light curve products at different detrending stages. From left to right, we show the RAWLC the EPDLC, the COSLC the TFALC and the VALC. The dashed red line is the fitted function for the magnitude versus the median MAD in the magnitude bin. The solid horizontal line indicates a scatter of 10−4 (100 ppm). The vertical scale is logarithmic, and is the same for each panel. note, in the subsequent detrending, VA14 used the centroid of star EPIC 201611708 instead of the cen- troids of individual stars. In this work we used 36 moving circular apertures, with the centroids of apertures determined by precise astro- metric solutions, to determine the photometry of each star. 5.2. Detrending Methods of previous work There are three different types of "detrending" meth- ods used by other authors. • Decorrelation: VA14 , Armstrong et al. (2014) and Lund et al. (2015) used a self-flat-field method to decorrelate the aperture photometry from centroid position of the image. There are, however, subtle differ- ences between these studies. VA14 used the cen- troids from a representative star, Armstrong et al. (2014) seemed to use centroids for individual stars, while Lund et al. (2015) used the weighted light centroids derived for individual stars. VA14 used a 1-d decorrelation along the trajectory of the drift, Armstrong et al. (2014) used 2-d centroid surface to decorrelate with the flux, and Lund et al. (2015) used both the 1-d and 2-d approach in their pipeline. • Gaussian Process: Aigrain et al. (2015) and Crossfield et al. (2015) used Gaussian process model, with the rolling an- gle as the input variable to detrend the light curves. They assume the systematics can be modelled as a function form of the rolling angle, and that func- tion's form can vary from star to star. • Not Detrending: Foreman-Mackey et al. (2015) and Angus et al. (2015) choose to not detrend their light curve prior to the search of signal, but instead, they simul- taneously fit for the systematics and the signal of interest. In this work, we applied three stages of detrending. In the first stage, we applied a similar method as the decor- relation detrend in VA14. We additionally applied TFA and COS filtering to further filter the data. TFA is aimed to correct for shared systematics between the stars ob- served on the same channel, while preserving the stellar variability. The COS filtering method aimed to correct for any variability in the light curves and is optimized for searching for transit signals. 5.3. Centroids Determination We took a similar approach as Aigrain et al. (2015) in the determination of centroids by deriving an astrometric solution for each image. We made use of a more precise catalogue, UCAC4 instead of 2MASS, and the Full Frame Image as a better initial guess, and achieved higher pre- cision in our astrometric solution. Aigrain et al. (2015) 7 Fig. 5. -- The 6.5 hour precision versus Kepler magnitude of all the light curve products at different detrending stages. From left to right, we show the RAWLC the EPDLC, the COSLC the TFALC and the VALC. The solid black horizontal line indicates a scatter of 10−4 (100 ppm). The vertical scale are the same for each panel, linear and in units of parts per million (ppm). The dashed red line is the fitted function for the magnitude versus the median 6.5 hour precision in the magnitude bin. The dashed blue line is the fitted function indicate the bottom envelope of the original Kepler 6.5 Hour precision based on Jenkins et al. (2010). Fig. 6. -- Ratio of point to point RMS and MAD versus Kepler magnitude. From left to right, we show the RAWLC the EPDLC, the COSLC and the TFALC. The red horizontal line indicates the value of √2, which should be the value or their ratio for pure white noise. reported a typical root mean square of the astrometric solution of 0.4′′, or approximately 0.1 pixel, ∼ 3 times larger than our typical astrometric residuals. 5.4. Photometric Precision To date, only VA14 have released their detrended K2 Campaign 1 light curves, therefore we will focus on com- paring our photometry precision with their work. We show in Figure 7 the precision ratio between our light curves and the VA14 light curves, for the same stars at both times scales. The EPDLC, COSLC and TFALC from this work have comparable precision com- pare to VA14 light curves on the 6.5 hour time scale, and a smaller point-to-point scatter over the entire observa- tion length. Aigrain et al. (2015) presented their σMAD (similar to MAD) and 6.5 hour Combined Differential Photometric Precision (CDPP) for K2 engineering data photometry. Their best precision for σMAD is ∼ 300 ppm for the bright stars, and 60 ppm for 6.5 hour CDPP. In this work, we 8 Fig. 7. -- Photometry precision of our light curves compared to VA14 K2 Campaign 1 light curves. Top panel: 6.5 hour precision ratio between our light curves and VA14 light curves versus Kepler magnitude. Bottom panel: Per point MAD ratio between our light curves and VA14 light curves versus Kepler magnitude. From left to right, we show the RAWLC the EPDLC, the COSLC and the TFALC. The red horizontal line indicates the value of 1. We note the vertical scale in the two panels are different, and both in log scale. achieved higher precision on both time scales (50 ppm and 15 ppm, respectively). Although, we caution that the noise characteristic for K2 engineering data and K2 Campaign 1 data could be different. 5.5. Power Spectrum Previous works (Lund et al. 2015; Angus et al. 2015) noted the detrended light curves from a 1D decorrela- tion may still have residual spikes around the harmonics of ∼ 47.2271 µHz in their power spectra. These residu- als may be largely due to aliasing of the low frequency power, induced by data gaps from rejected points during thruster firing. These harmonics are less obvious in our light curves. We computed the Discrete Fourier Transfor- mation (DFT) power spectrum using vartools (Hartman et al. 2008) for the star EPIC 201183188, following Angus et al. (2015). We find our EPDLC does not show the frequencies corresponding to the 6 hour roll in the DFT power spectrum present in the VA14 light curves. Figure 8 compares our EPDLC for EPIC 201183188 with the VA14 reduction, and Figure 9 compares the DFT power spectrum of our EPDLC with that of the VA14 light curve. The offset around epoch BJD 2456015 in the VALC could be blamed for contributing to noise peaks corresponding to the space craft rolling frequency. We further investigate the noise characteristic of the light curves by computing the median of DFT power spectra of 1661 stars in the magnitude range of 10 -- 12. To eliminate the influence from the long term trend, we first filter out the strongest low frequency peak in the light curves, and then recompute the DFT power spec- tra of each star before taking the median. We show this median spectrum in Figure 10. The peaks related to the rolling frequency are still present, but have less power compared to the median DFT spectrum computed with the same set of stars from VA14. When the signal from stellar variability is strong, such as in the case of EPIC 201183188, the systematic noise peaks are negligible. 5.6. Example Light Curves To demonstrate the products of our photometric ex- traction and detrending in the context of stellar variabil- ity and transit searches, we show example light curves for stars with known variability and transiting signals. EPIC 201711881, is a Cepheid candidate discovered by the ASAS project (Schmidt et al. 2009). The original discovery paper reported a period of 2.7353±5 × 10−4 days. The period we detected is ∼ 2.735 day, consistent with (Schmidt et al. 2009). The RAWLC and EPDLC folded with the Cepheid period, are shown in Figure 11. Our RAWLC without any detrending, already shows a clean periodic signal. In addition, our EPDLC preserves the amplitude of the pulsation after removing the sys- tematics. We also find a periodic eclipsing signal in the light curve with twice the pulsation period. This eclips- ing signal can be visually seen, even in the RAWLC. 10.644 EPIC 201183188 this work,EPD Mag VA14 e d u t i n g a M e v i t a l e R 10.646 10.648 10.65 10.652 10.654 10.656 1990 2000 2010 2020 2030 2040 2050 2060 BJD-2454000 Fig. 8. -- The EPDLC for EPIC 201183188 from this work (black), and VA14 (red). The DFT power spectrum of this star is shown in Figure 9. x 10-8 EPIC 201183188 VA14 0 x 10-9 50 100 150 200 250 this work,EPDMag 2 1 0 3 2 1 r e w o p T F D r e w o p T F D 0 0 50 100 150 frequency (µHz) 200 250 Fig. 9. -- The DFT power spectrum for EPIC 201183188 from this work (top panel) and VA14 (bottom panel). Note the vertical scales are different in the two panels. The dash lines indicate the corresponding frequency associated with the spacecraft roll motion. For more discussion, see Angus et al. (2015). VA14 This work x 10-9 r e w o p T F D 4 3 2 1 40 42 44 10-5 10-6 10-7 10-8 10-9 10-10 10-11 r e w o p T F D 48 46 frequency (µHz) 50 52 54 0 50 100 frequency (µHz) 150 200 250 Fig. 10. -- The median DFT power spectrum of 1661 stars in the magnitude range 10-12. Red dashed lines show the result from the VA14 light curves. Black solid lines show the result from the EPDLC of this work. In the small window of this figure, we show the zoom-in of these spectra around the space craft roll frequency ∼ 47.2271µ Hz. 9 The COSLC, phase folded with the eclipse period, is also plotted in the bottom panel of Figure 11. It seems that this system is more likely to be an eclipsing binary system with large out-of-transit variation rather than a real Cepheid. To further characterize this signal, the COSLC from the general detrending pipeline is not good enough. The COSLC we show in Figure 11 have been reconstructed after the discovery of the eclipsing signal. Only the out-of-eclipse part have been filtered by the co- sine filters in order to preserve the shape and amplitude of the eclipsing signal. WASP-85b: WASP-85 (EPIC 201852715), was ob- served in K2 Campaign 1, in module 15, channel 49. 1182 other stars were observed in the same module. We show in Figure 12 the light curves of WASP-85 at dif- ferent detrending stages from this work. In the second panel from the top, we also overlaid the VA14 detrended light curve. WASP-85b has a period of ∼2.65 days, and known depth of ∼ 1.6%. We show the WASP-85 light curve folded with the detected period and epoch in phase space for both the COSLC and TFALC in the bottom of Figure 12. K2-3: K2-3 (EPIC 201367065), was observed in K2 Campaign 1, in module 12, channel 40. The host star is an M dwarf, with three transiting super-earths discov- ered by Crossfield et al. (2015). We show the light curves for K2-3 in Figure 13. The transits of the biggest planet (1 mmag) is visible in our RAWLC, and the transits of all three planets are visible in all the other light curves. We also show the phase folded COSLC and TFALC for all three planets in the bottom panel of Figure 13. EPIC 201613023: This star was identified as a tran- siting planet candidate system by Foreman-Mackey et al. (2015). We show our light curves in Figure 14. The transit signal has depth of 400 ppm. Individual transits from the planets are visible in the EPDLC, TFALC and COSLC. The phase folded COSLC and TFALC with the detected epoch and period are shown in the bottom panel. 6. CONCLUSION AND DISCUSSION In this article, we present our effort to extract high precision photometry from K2 Campaign 1 data. Our method has three distinct advantages: • Making use of accurate astrometric solution (0.127′′ or 0.034 pixels) from the FFIs for aperture centroiding; • Providing photometry for all sources on the stamps, not only for the proposed targets from the input catalogue; • Presenting light curves with very low systematic variations. Our extracted light curves are of high precision at both the long (entire campaign) and short (6.5 hours) time scales, even for the raw light curves without any detrend- ing. Light curves derived from all 36 photometric aper- tures at all four detrending stages are provided for the public at http://k2.hatsurveys.org. ACKNOWLEDGMENTS 10 EPIC 201711881 g a M ∆ g a M ∆ g a M ∆ g a M ∆ g a M ∆ g a M ∆ Raw LC EPD LC VA LC TFA LC COS LC 10.1 10.15 10.2 10.25 10.1 10.15 10.2 10.25 10.1 10.15 10.2 10.25 10.12 10.14 10.16 10.18 (A) (B) (C) (D) 1970 1980 1990 2000 2010 2020 2030 2040 2050 2060 BJD-2454833 (Day) EPD LC VALC (E) 1 g a M ∆ 10.05 10.1 10.15 10.2 10.25 10.3 0 (F) 1 0 Phase (G) Raw LC 0 Phase COS LC 0 10.05 10.1 10.15 10.2 10.25 10.3 10.12 10.14 10.16 -0.5 0 Eclipse Phase 0.5 Fig. 11. -- The light curves of EPIC 201711881, a Cepheid Candidate discovered by Schmidt et al. (2009). From the top to bottom, we show the RAWLC (A), EPDLC (B), TFALC (C) and COSLC (D). In panel (E) and (F), we show RAWLC and EPDLC folded with the period of the pulsation period (∼2.735 day) in the phase space. The last panel we show COSLC (G) folded with twice the period of the pulsation. The eclipse events can be seen at phase 0. The COSLC in panel (D) and (E) have been reconstructed after the discovery of the eclipsing signal. Only the out-of-eclipse part have been filtered by the cosine filters in order to preserve the signal. The red light curve in figure (B) and (F) is from VA14. We note that the vertical scales are different for each panel. EPIC 201862715, WASP-85 Raw LC EPD LC VA LC TFA LC COS LC 11 (A) (B) (C) (D) g a M ∆ g a M ∆ g a M ∆ g a M ∆ 10.21 10.22 10.23 10.24 10.25 10.19 10.21 10.23 10.25 10.27 10.29 10.21 10.22 10.23 10.24 10.25 10.21 10.22 10.23 10.24 10.25 10.26 1970 1980 1990 2000 2010 2020 2030 2040 2050 2060 BJD-2454833 (Day) g a M ∆ 10.22 10.23 10.24 10.25 10.26 COS LC (E) g a M ∆ -2 -1.5 -1 -0.5 0 0.5 1 1.5 2 10.21 10.22 10.23 10.24 10.25 10.26 TFA LC (F) -2 -1.5 -1 -0.5 0 0.5 1 1.5 2 Hours from Mid-Transit Hours from Mid-Transit Fig. 12. -- The light curves of WASP85 (black). From the top to bottom, we show the RAWLC, EPDLC, TFALC and COSLC. The bottom most panel, we show COSLC and TFALC of WASP85 folded with the epoch and period of WASP-85b (Brown et al. 2014). The red light curve in the second panel is from VA14. We note that the vertical scales in different panels are different. 12 g a M ∆ g a M ∆ g a M ∆ g a M ∆ g a M ∆ Raw LC EPD LC VA LC TFA LC COS LC 11.578 11.58 11.582 11.584 11.578 11.58 11.582 11.584 11.578 11.58 11.582 11.584 11.58 11.582 EPIC 201367065 (A) (B) (C) (D) 11.584 1970 11.58 11.582 11.584 11.586 -4 1980 1990 2000 2010 2020 2030 2040 2050 2060 BJD-2454833 (Day) COS LC (E) TFA LC (F) K2-3 b K2-3 c K2-3 d g a M ∆ 11.58 11.582 11.584 K2-3 b K2-3 c K2-3 d -1 -2 -3 3 Hours from Mid-Transit 0 2 1 4 11.586 -4 -1 -2 -3 3 Hours from Mid-Transit 0 2 1 4 Fig. 13. -- Same as Figure 12, but for K2-3 (Crossfield et al. 2015). In the bottom panel, we show all the three planets detected in this system folded with their own periods and epochs. K2-3 b (P∼ 10 day, R∼2R⊕ ), c (P∼ 25 day, R∼1.7R⊕ ), d (P∼ 44 day, R∼1.5 R⊕ ) are presented in red,blue and black curves, respectively. We note that the vertical scales in different panels are different. 13 (A) (B) (C) (D) 12.05 1970 1980 1990 2000 2010 2020 2030 2040 2050 2060 BJD-2454833 (Day) 12.048 12.049 12.05 -6 COS LC (E) 12.048 12.049 g a M ∆ TFA LC -2 4 -4 Hours from Mid-Transit 2 0 6 12.05 -6 -2 4 -4 Hours from Mid-Transit 2 0 (F) 6 Fig. 14. -- Same as Figure 12, but for EPIC 201613023 (Foreman-Mackey et al. 2015). In the bottom panel we show the detected planet candidate folded with its period (∼8.28 days). The depth of the transit is ∼ 400 ppm. We note that the vertical scales in different panels are different. EPIC 201613023 12.045 Raw LC g a M ∆ 12.047 12.049 12.051 EPD LC VA LC TFA LC COS LC 12.045 12.047 12.049 12.051 12.048 12.05 12.048 12.049 g a M ∆ g a M ∆ g a M ∆ g a M ∆ 14 We would like to thank the referee for their helpful comments. We also thank A.V for his thoughtful sug- gestions. G..B. and X.H acknowledge funding from the Packard Foundation. This work was also supported by NASA grant NNX13AJ15G. K.P. acknowledges support from NASA grant NNX13AQ62G. The K2 data pre- sented in this paper were obtained from the Mikulski Archive for Space Telescopes (MAST). STScI is operated by the Association of Universities for Research in Astron- omy, Inc., under NASA contract NAS5-26555. Support for MAST for non-HST data is provided by the NASA Office of Space Science via grant NNX09AF08G and by other grants and contracts. This paper includes data collected by the Kepler telescope. Funding for the K2 Mission is provided by the NASA Science Mission direc- torate. REFERENCES Ahn, C. P., Alexandroff, R., Allende Prieto, C., et al. 2012, ApJS, 203, 21 Huber, D., & Bryson, S. T. 2015, KSCI-19082-008 Jenkins, J. M., Caldwell, D. A., Chandrasekaran, H., et al. 2010, Aigrain, S., Hodgkin, S. T., Irwin, M. J., Lewis, J. R., & Roberts, ApJL, 713, L120 S. J. 2015, MNRAS, 447, 2880 Kipping, D. M., Hartman, J., Buchhave, L. A., et al. 2013, ApJ, Angus, R., Foreman-Mackey, D., & Johnson, J. A. 2015, ArXiv 770, 101 e-prints, 1505.07105 Lang, D., Hogg, D. W., Mierle, K., Blanton, M., & Roweis, S. Armstrong, D. J., Osborn, H. P., Brown, D. J. A., et al. 2014, 2010, AJ, 137, 1782, arXiv:0910.2233 ArXiv e-prints, 1411.6830 Bakos, G. ´A., Torres, G., P´al, A., et al. 2010, ApJ, 710, 1724 Bertin, E., & Arnouts, S. 1996, A&AS, 117, 393 Brown, D. J. A., Anderson, D. R., Armstrong, D. J., et al. 2014, ArXiv e-prints, 1412.7761 Bryson, S. T., Tenenbaum, P., Jenkins, J. M., et al. 2010, ApJL, 713, L97 Crossfield, I. J. M., Petigura, E., Schlieder, J. E., et al. 2015, ApJ, 804, 10 Ester, M., Kriegel, H.-p., Sander, J., & Xu, X. 1996, AAAI Press, 226 Foreman-Mackey, D., Montet, B. T., Hogg, D. W., et al. 2015, ApJ, 806, 215 Hartman, J. D., Gaudi, B. S., Holman, M. J., et al. 2008, ApJ, 675, 1254 Høg, E., Fabricius, C., Makarov, V. V., et al. 2000, A&A, 355, L27 Howell, S. B., Sobeck, C., Haas, M., et al. 2014, PASP, 126, 398 Huang, X., Bakos, G. ´A., & Hartman, J. D. 2013, MNRAS, 429, 2001 Lund, M. N., Handberg, R., Davies, G. R., Chaplin, W. J., & Jones, C. D. 2015, ApJ, 806, 30 Monet, D. G., Jenkins, J. M., Dunham, E. W., et al. 2010, ArXiv e-prints, 1001.0305 P´al, A. 2012, MNRAS, 421, 1825 Schmidt, E. G., Hemen, B., Rogalla, D., & Thacker-Lynn, L. 2009, AJ, 137, 4598 Skrutskie, M. F., Cutri, R. M., Stiening, R., et al. 2006, AJ, 131, 1163 van Leeuwen, F., ed. 2007, Astrophysics and Space Science Library, Vol. 350, Hipparcos, the New Reduction of the Raw Data Vanderburg, A., & Johnson, J. A. 2014, PASP, 126, 948 Zacharias, N., Finch, C. T., Girard, T. M., et al. 2013, AJ, 145, 44 UCAC4ID RA DEC J H K B V g r i x(t0) b y(t0) b channel c K2ID d flaga TABLE 1 K2 target list UCAC4-555-033290 UCAC4-555-033327 UCAC4-555-033328 UCAC4-555-033330 UCAC4-555-033335 UCAC4-555-033336 UCAC4-555-033338 UCAC4-555-033418 UCAC4-555-033420 UCAC4-555-033425 ... 101.764743 101.789413 101.789718 101.790216 101.795353 101.795597 101.799223 101.876901 101.876908 101.881285 20.947725 20.952493 20.941804 20.942716 20.945481 20.950077 20.951700 20.980552 20.972664 20.975126 11.705 14.487 15.207 16.218 11.554 16.008 15.611 14.091 12.859 16.224 11.192 12.700 14.082 14.420 11.266 14.560 14.331 13.383 11.979 13.317 11.067 12.157 13.808 13.828 11.224 14.051 14.073 13.252 11.702 12.709 11.927 15.417 15.967 17.923 11.589 17.444 15.935 14.427 13.485 17.317 11.661 14.481 15.332 16.773 11.466 16.468 15.416 14.080 12.948 16.082 11.756 14.920 15.603 17.021 11.478 16.771 15.625 14.193 13.181 16.193 11.696 14.144 15.144 16.130 11.559 16.030 15.249 14.071 12.839 15.201 11.726 13.857 14.984 15.738 11.677 15.720 15.125 14.044 12.712 14.762 41.098906 42.030470 32.436217 33.183423 34.961871 39.036805 39.999106 55.344496 48.289514 49.903488 885.251787 863.957681 865.242115 864.694662 860.006732 859.140643 855.879192 786.888096 788.011635 784.005099 24 24 24 24 24 24 24 24 24 24 a The photometry flag indicator: A-APASS photometry; T-Tycho photometry; e-estimated with 2MASS photometry b the X and Y coordinate of star at time 0, time 0 is defined as cadence 1 (BJD-2455895.528) for K2.0, and cadence 102 (BJD-2455975.178) for K2.1 c the channel number of which the star is observed d the given K2 ID of the star eeeeeeeee eeeeeeeee 202071861 AAeeeAAAe 202071849 AAeeeAAAe 202071849 202071849 202071849 AAeeeAAAe 202071849 202071849 202068459 AAeeeAAAe 202068459 AAeeeAAAe 202068459 eeeeeeeee eeeeeeeee eeeeeeeee 1 5 16 Light Curve Segments used in EPD TABLE 2 Segment No start Cadence end Cadence 1 2 3 4 5 6 0 455 1006 2050 2315 2998 454 1005 1989 2314 2997 4020
1804.03006
1
1804
2018-04-09T14:13:14
Small impacts on the giant planet Jupiter
[ "astro-ph.EP" ]
Video observations of Jupiter obtained by amateur astronomers over the past eight years have shown five flashes of light of 1-2 s. The first three of these events occurred on 3 June 2010, 20 August 2010, and 10 September 2012. Previous analyses showed that they were caused by the impact of objects of 5-20 m in diameter, depending on their density, with a released energy comparable to superbolides on Earth of the class of the Chelyabinsk airburst. The most recent two flashes on Jupiter were detected on 17 March 2016 and 26 May 2017 and are analyzed here. We characterize the energy involved together with the masses and sizes of the objects that produced these flashes. The rate of similar impacts on Jupiter provides improved constraints on the total flux of impacts on the planet, which can be compared to the amount of exogenic species detected in the upper atmosphere of Jupiter. We extracted light curves of the flashes and calculated the masses and sizes of the impacting objects. An examination of the number of amateur observations of Jupiter as a function of time allows us to interpret the statistics of these detections. The cumulative flux of small objects (5-20 m or larger) that impact Jupiter is predicted to be low (10-65 impacts per year), and only a fraction of them are potentially observable from Earth (4-25 per year in a perfect survey). More impacts will be found in the next years, with Jupiter opposition displaced toward summer in the northern hemisphere. Objects of this size contribute negligibly to the exogenous species and dust in the stratosphere of Jupiter when compared with the continuous flux from interplanetary dust punctuated by giant impacts. Flashes of a high enough could produce an observable debris field on the planet. We estimate that a continuous search for these impacts might find these events once every 0.4 to 2.6 years.
astro-ph.EP
astro-ph
Astronomy & Astrophysics manuscript no. 32689_astroph April 10, 2018 c(cid:13)ESO 2018 8 1 0 2 r p A 9 . ] P E h p - o r t s a [ 1 v 6 0 0 3 0 . 4 0 8 1 : v i X r a Small impacts on the giant planet Jupiter R. Hueso1,(cid:63), M. Delcroix2, A. Sánchez-Lavega1, S. Pedranghelu, G. Kernbauer, J. McKeon3, A. Fleckstein, A. Wesley4, J.M. Gómez-Forrellad5, J.F. Rojas1, and J. Juaristi1 1 Física Aplicada I, Escuela de Ingeniería de Bilbao, UPV/EHU, Alameda Urquijo s/n, 48013, Bilbao, Spain 2 Societé Astronomique de France, France 3 Meath Astronomy Group, Dublin, Ireland 4 Astronomical Society of Australia, 1502 Rubyvale Rd, Rubyvale Queensland 4702 Australia 5 Fundacio Observatori Esteve Duran, Spain April 10, 2018 ABSTRACT Context. Video observations of Jupiter obtained by amateur astronomers over the past eight years have shown five flashes of light with durations of 1-2 s. The first three of these events occurred on 3 June 2010, 20 August 2010, and 10 September 2012. Previous analyses of their light curves showed that they were caused by the impact of objects of 5-20 m in diameter, depending on their density, with a released energy comparable to superbolides on Earth of the class of the Chelyabinsk airburst. The most recent two flashes on Jupiter were detected on 17 March 2016 and 26 May 2017 and are analyzed here. Aims. We characterize the energy involved together with the masses and sizes of the objects that produced these flashes. The rate of similar impacts on Jupiter provides improved constraints on the total flux of impacts on the planet, which can be compared to the amount of exogenic species detected in the upper atmosphere of Jupiter. Methods. We extracted light curves of the flashes and calculated the masses and sizes of the impacting objects after calibrating each video observation. An examination of the number of amateur observations of Jupiter as a function of time over the past years allows us to interpret the statistics of these impact detections. Results. The cumulative flux of small objects (5-20 m or larger) that impact Jupiter is predicted to be low (10-65 impacts per year), and only a fraction of them are potentially observable from Earth (4-25 per year in a perfect survey). Conclusions. We predict that more impacts will be found in the next years, with Jupiter opposition displaced toward summer in the northern hemisphere where most amateur astronomers observe. Objects of this size contribute negligibly to the abundance of exogenous species and dust in the stratosphere of Jupiter when compared with the continuous flux from interplanetary dust particles punctuated by giant impacts. Flashes of a high enough brightness (comparable at their peak to a +3.3 magnitude star) could produce an observable debris field on the planet. We estimate that a continuous search for these impacts might find these events once every 0.4 to 2.6 years. Key words. Planets and satellites: Jupiter, Planets and satellites: atmospheres, Meteorites, meteors, meteoroids 1. Introduction Because of its large gravitational attraction and effective cross section, the giant planet Jupiter is the most likely place to re- ceive impacts in the solar system. Direct and dramatic evidence of impacts on Jupiter was acquired with the observations of the series of impacts from the comet Shoemaker-Levy 9 (SL9) in July 1994 (Harrington et al. 2004). On July 19, 2009, an un- known body collided with Jupiter on its night side (Sánchez- Lavega et al. 2010). In both cases, the impacts produced large spots of material that were dark in the visible wavelength range and bright in methane absorption bands because of the high alti- tude of the debris fields. These spots remained visible for weeks to years (Hammel et al. 1995, 2010; Sanchez-Lavega et al. 1998; Sánchez-Lavega et al. 2011). Impacts supply disequilibrium species to the upper atmo- sphere of Jupiter, which in the case of SL9 are still unam- (cid:63) e-mail: [email protected] biguously observable today because of the higher concentra- tion of water, CO, and other chemical species in the south- ern hemisphere of Jupiter (Lellouch et al. 2002; Cavalié et al. 2013). Spectroscopic observations of the planet allow inferring the amount of exogenic molecules in the upper atmosphere of Jupiter. However, the relative contribution to the abundance of exogenic water and carbon dioxide from impacts of very differ- ent size range from a continuous supply of interplanetary dust to rare impacts of large objects are not yet well characterized (Lellouch et al. 2002; Bézard et al. 2002; Poppe 2016; Moses & Poppe 2017). Impacts from objects of about 10 m in diameter have been detected in telescopic observations of Jupiter from the sudden release of luminous energy when the impacting objects enter the atmosphere of Jupiter and explode as atmospheric bolides (Hueso et al. 2010b, 2013). Five impacts have been detected in this way since 2010. Three of them have been examined previ- ously in the literature (Hueso et al. 2010b, 2013), and two more have occurred since then. Each of these impacts has been de- Article number, page 1 of 14 A&A proofs: manuscript no. 32689_astroph tected simultaneously by more than one observer (12 observers recorded 11 video acquisitions for a total of five impact bolides). Hueso et al. (2010b) presented the analysis of the first bolide impact on Jupiter and the general method for calibrating light curves from amateur observations of Jovian flashes. They also converted them into energies, masses, and sizes of the impact- ing object. The first Jovian bolide was caused by an object with a diameter in the range of 10 m, releasing an energy compara- ble to an object of about 30 m impacting Earth's atmosphere. Follow-up observations with telescopes such as the Very Large Telescope (VLT) or the Hubble Space Telescope (HST) showed no evidence of atmospheric debris left by the impact, confirm- ing the small size of the object. Hueso et al. (2013) extended this study to a characterization of the three known bolides in 2013. The combined analysis of these three impacts allowed a first quantification of the flux of similar impacts on Jupiter. The expected number was 12-60 impacts per year for objects larger than 5-20 m in diameter. This impact rate is close to expectations based on an extrapolation of dynamical models of comets and as- teroids in orbits prone to Jupiter encounters (Levison et al. 2000) (30-100 collisions per year for objects with diameters larger than 5-20 m). Hueso et al. (2013) also presented model simulations of airbursts caused by these small-sized objects following similar techniques to those used in simulations of larger impacts (Ko- rycansky et al. 2006; Palotai et al. 2011; Pond et al. 2012). We here update previous results presented in Hueso et al. (2010b, 2013) with the analysis of the latest two impacts de- tected in March 2016 and May 2017. We examine how the new observations constrain the flux of impacts on Jupiter similarly to the estimate in Hueso et al. (2013). We also discuss the implica- tions of the predicted flux of impacts on the amount of exogenic species (water and carbon monoxide) and dust in the upper atmo- sphere of Jupiter, and we discuss the probability of finding more intense flashes from larger objects that could leave observable traces in the atmosphere of Jupiter from follow-up observations. The outline of this paper is the following: In section 2 we summarize previous observations of the first three impact fire- balls (Hueso et al. 2010b, 2013), and we give observational de- tails of the latest two impacts found on Jupiter in March 17, 2016, and May 26, 2017. In section 3 we present light curves of these two impacts and calibrate the images to obtain size and mass estimations of these objects. In section 4 we examine the sizes and masses of impacts required to leave an observable de- bris field in the planet atmosphere. In section 5 we discuss cur- rent efforts of detecting new observation flashes on Jupiter. In section 6 we present an statistical analysis of the amateur ob- servations to infer the statistical significance of these impact de- tections in a larger context. In section 7 we present an updated estimate of the impact flux on Jupiter and discuss the implica- tions for exogenous water and carbon monoxide on the upper atmosphere of Jupiter and the probability of finding observable debris fields in the atmosphere of Jupiter after more intense su- perbolides. We present our conclusions and a summary of our findings in section 8. 2. Observations 2.1. Summary of previous impacts On June 3, 2010, Anthony Wesley (Australia) and Christopher Go (Philippines) recorded a short flash while taking video ob- servations of Jupiter. The flash lasted about two seconds and was observed by A. Wesley, who issued an e-email alert that was confirmed later by C. Go. A quick and large follow-up cam- Article number, page 2 of 14 Fig. 1. First three fireballs found on Jupiter from observations by A. Wesley (top panel), M. Tachikawa (middle panel), and G. Hall (bottom panel). Arrows show the position of the impacts. Images have been pro- cessed by stacking the frames where the flash is visible and adding the result over a Jupiter image built from the stack of the full video of the planet at the time of the impact. The color in the first panel comes from acquisitions obtained with a filter wheel in the minutes before and after the flash. The color in the second panel comes from the detector, which makes use of a Bayer mask to produce color images. The diffraction- like ring patterns in the last panel show the brightest flash. R. Hueso et al.: Small impacts on the giant planet Jupiter paign was organized, and it obtained observations of the planet within a few days with telescopes such as the VLT and HST. None of these observations found any debris field in the region hit by the impact. A later analysis of the flash light curve resulted in the conclusion that it was caused by an object of 8-13 meters in diameter impacting the atmosphere of Jupiter and producing a giant fireball (Hueso et al. 2010b). Two months later, on August 20, 2010, another flash on Jupiter was detected by amateur astronomer Masayuki Tachikawa of Japan, and it was confirmed by Kazuo Aoki and Masayuki Ishimaru. This flash was found in videos of lower quality in the wake of published news in amateur astronomy journals about the first flash. Two years later, on September 10, 2012, a new fireball on Jupiter was discovered by a visual ob- server, who issued an alert in astronomical forums (Dan Petersen from Racine in Wisconsin). The flash was later confirmed by a video observation obtained by George Hall from Dallas, Texas. This flash was significantly brighter than the previous two im- pacts. Subsequent analysis of the video observation agreed re- markably well with the brightness estimate from Dan Petersen, who had observed the flash visually in the telescope eyepiece. Figure 1 shows processed versions of the observations of the three flashes. In these three cases (as well as in the next two flashes discussed below), an observer raised the alert to the amateur community after observing the impact flash. Confirma- tions from other observers who had been taking data at the same time quickly followed, but we remark that most observers did not see the flash originally when they were at the telescope, or when they first analyzed their video observations with au- tomatic stacking software tools. Amateur astronomers combine thousands of frames from a single video into a stacked image with high signal-to-noise ratio using automatic software tools (Mousis et al. 2014), where the light of any possible short flash dilutes within the rest of the frames, rendering it invisible in the final image. In all cases, the regions that were hit did not show any trace of the bolide material in later observations either by large telescopes such as in June 2010 and September 2012 or by fast amateur follow-ups in August 2010. Therefore these events can only be discovered if it is spotted in the few seconds during which each impact produces a bright fireball. 2.2. Impact on March 17, 2016 A new impact was detected on March 17, 2016, by Gerrit Kern- bauer (Austria). The impact was announced ten days after the ob- servation because the relatively poor seeing of that night caused G.K. to delay an analysis of the video observations. The an- nouncement was noted by John McKeon (Ireland), who had been observing the planet in the same night for 3.5 hours, building a time-lapse video of Jupiter and its moons. The second video confirmed the finding with better image quality. It is difficult to visually find a short flash of light of 2 seconds in a sequence of video observation that lasts several hours. 2.3. Impact on May 26, 2017 The most recent impact on Jupiter was found by Sauveur Pe- dranghelu from Corsica (France) on May 26, 2017. The impact was announced the next day and was quickly confirmed by two German observers, Thomas Riessler and André Fleckstein, both after reading news of the impact posted on German astronomi- cal forums. The videos by S.P. and T.R. were of excellent quality, and the video observation by A.F. was not as good because of the Fig. 2. Latest two fireballs found on Jupiter. Arrows show the position of the impacts. Upper panel: Impact on March 17, 2016, observed by G. Kernbauer and J. McKeon. The image has been processed for aes- thetics by S. Voltmer using data from the two video observations. The color comes from the G.K. video and the luminance from J. McKeon. Bottom panel: Impact on May 26, 2017, observed by S. Pedranghelu, T. Riessler, and A. Fleckstein. The image is a combination of video obser- vations in color from S.P. and T.R. All image combinations were done with stacking software and wavelet processing to increase the sharp- ness of atmospheric features, and the impact was added from a separate processing of the frames where the flash is visible. atmospheric seeing at his location. It was almost impossible to observe the flash in this video without previous knowledge of the moment where the impact had occurred, meaning that good atmospheric seeing is a critical factor in the discoveries of these events. Figure 2 shows these two impacts from a variety of video observations. Table 1 summarizes the dates of all the impacts, observers, and equipment. 2.4. Follow-up observations For the 2016 and 2017 impacts, observations by a variety of amateur astronomers were obtained from between 10 minutes to a few rotations after the impact. None of these observations Article number, page 3 of 14 Table 1. Jovian bolides detections A&A proofs: manuscript no. 32689_astroph Date (yyyy-mm-dd) Observers Time (hh-mm-ss) (and locations) (UT) 2010-06-03 (20:31:20) 2010-08-20 (18:21:56) 2012-09-10 (11:35:30) 2016-03-17 (00:18:39) 2017-05-26 (19:24:50) A.Wesley (Australia) C. Go (Phillipines) M. Tachikawa (Japan) K. Aoki (Japan) M. Ishimaru (Japan) D. Petersen (USA) G. Hall (USA) G. Kernbauer (Austria) J. McKeon (Ireland) S. Pedranghelu (France) T. Riessler (Germany) A. Fleckstein (Germany) Telescope diameter (cm) 37 28 15 23.5 12.5 30.5 30.5 20 28 20.3 20.3 28 Detector Filters Point Grey Flea3 Point Grey Flea3 650 nm 435 nm Philips ToUCam II Bayer RGB Philips ToUCam II Bayer RGB Philips ToUCam II Bayer RGB Visual observation Point Grey Flea3 - 640 nm QHY5LII ASI120MM ASI224MC ASI120MC ASI120MM Bayer RGB IR742 Bayer RGB Bayer RGB IR742 Flash duration (s) 1.9 0.95 Sampling rate (fps) 60 55 1.4 1.9 1.1 -- 1.7 1.15 1.30 1. 38 0.88 0.92 30 15 30 - 15 47 26 61.79 30.78 30 Notes. * fps stands for frames per second showed any debris field on the surface of the planet at the loca- tions of each impact. We surveyed the Planetary Virtual Obser- vatory and Laboratory (PVOL) (Hueso et al. 2010a, 2018) and the Association of Lunar and Planetary Observers (ALPO) Japan databases of amateur observations as well as several amateur as- tronomical forums in search of images covering these areas. In the case of the March 2016 impact, the impact area was just dis- appearing behind the east limb, and the high-resolution observa- tion closest in time was obtained by Randy Christensen (USA) one Jovian rotation later (10 hr). For the May 2017 impact, the geometry was better suited. Images were acquired inmediately after the impact by A.F., but showed no brightening or a darkening of the impact area. Images with some better seeing were acquired 10 minutes after the im- pact by Giancarlo Rizatto (Italy) and Philipp Salzgeberg (Aus- tria) without any observable impact feature with a good image quality. A better observation was obtained by William Pelissard (France) 30 minutes after the impact and did not show an ob- servable debris field. A high-resolution observation obtained 10 hours later by Randy Christensen did not show any observable perturbation at the impact location. Finally, a methane band ob- servation by Christopher Go 40 hours after the impact did not show any bright feature in the planet. Figure 3 shows a selection of these images. 3. Analysis of the impacts in March 2016 and May 2017 3.1. Light curves For each video observation, we transformed the initial video files into a sequence of numbered frames that were analyzed with a software pipeline written in IDL. The pipeline coregisters all frames by calculating the relative motions of the frames caused by the atmospheric seeing. A reference image is calculated from a stack of coregistered images and normalized by taking into account the number of frames used. The coregistration is done by an image correlation algorithm and is loosely based on the PLAYLIST pipeline for lucky imaging of planets (Mendikoa et al. 2016). The impact location is found by calculating an im- Article number, page 4 of 14 Fig. 3. Selected follow-up observations of the events in March 2016 (top panel) and May 2017 (bottom panel). In the follow-up observation by A.F., differential images with images acquired just before the flash do not show any significant difference at the impact location. age built from the maximum brightness of each pixel and sub- stracting the average brightness for each location. This generally produces an image where the impact location is well contrasted and can be found automatically. However, atmospheric seeing not only moves the planet from one frame to another, it also dis- torts the planet shape, causing the flash light to apparently move around the main impact location. In order to calculate the light from the flash, the software calculates the difference between each frame and the reference R. Hueso et al.: Small impacts on the giant planet Jupiter Fig. 4. Light-curve analysis pipeline. (a) Section of the original frame; (b) differential image for that frame, i.e., the current frame minus the reference image; (c) circular mask around the mean location of the im- pact; and (d) circular mask within a smaller radius containing all the light from the impact and recentered on the impact. Seeing distorts each image differently, and the location of the flash needs to be adjusted frame by frame by the software. The example is from the S.P. video observation on 2017-05-26. image after coregistering each individual frame with the refer- ence image. Differential photometry images are used to calcu- late aperture photometry over the impact location. Aperture pho- tometry is done using a circular mask plus a ring to substract contributions from the background in the differential photome- try images. The circular mask is recentered in each frame above the flash location by correcting distortions caused by the see- ing. Additionally, the software also calculates the integrated flux of Jupiter so that the brigthness of the impact can be compared with the total brightness of the planet. Figure 4 shows examples of the pipeline processing. Images like these are generated by the pipeline and are used to check the correct positioning of the moving-aperture photometry mask. For video observations with cameras that use a Bayer RGB filter, we convert the color frames into black-and-white versions using the average of all three channels. This allows better visibil- ity of the impact and a more detailed and less noisy light-curve. Figures 5 and 6 show raw light curves of the impacts in March 2016 and May 2017, respectively. In this case the light curve from the analysis of S.P.'s video observation shows signif- icant temporal structure with a double central flash and an ex- tended tail of brighness decay lasting for about 0.6 s. The double central flash is also partially distinguishable in the second video of this event by T.R., but the smaller pixel size of the optical setup prevents us from extracting more accurate information of this video. The two flashes are readily apparent when examin- ing the video observation frame by frame and can be related to fragmentations of the impact object. The same types of features with similar timescales are observed on light curves of super- bolides on Earth (see, e.g., Figure 3 in Borovicka et al. 2017 and the extended figure 2 in Borovicka et al. 2013). This might therefore be the first case for fragmentation of Jupiter bolides Fig. 5. Raw light curves of the impact in March 2017. Top panel: Data from Gerrit Kernbauer. Bottom pannel: Data from John McKeon. and suggests that observations of future impacts should involve a fast frame rate of at least 30 fps to observe these characteristics. The fragmentation history of a bolide depends on the entry mass, physical nature of the meteorite, speed, and angle of the impact. A sophisticated light curve analysis can be made for Earth super- bolides resulting in the physical characterization of the impactor (Borovicka et al. 2017) (i.e., determining the physical class of the impacting object, which can be stony, metallic, icy compact, or icy porous). The fast cameras currently used by most ama- teurs might start to produce such data in observations of Jupiter impacts for objects more massive than the impact in May 2017. 3.2. Image calibration The total intensity from each flash, I∗, is calculated from the inte- grated data numbers (DNs) of each light curve. This is computed by adding the excess DNs from the minimum to the maximum times marked in figures 5 and 6. This number is compared with the total DNs associated to the full disk of Jupiter over the ref- erence image built for each video, IJ. Transforming I∗ into lu- minous energy, L∗ (measured in Joules), is a simple problem of scaling the flash light with the Jupiter brightness as detected in each observation, L∗ = K · (cid:32) I∗ (cid:33) , IJ (1) where K is a conversion factor different for each video obser- vation and proportional to the total flux of light reflected from Jupiter and detected with the camera. The conversion factor was calculated in the following way: for each date, we calculated the effective solar constant at the Article number, page 5 of 14 A&A proofs: manuscript no. 32689_astroph Fig. 7. Spectral responses of the combination of cameras and filters for the impacts in March 2016 (top) and May 2017 (bottom). Note the better quantum efficiency of the ASI224MC camera used by S.P. for the May 2017 impact when compared with the ASI120MC used by T.R. and the QHY5LII camera used by G.K. for the impact in March 2016. All the color filters used correspond to Bayer mask filters on the CCD. Fig. 6. Raw light curves of the May 2017 impact. Top panel: Data from Sauveur Pedranghelu. The light curve has different phases that are iden- tified with vertical orange lines: A first flash of 0.21 s, a second phase with constant flux for 0.16 s, a central flash of 0.40 s, and a extended decay for another 0.32-0.61 s for a total flash duration of 1.38 s. Mid- dle pannel: Data from Thomas Riessler showing the double flash with a slightly shorter duration. Bottom panel: Data from André Fleckstein. The video by T.R. does not show the same amount of structure visible in the first light curve, possibly because of the different frame rates and smaller pixel size. The video by A.F. was acquired under poorer seeing conditions. distance of Jupiter, S J, by scaling the solar constant at Earth, S E = 1361 W/m2 to the distance of Jupiter to the Sun, dJ, us- ing the ephemeris computed with the JPL HORIZONS system at https://ssd.jpl.nasa.gov/. For each video observation we convolved the solar spectrum from Colina et al. (1996) with the camera and filter response and the reflectivity spectrum of Jupiter from Karkoschka (1994) by computing the amount of en- ergy reflected from Jupiter and detected by the camera. Because the impact photometry is computed with respect to the full disk Article number, page 6 of 14 brightness of Jupiter, the absolute values of the camera and fil- ter response are not needed, only their relative values at different wavelengths. Thus, (cid:16) L∗ = S J · πReqRp (cid:17)·  (cid:82) ∞ 0 F(cid:12) (λ) · I/F(λ) · C(λ)dλ (cid:82) ∞ 0 F(cid:12) (λ)dλ (cid:33) · (cid:32) I∗ IJ ·∆t, (2) where Req and Rp are the equatorial and polar radius of the planet, F(cid:12)(λ) is the solar spectral radiance, I/F(λ) is the reflec- tivity of Jupiter, C(λ) is the spectral response of the camera and filter (figure 7 shows the spectral responses of the cameras and filters used in the detection of the two impacts), and ∆t is the ex- posure time for a single frame. The first term represents the flux of solar light intercepted by Jupiter. The second term is the pro- portion of this energy that the system can detect and contains the spectrum of Jupiter and the spectral response of the camera. The third term contains the normalization factor from the analysis of the light curve and the integrated light of the Jupiter disk. The ∆t term is used to transform Watts into Joules. The result is the "detected" luminous energy of the flash given in Joules. The flash behaves as a punctual source of light releasing energy in all directions. Part of this light (almost 50%) illumi- R. Hueso et al.: Small impacts on the giant planet Jupiter nates the upper clouds of Jupiter and is reflected with an aver- age approximate albedo of 0.5. This added contribution implies that a geometric correction needs to be introduced in the lumi- nosity evaluation. The magnitude of the correction depends on the viewing geometry of the impact, and its exact evaluation is a complex problem of radiative transfer. For consistency with Hueso et al. (2013), this correction is here simply computed as 1.3, Lcor∗ ≈ L∗/1.3. (3) This approximate correction is applied to the impact in May 2017, but not to the impact in March 2016, which occurred too close to the planet limb, to required adding this correction from reflections of light in the Jupiter clouds. The "detected" and corrected luminous energy Lcor∗ corre- sponds to a part of the total emitted energy in form of light. Depending on the spectral energy distribution of this light, the Lcor∗ can correspond to a higher or lower amount of total emitted light. If we assume that the light is emitted at a given temperature following Planck's blackbody law, we can compute an efficiency factor for different temperatures and detectors,  (cid:82) ∞ (cid:82) ∞ 0 FBB(T, λ) · C(λ)dλ 0 FBB(T, λ)dλ  , L f∗ (T) = Lcor∗ (4) where L f∗ (T) is the total luminous energy of the impact as a func- tion of temperature, T is temperature, and FBB is Planck's law of radiation for a blackbody. We considered that blackbody brigth- ness temperatures of the flash are in the range of [3500-8500] K. These temperatures come from values of Earth's fireballs, SL9 impacts observed by the Galileo spacecraft, and analysis of the 2010 fireball on Jupiter, which was observed simultaneously at high quality with a red and blue filter (Hueso et al. 2010b). This temperature range produces a factor of two uncertainty in the energy calculation, which is larger than the uncertainties in the light-curve calculation or the geometric factor correction. In video observations obtained with cameras that use a Bayer mask to build RGB images, we considered that the image is the sum of the three red, green, and blue channel images, so that C(λ) is the sum of the curves representing the spectral responses of each channel. Finally, in order to transform the total luminous energy into kinetic energy of the impactor, we need to know the efficiency of the impact to convert kinetic energy into luminous energy. For meteoroids and fireballs entering Earth's atmosphere, an empir- ical efficiency formula has been derived by Brown et al. (2002). µ = 0.12E0.115 0 , (5) where E0 is the optical energy measured in kilotons of TNT (1 kton=4.185 × 1012 J). Values of µ from this formula for Jovian impacts range from 0.15 to 0.20. We caution that this formula is calibrated from ob- servations of Earth impacts with optical energies from 0.001 to 1 kiloton, while Jovian impacts release optical energies in the range of 5-60 kiloton. Additionally, the impacts on Jupiter occur at a different velocity with an atmosphere of a different compo- sition. These factors introduce an additional uncertainty in the size of the impact object that is currently unconstrained. 3.3. Masses and sizes of the impacting objects We assumed impacts at a velocity of 60 km s−1 close to the es- cape velocity of Jupiter and densities from 2.0 to 0.25 g cm−3. The results are summarized in Table 2 in comparison with de- terminations of energies and masses of previous impacts. The impact in May 2017 was approximately 4.5 times less energetic than the impact in March 2016. When examining the ensemble of impacts on Jupiter given in Table 2, kinetic energies range 32-405 ktn close to Chelyabinsk- like events, which was considered to release about 450 ktn of en- ergy (Brown et al. 2013) and an order of magnitude lower than the Tunguska impact (5,000-15,000 ktn) (Boslough & Crawford 2008) or 1-3 million times lower than the combined SL9 im- pacts (estimated to release 300,000 kTn of energy) (Boslough & Crawford 1997). 4. Visibility of debris fields caused by intermediate-sized impacts The impacts we characterized have a remarkably small diversity of sizes. The largest of them was caused by an object of at most 19 m in diameter when considering a density of 0.25 gcm−3 sim- ilar to the assumed density for SL-9 fragments (Crawford 1997). The smallest debris field left in the atmosphere of Jupiter by one of the SL-9 fragments was caused by fragment N. This fragment was estimated to have a size of about 45 m in diameter (Craw- ford 1997). For equal density, this is about 12 times more mas- sive than the impact flash detected in September 2012, or 180 times more massive than the smallest impact flash in May 2017. The dark debris of fragment N was observed in HST images be- fore it mixed with debris from other fragments (Hammel et al. 1995). Fragment N would have caused a flash 12 times brighter than the impact in September 2012, producing significant saturation over a standard impact video record (most amateurs expose each frame, so that the brightest part of Jupiter reaches about 70% - 80% of the saturation level in their detector). When we com- pare the estimated mass of fragment N and the impact in 2012, the video observation of that fragment would amount to a star of +3.3 magnitude. This is equivalent to three times the visible magnitude of Ganymede, 1/315 of the total flux of Jupiter, or 2 × 1016 J. A flash of this energy would produce a debris in the atmosphere of Jupiter within the observable reach of HST and Earth's largest telescopes. Material from the N impact and other small SL9 impacts could be observed for at least two days, but there are no reports of them a week after the impact (Spencer & Mitton 1995). Dissipation times of impact material in the atmosphere for previous impacts were on the order of a few months for the 2009 impact (Sánchez-Lavega et al. 2011) with an e-folding time of 10 days in the debris particles concentration (Pérez-Hoyos et al. 2012) and longer for the SL9 largest fragments (Sanchez- Lavega et al. 1998). In these impacts, the debris left in the at- mosphere was spectrally dark in the continuum (with maximum contrast with the environment clouds at red wavelengths), but bright in methane absorption bands. This is so because the dark particles were deposited in the atmosphere at high altitude (pres- sures lower than 10 mbar) (Hammel et al. 1995, 2010; de Pater et al. 2010; Pérez-Hoyos et al. 2012). The impact debris was also bright in the thermal infrared because of the heating effect of an impact and the long radiative time constant of the stratosphere of Jupiter (Harrington et al. 2004; de Pater et al. 2010). Article number, page 7 of 14 Table 2. Jovian bolide analysis A&A proofs: manuscript no. 32689_astroph (J) Mass (103kg=ton) Diameter (m) ρ = 2.0 gcm−3 (J) Optical Energy 0.3 − 2.5 × 1014 0.6 − 2.0 × 1014 1.6 − 3.2 × 1014 1.3 − 2.8 × 1014 1.9 − 3.6 × 1013 Date (yr-mm-dd) 9.3-18 2010-06-03* 12-17 2010-08-20* 15-19 2012-09-10* 14-19 2016-03-17 2017-05-26 8.3-10 Note: (*) Data from Hueso et al. (2013). Densities of 0.25 gcm−3 are considered as representative of the SL-9 impact (Crawford 1997). Kinetic Energy 1.9 − 14 × 1014 3.7 − 11 × 1014 9.0 − 17 × 1014 7.3 − 14 × 1014 1.3 − 2.3 × 1014 Energy (ktn) 46-350 88-260 215-405 175-350 32-55 105-780 205-610 500-950 403-805 75-130 4.7-9.1 5.8-8.4 7.8-9.7 7.3-9.2 4.1-5.0 Diameter (m) ρ = 0.6 gcm−3 Diameter(m) ρ = 0.25 gcm−3 7.0-14 8.7-13 12-14 10.9-13.7 6.1-7.4 To the best of our knowledge, the long-term visibility of a small impact debris field has not yet been explored in the liter- ature. An empirical estimation between the debris lifetime and the size of the impactor would also depend on the nature of the impactor (stony or icy), location of the impact in the atmosphere in a region with higher or lower wind shear, and many other pos- sible parameters (such as the impact trajectory angle with the planet). However, in the event of an impact 10 times brighter than the flashing impact that occurred in September 2010, we predict that quick follow-up observations could detect an atmo- spheric debris field. We examine the detectability of impact debris on Jupiter caused by larger impacts in section 7.3. 5. Searches for new impacts 5.1. Dedicated detection campaigns Professional dedicated campaigns to detect impacts on Jupiter are difficult to carry out because such campaigns must involve the capacity of acquiring and analyzing images over many dif- ferent nights. Our team runs frequent observations of Jupiter with 1-2 m size telescopes using PlanetCam UPV/EHU, a lucky- imaging instrument (Sanchez-Lavega et al. 2012; Mendikoa et al. 2016). In the past five years, we have observed Jupiter over 20 different campaigns over an average of two nights per cam- paign, acquiring about 2.0 hours of data for each night. These observations have been checked for impacts without observing an impact flash for an accumulated observing time of 80 hours. Details of the observations are given in Mendikoa et al. (2016, 2017). The lack of impact detections in this survey imposes a weak constraint over the maximum impact rate on the planet. We used a simple Monte Carlo simulation to calculate the significance of this negative detection. We tested different values of the num- ber of observable impacts per year, and for each, we launched a large-number (5,000) of Monte Carlo simulations where we examined the number of impacts that would occur in 80 hours. To do this, we divided 80 hours into 4,800 opportunities of 1 minute to detect an impact. We examined the statistics of the Monte Carlo simulation and searched for the number of observ- able impacts per year that result in 66% of the simulations pro- ducing at least one impact in an accumulated observing time of 80 hours. The statistical result is that an impact rate of 120 de- tectable impacts per year would be needed to find one impact in this observing time. The Monte Carlo simulation shows that for a probability of 90% to find an impact in a survey of 80 hours, we would require an impact rate of 250 observable impacts per year. This analysis suggests an upper limit on the impact rate smaller than 120 impacts per year. Article number, page 8 of 14 A project run by Japanese amateurs called "Find Flash" and run by the Association of Lunar and Planetary Observers (ALPO) in Japan with more than 50 amateur observers and about 10 nights per year observation time on 1 m size telescopes did not find impacts for about four years of observations, placing a similar constraint (I. Tabe, private communication). 5.2. Filters and technology Flashes can be best detected in filters where the planet is dark and the integration time of the camera is not too long. Blue fil- ters and relatively wide filters centered on the methane absorp- tion band at 890 nm are best suited for a flash detection because the planet is dark and the flash should be bright. Integration times lower than 0.2 s are needed, suggesting that while blue filters can be used with small telescopes, an instrument with a minimum diameter of 30 cm might be required to perform a flash detec- tion campaign for the 890 nm methane band. Dual observations in blue and 890 nm would highly constrain the brigthness tem- perature of the flash, allowing for a determination of the impact energy with lower ambiguities. The cameras used in the amateur community have experi- enced significant improvements over the past decade from read noises of ∼ 8e- at 0 gain (e.g., in the popular Point Grey Flea3 camera used for the first detections of impacts in Jupiter) to ∼ 3e- at 0 gain (in the ASI cameras now used by most amateurs). Quantum efficiencies have improved at least by a factor of two in the near-infrared, where most observers concentrate their obser- vational efforts, and have increased from 5% to 30% in the 900 nm range. Better sensitivity results in impacts being detectable by smaller telescopes with faster exposures and better temporal resolution of the light curves. 5.3. Impacts DeTeCtion software We have written a software tool to inspect amateur video obser- vations of Jupiter capable of detecting impacts in the planet. The latest version of this software, DeTeCt3.1, is an open-source ap- plication that amateur observers can use on their own computers (Delcroix et al. 2017). The software constitutes one of the activ- ities of the "Planetary Space Weather Services" (PSWS) (André et al. 2018) and is based on differential photometry of coregis- tered images. This software is essentially different from "flash detection" software developed for search of impact flashes on the Moon (Madiedo et al. 2015) because these tools generally examine videos acquired over a large field of view (almost the full non-illuminated side of the Moon) without the need to cor- rect for seeing effects (coregistration) and because the flash is R. Hueso et al.: Small impacts on the giant planet Jupiter recorded over a dark object and not the bright background of Jupiter. The DeTeCt software has been used regularly by dozens of observers examining about 70,000 video files, which is equivalent to more than 75 days of observations distributed unevenly over several years. The software produces log files that are later analyzed to examine the statistics of nondetec- tions when comparing with the fortuitous detections of im- pacts. The analysis shows that about 5% of all observations were acquired at the same time by the collaborating observers. The software and statistical analysis of its results can be ac- cessed at http://www.astrosurf.com/planetessaf/doc/ project_detect.shtml. A similar analysis of video observa- tions of Saturn is also available at that web site, with negative results so far. From the statistics of the time covered by these observa- tions, and because one of the observers of impacts in Jupiter (J. McKeon) uses this software regularly, one can consider a de- tectable impacts rate of 1/75 days−1), which is equivalent to 4.9 observable impacts per year. This is a minimum number because most video files are acquired with regular sky conditions, and we cannot measure the eficiency of finding impacts on low-quality video files. 6. Statistical analysis of the impacts 6.1. Continuous flux or meteor showers? Several works have predicted the existence of meteor show- ers on other planets based on the characteristics of known comet/meteoroid orbits that cross the nodal points of planets. Most of them considered only the inner planets (e.g., Christou 2010), but Selsis et al. (2004) presented a study for all solar sys- tem planets, including Jupiter and Saturn. They found 48 "comet candidates" that could produce meteor showers on Jupiter. In contrast to what happens on the inner planets, meteor showers in Jupiter largely overlap in time because of the very long duration of close comet passages. We therefore assume that impacts of the class detected by amateur astronomers should be distributed randomly in time and not be clustered when the Jupiter orbit tra- verses a cometary tail. We also assume that these impacts occur close to the escape velocity of Jupiter of 60 km s−1 , in agreement with observations of the SL9 impact (Harrington et al. 2004). 6.2. Temporal survey of Jupiter from amateur observers Understanding the significance of the five flashes requires knowledge about when the amateur community performs their observations of Jupiter. Figure 8 shows a statistical analysis of Jupiter observations archived at the PVOL database (Hueso et al. 2010b, 2018). This is one of the most popular databases of am- ateur observations of solar system planets and can be searched with very many parameters. In the period from January 2010 to December 2017, the PVOL database contains 17,643 Jupiter observations that are representative of global trends in amateur observations of Jupiter. Each year, the observations cluster more abundantly close to the opposition of Jupiter. In years where this opposition is close to winter in the north hemisphere, fewer am- ateurs are able to observe the planet regularly. This is due to the geographical distribution of most amateur observers. About 65% of all Jupiter observations come from observers in the north hemisphere, with about 21% of observations contributed from ±30◦ latitudes and 16% from south hemisphere latitudes. The number of Jupiter observations for the 2010-2011 Jupiter appari- tion was about 2,400. This number reduced by about 25% in the three Jupiter apparitions in 2013-2015 and increased by about 15% in the latest 2016 and 2017 Jupiter apparitions. This trend with Jupiter oppositions helps to explain the gap in the detection of Jupiter flashes in the period 2013-2015 when Jupiter opposition resulted in most observers having difficulties to find good weather and with generally fewer observations (Fig- ure 8). Jupiter oppositions in 2018-2022 will occur from May to September, offering increased capabilities of detecting impacts. 6.3. Statistical interpretation of the flashes It is difficult to make statistical analysis of events that have been observed only a few times. The results in section 5 provide abso- lute upper limits and weak lower limits to the observable impacts on Jupiter. Here we present different arguments to infer the num- ber of detectable impacts on Jupiter. 1. An absolute minimum flux of impacts on Jupiter of 0.63 im- pacts per year is found based on the five flashes detected in eight years (2010-2017). 2. The statistical analysis from DeTeCt can also be understood as a minimum flux of 4.9 impacts per year. Only one of the 11 observers that have successfully found an impact on Jupiter collaborates with this project regularly. If this fraction is rep- resentative of impacts on the planet, then the 4.3 impacts per year could scale up to 52 impacts per year. 3. Hueso et al. (2013) gave order-of-magnitude estimates con- sidering the geographical distribution of observers and the number of observations per year of the amateur community, inferring between 6-30 detectable impacts per year based on the three impacts detected from 2010-2012. We here correct these estimates with an update of the number of impacts de- tected in the period 2010-2017. We assume that the total sur- veyed time is given by T = N · t1 · e, (6) where N is the total number of images, 17,643 as stated above, t1 is the time accumulated to form each image, and e is the efficiency for each image to have enough quality to show an impact. t1 can be from 5 min. to 15 min. since each image is the result of a longer observing session. The efficiency e in which a video observation can have enough quality so as to show an impact was estimated to be from 0.3-0.5 in Hueso et al. (2013). Then T is approximately 18- 91 days over an accumulated time of eight years. Only 2 of the 11 observers who have found impacts in Jupiter are reg- ular contributors to this database, and we estimate that the global network of amateur astronomers can be represented by an increase in Jupiter observing time by a factor of 11/2. Therefore, we consider that the global survey of Jupiter ob- servations by the amateur community can be globally rep- resented by a total observation time of 99-500 days obtained over the past eight years. This represents a global observation efficiency of Jupiter of 3.4-17%. In this way, the five impacts detected in eight years may scale up to an estimate of 4-18 "detectable" impacts per year. Even if we were to consider the amateur observations as a "perfect survey" with a detec- tion efficiency of e = 1.0, the detectability of impacts would still be limited by the geographical distribution of observers clustered in North and South America, Europe, and Japan- Australia. This would result in an observing time efficiency of 33% of the total available time and a minimum number of 2.5 detectable impacts per year. Article number, page 9 of 14 A&A proofs: manuscript no. 32689_astroph Fig. 8. Number of Jupiter images per month archived in the PVOL database (top) and observation conditions (bottom) as a function of time. Jupiter oppositions are shown with solid green lines and are labeled on the uppermost horizontal axis. Dates of Jupiter impacts are shown with vertical dotted magenta lines and yellow stars. The region where the impact was observed labels each line. Spring and summer months in the north hemisphere are shown in gray, and winter and fall in the north hemisphere are shown in white. Bottom panel: Jupiter declination (left axis and black line) compared with the planet apparent size in arcsec (right axis and orange line) compared with its visual magnitude (blue line and axis). The horizontal dotted gray line shows the zero-declination line. 4. Of the 11 observers that have detected impacts in Jupiter, 2 can be considered as very regular observers performing an outstanding number of Jupiter observations every year and participating in several research projects (A.W. and C. G.). One of them (A.W.) discovered the debris of an impact in 2009 (Sánchez-Lavega et al. 2010) and the first flash of light the next year. A.W. accumulates 180 hours of Jupiter ob- servations per year, personally looking at every video, and half of these data have a quality good enough to visualize a small impact (90 hours per year). This means that he has found one (two) impact(s) in an accumulated observing time of 720 (810) hours over the past eight (nine) years, where the number in parentheses indicate whether we also consider his initial finding of the 2009 large impact. For A.W. alone, this is about 35-40% of all the time covered by the DeTeCt program. This suggests detectable impacts with a rate close to 10-20 impacts per year. 5. The second impact on Jupiter was found only two months af- ter the first. The clustering of these two events close in time can be examined with Monte Carlo simulations. We simu- lated impacts considering different impact frequencies in a Monte Carlo simulation representative of eight years, where for each day, we calculated the random probability of hav- ing found an impact. Detection was examined considering (i) that each day, the planet could only be observed 33% of the time because of the longitudinal distribution of Earth ob- servers that peaks over Europe, North America, and Japan meridians; (ii) that observations only cover nine months of a year; (iii) that detecting the impact was not possible be- cause of poor weather 50% of the time; and (iv) that detec- tions were not possible because of poor atmospheric seeing 50% of the time. This renders the detection probability of any given impact as 6.3%. Two impacts occur with a time differ- ence shorter than three months in about 50% of the Monte Carlo simulations, with an impact flux of 5-15 impacts per year. All in all, we consider that a reasonable estimate of the num- ber of potentially detectable impacts per year in Jupiter from objects of 5-20 m size or larger can be on the order of 4-25. However, since we only observe the planet nine months every year and we can only observe about one half of its surface at any given time, the "detectable" number of impacts corresponds to a higher number of objects colliding with Jupiter. This correction means that the accumulated flux of objects larger than 5-20 m in diameter that hit Jupiter every year is estimated to be 10-65, compared with 12-60 from Hueso et al. (2013). 7. Impact flux on Jupiter 7.1. Consequences for the chemichal composition and dust abundance in the Jovian stratosphere Based on the flux rate derived in the previous section, the contri- bution of impacts of this size range to the delivery of chemical species and dust to the upper atmosphere of Jupiter is expected to be on the order of 8 × 105 − 7 × 107 kg yr−1. Recent research on Earth large-size bolides shows that the largest fraction of the impacting object is deposited in the atmosphere in the form of micrometer dust (Klekociuk et al. 2005). This contribution to exogenous species and dust can be compared with the contin- uous contribution from interplanetary dust particles (IDPs) col- Article number, page 10 of 14 R. Hueso et al.: Small impacts on the giant planet Jupiter Fig. 9. Impact rates on Jupiter and Earth compared. The vertical gray region represents the sizes of the five bolides, and the blue box represents the impact rate of 5-20 m size objects on Jupiter. The lines represent impact rates from: (a) dynamical models of comets (red line with estimated uncertainties as the dotted line, Levison et al. 2000; (b) orange line and estimated uncertainties from corrections to that model introduced by Bézard et al. 2002, or; (c) the cratering record of Galilean moons (magenta line from Zahnle et al. 2003; blue line from Schenk et al. 2004). This is compared with estimates of the impact rate on Earth from Brown et al. 2002). The yellow regions indicate the limiting size of objects that might be discovered based on the dark debris that they are expected to leave in the atmosphere of Jupiter: The smallest SL9 fragment that produced a detectable debris field on HST images (fragment N) and long-standing debris features associated with objects of 170 m such as SL9 fragment D are highlighted. The green vertical region represents objects leaving a debris field that might be detected with amateur equipment. The light-brown box represents estimates of impact rates on Jupiter from Sánchez-Lavega et al. (2010) for 0.5-1.5 km size objects. Estimates of the impact rate of intermediate-size objects that hit the planet and form observable debris fields are plotted as red stars. The error bars represent an uncertainty factor of 2.5 upward and downward. Figure updated from Hueso et al. (2013). liding with Jupiter. Current models of interplanetary dust fluxes on the giant planets (Poppe 2016) predict about 10−13 g m−2 s−1 or 1.9 × 109 kg yr−1. The dust observations made by the Galileo Dust Detection System (DDS) of impact-induced ejecta clouds around the Galilean Moons (Krüger et al. 1998, 2000, 2003) re- sult in estimates of the total mass flux of IDPs on Jupiter of 3 × 10−13 g m−2 s−1 (Sremcevi´c et al. 2005), which is equiv- alent to 5.6 × 109 kg/yr. However, these estimates are probably uncertain by an order of magnitude (Poppe 2016). Thus, the con- tribution of impacts of the size range such as those discovered in observations of Jupiter fireballs contribute about 0.01-3.7 % of the IDP exogenous material and dust to the upper stratosphere of Jupiter and are not expected to have a strong impact even on local scales. Lellouch et al. (2002) have analyzed ISO observations and found an upper limit (8 × 104 cm−2 s−1) to the permanent water influx into the stratosphere of Jupiter. Bézard et al. (2002) have placed oxygen influx limits at (1.5− 10)× 106 cm−2 s−1 based on observations of CO. Various possibilities for the high exogenous CO/H2O ratio at Jupiter have been considered by Bézard et al. (2002) and are further discussed by Poppe (2016) and Moses & Poppe (2017). The main conclusion is that favorable production of CO over H2O during IDP ablation is needed to explain this result. In the case of the contribution of 5-20 m size impacts, the same conclusions apply, that is to say that the incoming water in impacts has to be transformed into CO. The water influx of 8×104 cm−2 s−1 from Lellouch et al. (2002) is equivalent to a flux of 143,000 kg of water per year. This amount of water could be supplied by a single impact of 10 m with 30% of water if water could be preserved during the impact. We conclude that the water molecules ablating from the incoming impacting object must be thermochemically converted into CO soon after ablation. Article number, page 11 of 14 Table 3. Summary of impacts on Jupiter Table 4. Predictions of impacts per year on Jupiter that leave observable debris fields, and their detectability A&A proofs: manuscript no. 32689_astroph Date 1981-03-05 1994-07-16 to 1994-07-24 2009-07-19 2010-06-03 2010-08-20 2012-09-10 2016-03-17 2017-05-26 Mass 11 kg 1.0 × 109 Tn 6.0 × 107 Tn 105 − 780 Tn 205 − 610 Tn 500 − 950 Tn 403 − 805 Tn 75 − 130 Tn Reference Cook & Duxbury (1981) Hammel et al. (1995) Harrington et al. (2004) Sánchez-Lavega et al. (2010) Hueso et al. (2010b, 2013) Hueso et al. (2013) Hueso et al. (2013) This work This work 7.2. Impact flux as a function of impactor size Table 3 summarizes the masses of all impacts observed on Jupiter from the SL9 series of impacts down to a very small meteor entering the atmosphere of Jupiter that was observed by Voyager 1 (Cook & Duxbury 1981). Other plausible but uncon- firmed collisions with Jupiter have been proposed, such as a dark spot on Jupiter observed by Cassini in 1690 with morphological characteristics and behavior similar to SL9 debris features (Tabe et al. 1997) that were proposed to be caused by an impact with a 600 m object (Zahnle et al. 2003). Figure 9 presents an update of results presented in Hueso et al. (2013) of our understanding of the current flux of impacts on Jupiter. The results from this work are very similar to our previous analysis. The deduced impact rate for 5-20 m size ob- jects compared with estimates of greater impacts from Sánchez- Lavega et al. (2010) can be interpolated to predict plausible im- pact rates of intermediate-size objects, as shown with stars in Figure 9. Error bars here represent uncertainties of 2.5 higher or lower from the overall uncertainty of the impact rate of small- size impacts (blue box in Figure 9). This is compared with ex- pectations from dynamical models of comets prone to Jupiter collisions (Levison et al. 2000) and the cratering record on the Galilean moons (Zahnle et al. 2003; Schenk et al. 2004). Predic- tions of impact rates from this study are comparable to the upper estimates of impact rates from Levison et al. (2000). Our results for small-size impacts clearly depart from impact rates estimated from cratering of the Jovian moons, which for small impacts are dominated by young craters on Europa (Schenk et al. 2004). 7.3. Searches of debris left by impacts We now focus on predictions based on Figure 9 of the plausible impact rate of larger objects that might be detected as an intense flash and might leave an observable trace in the atmosphere. The detectability of these impacts depends not only on the impact frequency in Figure 9, it also depends on their size and the bi- ases associated with their detection. Predicted impact rates in the planet and their detectatibility are summarized in Table 4 and are discussed below. – Large superbolides. Voyager and Cassini observed Jupiter at high resolution over a period of at least three months. For Cassini, a global coverage at spatial resolutions of about 140 km/pix or better was acquired for at least 15 days (Porco et al. 2003; Salyk et al. 2006). Objects of 5-20 m hitting the planet with the flux rate deduced from this work would give a non-negligible probability (0.5-3 impacts over 15 days) to have occurred in the course of this Cassini 15-day window. Higher resolutions over particular regions were acquired for another 30 days. Thus, small debris fields in the methane Article number, page 12 of 14 Size (m) Mean Min (yr−1) 1.1 0.17 0.06 (yr−1) 2.8 0.43 0.15 45 170 380 Max (yr−1) 7.0 1.1 0.37 Detectability (yr) 0.4-2.6 2-12 6-30 Note: The impact frequency is the number of impacts that we estimate to occur on Jupiter every year. Their detectability (the mean number of years between observable impacts of a given size and larger) is affected by observational biases, as discussed in detail in section 7.3. band and ultraviolet images where the debris maximizes its contrast might exist in the Cassini imaging data. On Earth, many satellites have observed high-atmosphere debris asso- ciated with impacts. The best case are the satellite observa- tions of the Chelyabinsk impact, which was observed at a variety of spatial resolutions 1-10 km/pix for at least 3 hr (Miller et al. 2013). This impact was similar in energy to the Jupiter impacts we discussed here. Although other missions have imaged Jupiter (Pioneers, New Horizons, and Juno), the number of images from these missions is too low to merit a specific analysis, but for Cassini and the Voyagers, a search for tiny- and small-debris fields might place an important constraint in the rate of impacts on the planet. – Small impacts. Based on this analysis, objects of 45 m or larger that leave a short-lived debris field that is only observ- able with large telescopes may impact Jupiter once every 0.36 years with uncertainties from 0.14 to 0.9 years. Since they can impact on the far side of the planet or in months when Jupiter is not observable, even a perfect survey of im- pacts on Jupiter could only find these events once per year with estimated uncertainties from 0.4 to 2.6 years. HST ob- servations of the fragment N impact site did not allow deter- mining for how long debris from this impact might be ob- servable. The visibility of such an impact may also depend on its latitudinal location and on dissipation effects such as local wind shear. A careful examination of archived HST images of Jupiter acquired since 1991 (the date of the first Jupiter observation) and a search for tiny dark spots in the visible or that are bright in methane may place additional constraints on the impact rate on Jupiter or serve to lower the impact rate deduced from small-size flashing impacts. This is a non-trivial effort because archived HST Jupiter images encompass more than 350 target names, several instruments, tens of filters, and observing programs that range from global coverage to snapshots, and the debris field might be within the limit of detectability in most filters. Additionally, small impact debris fields would not be observable in subpolar lat- itudes. – Intermediate-size impacts. Objects larger than 170 m that leave a debris field that might be observable with small tele- scopes over weeks and months with professional telescopes may occur once every 2.3 years with estimated uncertainties from 1 to 6 years. Their detectability is difficult to ascertain because these events might be observed in only about half of each year when Jupiter is well placed in the night sky for as- tronomical observation. Therefore a timescale of 2-12 years seems reasonable for the detectability of these events. R. Hueso et al.: Small impacts on the giant planet Jupiter – Large impacts. Objects larger than 380 m that are able to leave a debris field that is observable in standard amateur images over weeks and during months in observations with professional telescopes may only occur once every 7 years with uncertainties from 3 to 16 years. Again, since Jupiter is only observable nine months every year and the quality of observations is a function of the proximity to Jupiter op- position, these events may be discovered by amateurs about once every 6-30 years, similarly to the timescale separation between the SL-9 collision and the 2009 impact. 8. Conclusions The most recent two impacts on Jupiter in March 2016 and May 2017 had masses of 310-620 Tn and 75-130 Tn, respectively, with sizes ranging from 4.1 m to 17 m for object densities from 2.0 to 0.25 g cm−3. These masses are comparable to previous impacts on Jupiter that have also been found in flashes in video observations of the planet. The cumulative impact rate of objects of this size range or larger is predicted to lie in the range of 10-65 per year, with only 4-25 impacts per year being observable in a perfect survey of flashes because of the way they are distributed over Jupiter's visible and far sides and the Jupiter observation period per year. The overall impact rate for Jupiter according to this work is similar to the high impact rate limit of Levison et al. (2000) and compatible with the modifications discussed by Bézard et al. (2002). Although significant uncertainties exist, the observations of impact flashes on Jupiter disprove the low impact rate pre- dicted from the cratering record on Europa that was discussed by Schenk et al. (2004). 5-20 m size objects impacting Jupiter are expected to be detected yearly in the next Jupiter oppositions because of the improved observing conditions and the availability of software tools. Future observations of impacts with the modern cameras cur- rently available to the amateur community could be obtained at 60 fps or higher with a good signal-to-noise ratio. For an ener- getic impact like those of 2012 and 2016, this may allow explor- ing the fragmentation history of impacting objects, opening the possibility of studying the nature of the impacting object (stony, metallic, icy compact, or icy porous). The accumulated effect of these impacts on the chemistry of the upper stratosphere of the planet is negligible when compared with other sources of exogenous chemicals, such as interplane- tary dust particles and giant impacts. A "large flash" comparable to fragment N of SL9 leaving an observable debris field at the limit of spatial resolution with HST, VLT, or other large telescopes might occur on Jupiter with a typical timescale of once every 2-11 months. A perfect ob- servational survey of bright flashes would find these powerful flashes about once per year. A dedicated careful examination of all HST observations of Jupiter obtained since 1991 that would search for small dark spots on visible images and bright spots on methane images mighthelp to reduce the factor of 6 uncer- tainty on the impact rate from this work. A similar search for smaller impact debris from Voyager and Cassini images seems worthwhile, based on this study. An extremely intense flash leaving a standing debris field in the atmosphere of Jupiter that could be observed with HST or large ground-based telescopes over weeks might occur on Jupiter once every 1-6 years. Regular ground-based observations of de- bris fields on Jupiter might detect these events about once every 2-12 years and more efficiently than a survey of flashes. Impacts leaving a debris field that would be observable with amateur equipment might occur on Jupiter once every 3-16 years and might be observable once every 6-30 years when accounting for the time of the year when Jupiter can be observed at high resolution by amateur astronomers. Future observations will find increasingly smaller impacts as the technology improves. Specific searches in Voyager, Cassini, and HST images may contain small-debris fields that were not detected at the time of the acquisition of these observations. If these "small impact scars" are detected, they will largely con- strain the impact rate on Jupiter. The JUICE mission to Jupiter may also discover flashing impacts on the planet that are caused by objects of much smaller sizes through the long surveys of the Jupiter night side that are currently planned for studies of the Jovian magnetosphere and the deep lightening activity (Grasset et al. 2013). Acknowledgements. We are very grateful to Thomas Riessler for permission to work on his recording of the May 2017 impact. We are also grateful to Emmanuel Lellouch for a detailed and constructive review of this research. We thank C. Go, M. Tachikawa, K. Aoki, M. Ishimaru, D. Petersen, and G. Hall from their obser- vation of impacts in Jupiter as well as I. Tabe for help in communicating with Japanese amateur astronomers. We also thank Sebastian Voltmer for providing early information from Gerrit Kernbauer's video detection and image process- ing of this video, and Dan Fischer for his help in finding German-speaking ob- servers that had observed the May 2017 impact. We are also very grateful to the ensemble of amateur astronomers running DeTeCt on their video observations of Jupiter, and to the large community of observers providing Jupiter observa- tion data through PVOL and ALPO Japan. R.H. and A.S.L. were supported by the Spanish project AYA2015-65041 (MINECO/FEDER, UE), Grupos Gobierno Vasco IT-765-13 and UPV/EHU UFI11/55. This work has been supported by the Europlanet 2020 Research Infrastructure. Europlanet 2020 RI has received funding from the European Union's Horizon 2020 research and innovation pro- gramme under grant agreement No 654208. References André, N., Grande, M., Achilleos, N., et al. 2018, Planet. Space Sci., 150, 50 Bézard, B., Lellouch, E., Strobel, D., Maillard, J.-P., & Drossart, P. 2002, Icarus, Borovicka, J., Spurný, P., Brown, P., et al. 2013, Nature, 503, 235 Borovicka, J., Spurný, P., Grigore, V. I., & Svoren, J. 2017, Planet. Space Sci., 159, 95 143, 147 Boslough, M. B. E. & Crawford, D. A. 1997, Annals of the New York Academy Boslough, M. B. E. & Crawford, D. A. 2008, International Journal of Impact Brown, P., Spalding, R. E., ReVelle, D. O., Tagliaferri, E., & Worden, S. P. 2002, of Sciences, 822, 236 Engineering, 35, 1441 Nature, 420, 294 Brown, P. G., Assink, J. D., Astiz, L., et al. 2013, Nature, 503, 238 Cavalié, T., Feuchtgruber, H., Lellouch, E., et al. 2013, A&A, 553, A21 Christou, A. A. 2010, MNRAS, 402, 2759 Colina, L., Bohlin, R. C., & Castelli, F. 1996, AJ, 112, 307 Cook, A. F. & Duxbury, T. C. 1981, J. Geophys. Res., 86, 8815 Crawford, D. A. 1997, Annals of the New York Academy of Sciences, 822, 155 de Pater, I., Fletcher, L. N., Pérez-Hoyos, S., et al. 2010, Icarus, 210, 722 Delcroix, M., Hueso, R., Juaristi, J., & Sanchez-Lavega, A. 2017, European Plan- etary Science Congress, 11, EPSC2017 Grasset, O., Dougherty, M. K., Coustenis, A., et al. 2013, Planet. Space Sci., 78, Hammel, H. B., Beebe, R. F., Ingersoll, A. P., et al. 1995, Science, 267, 1288 Hammel, H. B., Wong, M. H., Clarke, J. T., et al. 2010, ApJ, 715, L150 Harrington, J., de Pater, I., Brecht, S. H., et al. 2004, Lessons from Shoemaker- Levy 9 about Jupiter and planetary impacts, ed. Bagenal, F., Dowling, T. E., & McKinnon, W. B., 159–184 Hueso, R., Juaristi, J., Legarreta, J., et al. 2018, Planet. Space Sci., 150, 22 Hueso, R., Legarreta, J., Pérez-Hoyos, S., et al. 2010a, Planet. Space Sci., 58, Hueso, R., Pérez-Hoyos, S., Sánchez-Lavega, A., et al. 2013, A&A, 560, A55 Hueso, R., Wesley, A., Go, C., et al. 2010b, ApJ, 721, L129 Karkoschka, E. 1994, Icarus, 111, 174 Klekociuk, A. R., Brown, P. G., Pack, D. W., et al. 2005, Nature, 436, 1132 Korycansky, D. G., Harrington, J., Deming, D., & Kulick, M. E. 2006, ApJ, 646, 1 1152 642 Article number, page 13 of 14 A&A proofs: manuscript no. 32689_astroph Krüger, H., Grün, E., Hamilton, D. P., et al. 1998, Planet. Space Sci., 47, 85 Krüger, H., Krivov, A. V., & Grün, E. 2000, Planet. Space Sci., 48, 1457 Krüger, H., Krivov, A. V., Sremcevi´c, M., & Grün, E. 2003, Icarus, 164, 170 Lellouch, E., Bézard, B., Moses, J. I., et al. 2002, Icarus, 159, 112 Levison, H. F., Duncan, M. J., Zahnle, K., Holman, M., & Dones, L. 2000, Icarus, Madiedo, J. M., Ortiz, J. L., Morales, N., & Cabrera-Caño, J. 2015, Planet. Space Sci., 111, 105 Mendikoa, I., Sánchez-Lavega, A., Pérez-Hoyos, S., et al. 2016, PASP, 128, Mendikoa, I., Sánchez-Lavega, A., Pérez-Hoyos, S., et al. 2017, A&A, 607, A72 Miller, S. D., Straka, W. C., Bachmeier, A. S., et al. 2013, Proceedings of the National Academy of Sciences, 110, 18092 Moses, J. I. & Poppe, A. R. 2017, Icarus, 297, 33 Mousis, O., Hueso, R., Beaulieu, J.-P., et al. 2014, Experimental Astronomy, 38, Palotai, C., Korycansky, D. G., Harrington, J., Rebeli, N., & Gabriel, T. 2011, 143, 415 035002 91 ApJ, 731, 3 221, 1061 Pérez-Hoyos, S., Sanz-Requena, J. F., Sánchez-Lavega, A., et al. 2012, Icarus, Pond, J. W. T., Palotai, C., Gabriel, T., et al. 2012, ApJ, 745, 113 Poppe, A. R. 2016, Icarus, 264, 369 Porco, C. C., West, R. A., McEwen, A., et al. 2003, Science, 299, 1541 Salyk, C., Ingersoll, A. P., Lorre, J., Vasavada, A., & Del Genio, A. D. 2006, Icarus, 185, 430 Sanchez-Lavega, A., Gómez, J. M., Rojas, J. F., et al. 1998, Icarus, 131, 341 Sánchez-Lavega, A., Orton, G. S., Hueso, R., et al. 2011, Icarus, 214, 462 Sanchez-Lavega, A., Rojas, J. F., Hueso, R., et al. 2012, in Society of Photo- Optical Instrumentation Engineers (SPIE) Conference Series, Vol. 8446, So- ciety of Photo-Optical Instrumentation Engineers (SPIE) Conference Series Sánchez-Lavega, A., Wesley, A., Orton, G., et al. 2010, ApJ, 715, L155 Schenk, P. M., Chapman, C. R., Zahnle, K., & Moore, J. M. 2004, Ages and interiors: the cratering record of the Galilean satellites, ed. F. Bagenal, T. E. Dowling, & W. B. McKinnon, 427–456 Selsis, F., Brillet, J., & Rapaport, M. 2004, A&A, 416, 783 Spencer, J. R. & Mitton, J., eds. 1995, The great comet crash Sremcevi´c, M., Krivov, A. V., Krüger, H., & Spahn, F. 2005, Planet. Space Sci., 53, 625 Tabe, I., Watanabe, J.-I., & Jimbo, M. 1997, PASJ, 49, L1 Zahnle, K., Schenk, P., Levison, H., & Dones, L. 2003, Icarus, 163, 263 Article number, page 14 of 14
1109.6350
2
1109
2012-02-15T00:49:59
Observational and Dynamical Characterization of Main-Belt Comet P/2010 R2 (La Sagra)
[ "astro-ph.EP" ]
We present observations of comet-like main-belt object P/2010 R2 (La Sagra) obtained by Pan-STARRS 1 and the Faulkes Telescope-North on Haleakala in Hawaii, the University of Hawaii 2.2 m, Gemini-North, and Keck I telescopes on Mauna Kea, the Danish 1.54 m telescope at La Silla, and the Isaac Newton Telescope on La Palma. An antisolar dust tail is observed from August 2010 through February 2011, while a dust trail aligned with the object's orbit plane is also observed from December 2010 through August 2011. Assuming typical phase darkening behavior, P/La Sagra is seen to increase in brightness by >1 mag between August 2010 and December 2010, suggesting that dust production is ongoing over this period. These results strongly suggest that the observed activity is cometary in nature (i.e., driven by the sublimation of volatile material), and that P/La Sagra is therefore the most recent main-belt comet to be discovered. We find an approximate absolute magnitude for the nucleus of H_R=17.9+/-0.2 mag, corresponding to a nucleus radius of ~0.7 km, assuming an albedo of p=0.05. Using optical spectroscopy, we find no evidence of sublimation products (i.e., gas emission), finding an upper limit CN production rate of Q_CN<6x10^23 mol/s, from which we infer an H2O production rate of Q_H2O<10^26 mol/s. Numerical simulations indicate that P/La Sagra is dynamically stable for >100 Myr, suggesting that it is likely native to its current location and that its composition is likely representative of other objects in the same region of the main belt, though the relatively close proximity of the 13:6 mean-motion resonance with Jupiter and the (3,-2,-1) three-body mean-motion resonance with Jupiter and Saturn mean that dynamical instability on larger timescales cannot be ruled out.
astro-ph.EP
astro-ph
Submitted, 2011-09-27; Accepted, 2012-02-14 Preprint typeset using LATEX style emulateapj v. 5/2/11 OBSERVATIONAL AND DYNAMICAL CHARACTERIZATION OF MAIN-BELT COMET P/2010 R2 (LA SAGRA) Henry H. Hsieh1,a, Bin Yang1, Nader Haghighipour1, Bojan Novakovi´c2, Robert Jedicke1, Richard J. Wainscoat1, Larry Denneau1, Shinsuke Abe3, Wen-Ping Chen3, Alan Fitzsimmons4, Mikael Granvik5, Tommy Grav6, Wing Ip3, Heather M. Kaluna1, Daisuke Kinoshita3, Jan Kleyna1, Matthew M. Knight7, Pedro Lacerda4,b, Carey M. Lisse8, Eric Maclennan9, Karen J. Meech1, Marco Micheli1, Andrea Milani10, Jana Pittichov´a1, Eva Schunova1,11, David J. Tholen1, Lawrence H. Wasserman7, William S. Burgett1, K. C. Chambers1, Jim N. Heasley1, Nick Kaiser1, Eugene A. Magnier1, Jeffrey S. Morgan1, Paul A. Price12, Uffe G. Jørgensen13,14, Martin Dominik15,c, Tobias Hinse16, Kailash Sahu17, and Colin Snodgrass18, 2 1 0 2 b e F 5 1 . ] P E h p - o r t s a [ 2 v 0 5 3 6 . 9 0 1 1 : v i X r a Submitted, 2011-09-27; Accepted, 2012-02-14 ABSTRACT We present observations of the recently discovered comet-like main-belt object P/2010 R2 (La Sagra) obtained by Pan-STARRS 1 and the Faulkes Telescope-North on Haleakala in Hawaii, the University of Hawaii 2.2 m, Gemini-North, and Keck I telescopes on Mauna Kea, the Danish 1.54 m telescope (operated by the MINDSTEp consortium) at La Silla, and the Isaac Newton Telescope on La Palma. An antisolar dust tail is observed to be present from August 2010 through February 2011, while a dust trail aligned with the object's orbit plane is also observed from December 2010 through August 2011. Assuming typical phase darkening behavior, P/La Sagra is seen to increase in brightness by > 1 mag between August 2010 and December 2010, suggesting that dust production is ongoing over this period. These results strongly suggest that the observed activity is cometary in nature (i.e., driven by the sublimation of volatile material), and that P/La Sagra is therefore the most recent main-belt comet to be discovered. We find an approximate absolute magnitude for the nucleus of HR = 17.9± 0.2 mag, corresponding to a nucleus radius of ∼ 0.7 km, assuming an albedo of p = 0.05. Comparing the observed scattering surface areas of the dust coma to that of the nucleus when P/La Sagra was active, we find dust-to-nucleus area ratios of Ad/AN = 30−60, comparable to those computed for fellow main- belt comets 238P/Read and P/2008 R1 (Garradd), and one to two orders of magnitude larger than for two other main-belt comets (133P/Elst-Pizarro and 176P/LINEAR). Using optical spectroscopy to search for CN emission, we do not detect any conclusive evidence of sublimation products (i.e., gas emission), finding an upper limit CN production rate of QCN < 6× 1023 mol s−1, from which we infer an H2O production rate of QH2O < 1026 mol s−1. Numerical simulations indicate that P/La Sagra is dynamically stable for > 100 Myr, suggesting that it is likely native to its current location and that its composition is likely representative of other objects in the same region of the main belt, though the relatively close proximity of the 13:6 mean-motion resonance with Jupiter and the (3,-2,-1) three-body mean-motion resonance with Jupiter and Saturn mean that dynamical instability on larger timescales cannot be ruled out. Subject headings: comets: general -- minor planets, asteroids [email protected] 1 Institute for Astronomy, University of Hawaii, 2680 Wood- lawn Drive, Honolulu HI 96822, USA 2 Department of Astronomy, Faculty of Mathematics, Univer- sity of Belgrade, Studentski trg 16, 11000 Belgrade, Serbia 3 National Central University, 300 Jhongda Rd, Jhongli City, Taoyuan 320, Taiwan 4 Astronomy Research Centre, Queens University Belfast, Belfast BT7 1NN, United Kingdom 5 Department of Physics, P.O. Box 64, 00014 University of Helsinki, Finland 6 Department of Physics and Astronomy, Johns Hopkins Uni- versity, 3400 North Charles Street, Baltimore, MD 21218, USA 7 Lowell Observatory, 1400 West Mars Hill Road, Flagstaff, AZ 86001, USA 8 Planetary Exploration Group, Space Department, Johns Hopkins University Applied Physics Laboratory, Laurel, MD 20723, USA 9 Northern Arizona University 10 Dipartimento di Matematica, Universit`a di Pisa, Largo Pontecorvo 5, 56127 Pisa, Italy 11 Department of Astronomy, Physics of the Earth and Meteo- rology, Comenius University, Mlynska dolina, 842 48 Bratislava, Slovakia 12 Department of Astrophysical Sciences, Princeton Univer- sity, Princeton, NJ 08544, USA 13 Niels Bohr Institute, University of Copenhagen, Juliane Maries Vej 30, 2100 Copenhagen, Denmark 14 Centre for Star and Planet Formation, Geological Museum, Øster Voldgade 5, 1350 Copenhagen, Denmark 15 SUPA, University of St Andrews, School of Physics & As- tronomy, North Haugh, St Andrews, KY16 9SS, United King- dom 16 Korea Astronomy & Space Science Institute (KASI), 776 Daedukdae-ro, Yuseong-gu, Daejeon, Republic of Korea 17 Space Telescope Science Institute, 3700 San Martin Drive, Baltimore, MD 21218, USA 18 Max-Planck-Institut fur Sonnensystemforschung, 37191 Katlenburg-Lindau, Germany a Hubble Fellow b Michael West Fellow c Royal Society University Research Fellow 2 1. INTRODUCTION Comet P/2010 R2 (La Sagra), hereafter P/La Sagra, was discovered by J. Nomen in images obtained on 2010 September 14.9 (UT) using the 0.45 m La Sagra Observa- tory in southern Spain (Nomen et al. 2010). With a semi- major axis of a = 3.099 AU, eccentricity of e = 0.154, inclination of i = 21.39◦, and a Tisserand parameter (with respect to Jupiter) of TJ = 3.099, it was immedi- ately suspected to be a member of the recently identified class of main-belt comets (MBCs; Hsieh & Jewitt 2006), which are objects that exhibit cometary activity likely due to the sublimation of volatile ice, yet are dynam- ically indistinguishable from main-belt asteroids. Prior to the discovery of P/La Sagra, four MBCs -- 133P/Elst- Pizarro, 176P/LINEAR, 238P/Read, and P/2008 R1 (Garradd) -- were recognized, while a sixth possible MBC (P/2006 VW139) has just recently been identified (Hsieh et al. 2011c). Two other comet-like main-belt ob- jects -- P/2010 A2 (LINEAR) and (596) Scheila -- have also been observed, but as their comet-like morpholo- gies are believed to be due to impact-generated ejecta clouds (Jewitt et al. 2010, 2011; Snodgrass et al. 2010; Bodewits et al. 2011; Yang & Hsieh 2011; Hainaut et al. 2012), and not cometary (i.e., sublimation-driven) dust emission, they are better characterized as disrupted as- teroids. Observationally distinguishing MBCs and disrupted asteroids is not straightforward as both types of objects exhibit visible dust emission in the form of comet-like features like comae and tails. Distinctive morphological features helped betray the true natures of P/2010 A2 and Scheila, however (Hsieh et al. 2012a, and references within). In the case of P/2010 A2, a gap between what appeared to be the nucleus of the "comet" and the dust tail indicated that the observed dust emission was possi- bly an impulsively-generated ejecta cloud that produced by an impact onto P/2010 A2's surface and was then drifting away. This hypothesis was later corroborated by numerical dust modeling and other physical arguments (Jewitt et al. 2010; Snodgrass et al. 2010). In the case of Scheila, no gap was seen between the nucleus and its emitted dust, but instead of forming a single tail, the dust emitted from Scheila appeared to form multiple plumes. Hsieh et al. (2012a) considered various scenarios that could produce multiple dust plumes, finding scenar- ios for sublimation-driven dust emission to be physically implausible as they required multiple active sites, which were argued to be unlikely on Scheila. Instead, a sim- pler scenario in which a single oblique impact caused dust emission in the form of a hollow cone which was then pushed back in the antisolar direction by radiation pressure was favored (Ishiguro et al. 2011a,b). In this scenario, what appeared to be multiple plumes of dust were proposed to simply be optical projection effects (i.e., limb brightening of the hollow dust structure). Early observation reports for P/La Sagra indicated that it exhibited a largely classical cometary morphology, with a visible coma and tail (Nomen et al. 2010), sug- gesting that it could be a true comet (Hsieh et al. 2012a), and was therefore likely to be just the fifth MBC to be discovered at that time. Numerical modeling by Moreno et al. (2011) furthermore showed that the evolution of P/La Sagra's dust tail was consistent with sublimation- driven dust emission, though they did not explicitly test impulsive impact-driven emission scenarios. In this pa- per, we seek to further assess whether P/La Sagra's dust emission is likely to be cometary (sublimation-driven) in nature, or if the object could instead be a disrupted as- teroid. We also conduct observational and dynamical characterization analyses of the object to place it in the context of the small but growing known population of comet-like objects in the main asteroid belt. 2. OBSERVATIONS Imaging observations of P/La Sagra were made in pho- tometric conditions on multiple occasions in 2010 and 2011 (Table 2) by the 2.0 m Faulkes Telescope North (FTN) on Haleakala in Hawaii, the University of Hawaii (UH) 2.2 m telescope, the 8 m Gemini-North Observa- tory, and the 10 m Keck I telescope, all on Mauna Kea in Hawaii, the 1.54 m Danish telescope at La Silla Obser- vatory in Chile, and the 2.5 m Isaac Newton Telescope (INT) at the Roque de los Muchachos Observatory on La Palma in the Canary Islands (Spain). In addition, precovery observations of P/La Sagra were also obtained with the 1.8 m Pan-STARRS 1 (PS1) survey telescope on Haleakala. To ensure proper calibration of PS1 data, we also obtained calibration observations using the Lowell Observatory 31(cid:48)(cid:48) telescope in Flagstaff, Arizona. PS1 is a wide-field synoptic survey telescope that em- ploys a 1.4 gigapixel camera consisting of a mosaic of 60 orthogonal transfer arrays (OTAs), each consisting of 64 590 × 598 pixel CCDs, giving a total field of view 3.2 degrees in diameter (image scale of 0.(cid:48)(cid:48)26 pixel−1). PS1 observations were made using a Sloan Digital Sky Survey (SDSS) r(cid:48)-band-like filter designated rP1 (Stubbs et al. 2010). Gemini observations (obtained as part of program GN-2011B-Q-17) were made using the Gemini Multi-Object Spectrograph (GMOS; Hook et al. 2004) in imaging mode. GMOS employs three 2048 × 2048 EEV CCDs behind a SDSS-like r(cid:48)-band filter, and has an image scale of 0.(cid:48)(cid:48)1456 pixel−1 when using 2 × 2 bin- ning as we did for our observations of P/La Sagra. Keck observations were made using the Low Resolution Imag- ing Spectrometer (LRIS; Oke et al. 1995) in both imag- ing and spectroscopic mode. LRIS employs a Tektronix 2048 × 2048 CCD with an image scale of 0.(cid:48)(cid:48)135 pixel−1 and standard Kron-Cousins B-band and R-band filters. The Danish 1.54 m telescope is currently exclusively used by the Microlensing Network for the Detection of Small Terrestrial Exoplanets (MiNDSTEp) consortium. Obser- vations on the Danish 1.54 m were made using the Dan- ish Faint Object Spectrograph and Camera (DFOSC) focal-reducing imager behind a standard Kron-Cousins R-band filter. The DFOSC imager has pixel dimensions of 2102 × 2102 and a pixel scale of 0.(cid:48)(cid:48)39 pixel−1. FTN observations were made using a 1024 × 1024 e2v CCD with an image scale of 0.(cid:48)(cid:48)2785 pixel−1 behind a Bessell R-band filter. Except for observations with PS1, FTN, and the INT, all other observations were conducted using non-sidereal tracking at the apparent rate and direction of motion of P/La Sagra on the sky. Sidereal tracking was used by PS1 because this was the standard mode of operation for survey operations. FTN observations were obtained using its web-based remote control interface, through which non-sidereal tracking is not possible. The INT is normally capable of tracking solar system objects non- sidereally, but was unable to do so during our observing run due to mechanical problems. We performed standard image preparation (bias sub- traction and flat-field reduction) for UH 2.2 m, Gem- ini, Keck, Danish 1.54 m, and INT data using Image Reduction and Analysis Facility (IRAF) software. For Keck data, flat fields were constructed from images of the illuminated interior of the Keck I dome, while for the other telescopes, dithered images of the twilight sky were used to construct flat fields. PS1 data were reduced using the system's Image Processing Pipeline (IPP; Mag- nier 2006) and then calibrated using field stars imaged later by the Lowell 31(cid:48)(cid:48) telescope. Calibration of Gem- ini data was performed using field star magnitudes pro- vided by the Sloan Digital Sky Survey (SDSS; York et al. 2000). Photometry of Landolt (1992) standard stars and field stars was performed for all data using IRAF and obtained by measuring net fluxes within circular aper- tures, with background sampled from surrounding cir- cular annuli. Conversion of magnitudes measured from PS1 and Gemini data obtained using SDSS-like r(cid:48)-band filters to their Kron-Cousins R-band equivalents was ac- complished using the transformation equations derived by R. Lupton and made available on the SDSS website (http://www.sdss.org/). Comet photometry was per- formed using both rectangular and circular apertures as described below, where to avoid dust contamination from the comet itself, background sky statistics are measured manually in regions of blank sky near, but not adjacent, to the object. Several (5 to 10) field stars in the comet images were also measured to correct for any extinction variation (typically negligible) during each night. In addition to imaging observations, we also secured optical spectra of P/La Sagra on 2010 October 05 with LRIS on Keck. We adopted LRIS's 1.(cid:48)(cid:48)0-wide long-slit mask, the 400/3400 grism, and the 460 dichroic on the blue side, giving a dispersion of 1.08 A pixel−1 and a spectral resolution of approximately 7 A. A total of 3600 s of data were obtained from four integrations on the comet while it was at an airmass of ∼1.1. Unfortunately, when taking calibration images, our observations were inter- rupted due to fog and we were forced to close the dome for over three hours. As such, we were only able to record a spectrum of a flux standard star and failed to obtain any spectra of nearby G-type standard stars, complicat- ing our reduction and analysis (Section 3.3.1). Data re- duction was accomplished using IRAF. 3. RESULTS & ANALYSIS 3.1. Morphological Analysis For each night during which observations were ob- tained, we construct composite images by aligning each night's set of images on the object's photocenter and adding them together. We see that P/La Sagra is clearly morphologically cometary between August 2010 and February 2011 (Figure 2), with a dust tail extending up to ∼ 30(cid:48)(cid:48) (∼ 4 × 104 km at the geocentric distance of the object at the time) from the nucleus's photocenter. The projected orientation of this dust tail on the plane of the sky is seen to rotate counterclockwise throughout this period, lagging slightly behind the similarly rotating antisolar vector (as projected on the sky). From August 3 2010 through November 2010, we observe a single dust tail aligned with the projection of the antisolar vector in the sky, while between December 2010 and February 2011, an apparent dust trail aligned with the projection of the heliocentric velocity vector (i.e., the plane of the object's orbit) appears in addition to the antisolar dust tail. By August 2011 (Figure 3), active dust emission appears to have ceased, though a residual dust trail (pre- sumably consisting of large dust particles ejected during earlier activity) aligned with the projection of the ob- ject's heliocentric velocity vector in the sky is observed extending at least ∼1 arcmin (∼ 1.5 × 105 km at the geocentric distance of the object at the time) from the nucleus. This dust trail peaks in brightness well away (∼ 20(cid:48)(cid:48)) from the nucleus. In the formulation of dust dynamics outlined by Fin- son & Probstein (1968), the trajectory of an emitted dust particle depends on the orbit of the emitting comet, the particle's ejection time, its ejection velocity, and the ra- tio, β, of solar radiation pressure, Frad, to solar grav- ity, Fgrav, on the particle. In cases where β is small (i.e., for massive particles), radiation pressure is negli- gible, causing particles follow orbits very close that of the comet. Over time, these large grains will spread out along the comet's orbit to form a potentially observable trail. In cases where β is large (i.e., for small particles), radiation pressure is dominant over solar gravity, giv- ing particles nearly directly antisolar motion in the rest frame of the comet. In the case of P/La Sagra, we ob- serve both limiting cases: a tail consistently oriented in the antisolar direction (likely composed of small, high- β particles) from August 2010 through February 2011 (Figure 2), and a trail confined to the orbit plane (likely composed of large, low-β particles) from December 2010 through August 2011 (Figures 2 and 3). This assessment is in agreement with the modeling results obtained by Moreno et al. (2011) who found dust grain sizes ranging from ∼5 µm to ∼1 cm. Similar dust structures compris- ing widely-varying particle sizes have also been observed for comets 2P/Encke (Reach et al. 2000; Lisse et al. 2004) and C/Austin 1990 V (Lisse et al. 1998). Due to the ease by which they are accelerated by radia- tion pressure, the small particles comprising an antisolar tail are expected to dissipate quickly (on timescales of several weeks at the distance of the main belt; cf. Hsieh et al. 2004) unless they are continually replenished. No such separation of the antisolar tail from the nucleus is observed at any time from August 2010 through Febru- ary 2011, a span of 5.5 months, strongly suggesting that continuous dust production was ongoing over this time period, consistent with cometary activity. We note that the persistence of the orbit-aligned dust trail until Au- gust 2011 does not mean that dust production contin- ued up until this time as well. In this latter case, the longevity of the dust trail is instead likely due to the large sizes and therefore slow dissipation rates of the particles in the trail (cf. P/2010 A2; Jewitt et al. 2010; Snodgrass et al. 2010), not continuous replenishment. Qualitatively, we find that P/La Sagra's morphological appearance and evolution strongly suggest that its dust emission is cometary in nature. We caution however that more detailed dust modeling (similar that performed by Moreno et al. 2011, but also considering impulsive emis- sion scenarios) will be required for quantitative confir- 4 mation of the conclusions reached here. 3.2. Photometric Analysis 3.2.1. Nucleus Size, Shape, and Color Vigorous activity, such as that exhibited by P/La Sagra during the majority of our 2010 and 2011 observa- tions, normally interferes with the measurement of phys- ical properties for cometary nuclei (e.g., size, rotation pe- riod, and phase darkening behavior) due to obscuration caused by coma when such activity is present. As such, in order to accurately measure the physical properties of comet nuclei, it is typically necessary to observe them once activity has mostly or entirely ceased, as we did with P/La Sagra in August 2011 using the Keck I and Gem- ini telescopes (Table 2). We measure a mean apparent R-band magnitude (averaged in flux space and weighted by nominal uncertainties) of m = 23.9 ± 0.1 mag, with a magnitude variation between the two nights of ∆m = 0.7 ± 0.1 mag (mavg = 24.1 ± 0.1 mag on 2011 August 26, and mavg = 23.4 ± 0.1 mag on 2011 Aug 31). Pro- vided that the photometric variation between these two nights is real (intra-night variations are on the order of ∼0.2 mag, with observations on each night only spanning 30 min or less) and due to rotation, the minimum ratio of the axes of the body as projected in the plane of the sky is then given by aN /bN = 100.4∆m (1) giving us a lower limit (due to unknown projection ef- fects) to the axis ratio for P/La Sagra of aN /bN = 1.9± 0.3. Given the extremely limited data set on which this calculation is based, this axis ratio should of course be considered as a preliminary result only. Better con- straints on the shape and structure of the nucleus of P/La Sagra will require additional observational sampling of the object's lightcurve (likely requiring large telescopes, given the small size of the object), ideally at multiple observing geometries. Using the range of slope parameter (G) values in the IAU H, G photometric system computed for fellow MBCs 133P, 176P, 238P, and P/Garradd (−0.03 < GR < 0.26; Hsieh et al. 2009a, 2011b; Maclennan & Hsieh 2012), we find an estimated absolute magnitude of HR = 17.9 ± 0.2 mag. Adopting a R-band albedo of pR = 0.05 (as measured for 133P and 176P; Hsieh et al. 2009a), we then find an approximate effective nucleus radius of rN = 0.7± 0.1 km, making it comparable in size to 238P, about three times larger than P/Garradd, and about three times smaller than 133P and 176P (Table 1). Using the approximate axis ratio derived above and assuming a triaxial ellipsoidal shape, we find semimajor and semimi- nor axes for the nucleus of aN = 1.0 km and bN = 0.5 km, respectively. We note that photometric variations on the scale of ∆m = 0.7 mag were not seen at any time while P/La Sagra was active (August 2010 through February 2011) for any observations that were comparably closely spaced in time as the August 2011 observations analyzed above. For example, observations on four separate occasions be- tween 2010 Sep 08 and Sep 29 exhibited a maximum variation of just ∼ 0.1 mag (Table 3). This lack of cor- roborating detections of significant photometric variation during that period does not invalidate our detection of a ∆m = 0.7 mag variation in August 2011, however, since rotational variations for a nucleus obscured by a steady- state coma (assumed to be optically thin) will be effec- tively damped by that coma. The extent of this damping can be determined using (cid:18) FN,max + Fd (cid:19) ∆mobs = 2.5 log Fobs,max Fobs,min = 2.5 log FN,min + Fd (2) where ∆mobs is the observed photometric range, Fobs,max and Fobs,min are the maximum and minimum fluxes ob- served for the active comet, FN,max and FN,min are the maximum and minimum fluxes for which the nucleus is responsible, and the flux due to dust, Fd, is assumed to be constant (Hsieh et al. 2011a). Assuming that the nucleus and dust grains have sim- ilar albedos, flux terms in Equation 2 can be consid- ered equivalent to the corresponding scattering surface areas of the comet components being considered. We find in the following section (Section 3.2.2) that the ratio of scattering surface areas of dust and the nucleus dur- ing the four September observations discussed above was Ad/AN ∼ 30 (Table 3). Thus, using AN,max = πaN rN , AN,min = πbN rN , and Ad = πr2 N , we find ∆mobs = 0.025 mag. In other words, during this period, even the minor ∼ 0.1 mag photometric variation we observed for the object is unlikely to be due to the rotation of the nucleus. Instead, we surmise that the observed variation is more likely due to fluctuations in the seeing conditions between the different nights, causing fluctuations in the amount of coma contained within the fixed 5.(cid:48)(cid:48)0 photom- etry apertures we used to measure the near-nucleus flux of the comet in all of our images. We also conclude that our non-detection of significant photometric variations in data obtained when P/La Sagra was active is not in- consistent with our apparent detection of a significant photometric variation while the object was inactive. Multi-filter observations using LRIS on Keck I also per- mitted us to measure a B − R = 0.9 ± 0.1 mag color for P/La Sagra's inactive nucleus, which within our uncer- tainties, is consistent with the color of the Sun (B − R = 1.03 mag; Hartmann et al. 1982, 1990). The colors we measure for P/La Sagra's dust coma while it was active are also approximately solar (B − R = 1.02 ± 0.05 mag on 2010 October 5 and B − R = 0.94 ± 0.05 mag on 2011 Feb 3), in good agreement with colors measured for the other MBCs (Jewitt et al. 2009; Hsieh et al. 2009b, 2010b, 2011a). 3.2.2. Activity Strength Using the nucleus size computed above in Section 3.2.1, the expected apparent nucleus magnitude for P/La Sagra at the time of our first observations in August 2010 is mR = 22.2 mag, not mR = 18.49 mag as measured (Ta- ble 3). This discrepancy indicates that there was signif- icant near-nucleus dust contamination even during our earliest observations of P/La Sagra. Assuming that the phase behavior of P/La Sagra's dust coma is similar to that of the nuclei of other MBCs whose phase functions have been measured (cf. Section 3.2.1; since we lack other meaningful constraints), we calculate the comet's corre- sponding absolute magnitudes, mtot(1, 1, 0), at heliocen- tric and geocentric distances of R = ∆ = 1 AU and phase angles of α = 0◦ for each set of photometric mea- surements (using photometry apertures with radii of 5.(cid:48)(cid:48)0) of the nucleus (Table 3). By comparing these absolute magnitudes to our estimated absolute magnitude for the nucleus, we then compute the ratio of the scattering sur- face area of dust in the coma, Ad, to that of the nucleus, AN , using Ad AN = 1 − 100.4(mtot(1,1,0)−HR) 100.4(mtot(1,1,0)−HR) (3) Then, assuming AN ∼ πr2 N and a = 10 µm-radius grains with bulk densities of ρ = 1300 kg m−3 (cf. Hsieh et al. 2004; Jutzi et al. 2008), we estimate the total dust mass, Md, in the coma using (cid:18) Ad (cid:19) Md = ρa · AN 4 3 AN (4) We list the values we compute for Ad/AN and Md using the above procedures in Table 3. Comparing the min- imum and maximum values of Md during our observa- tions, we find that P/La Sagra's coma increases in mass Md ∼ 0.1 kg s−1 (likely precise at an average net rate of to an order of magnitude, at best) between August and December 2010. This net increase in total dust mass represents the sum of the positive contribution of new dust production from the nucleus and the negative con- tribution of dust dissipation as individual grains disperse beyond the photometry aperture. As such, it represents a lower limit to the true dust production rate of the nu- cleus. In addition to performing nucleus photometry, we also measure the total flux from the comet in our compos- ite images from each night. We do so by using rectan- gular photometry apertures enclosing the entire visible dust cloud and oriented to avoid field star contamina- tion. Background sky levels were then measured from nearby areas of blank sky and subtracted to obtain net fluxes, which were then calibrated using standard stars to obtain absolute photometry (mtot). Unfortunately, the faintness of the residual dust trail observed on 2011 Au- gust 26 and 31, and extensive contamination by nearby bright field stars means that we are unable to perform these measurements for data obtained on those dates. Repeating the mass production rate analysis described above, we find that P/La Sagra's coma and tail increase in mass at about twice the rate measured for the near- Md ∼ 0.2 kg s−1. nucleus coma, or an average net rate of Since this analysis does not rely on a fixed aperture size, this net mass loss rate is closer to the true dust pro- duction rate of the nucleus than that computed above, though still represents a lower limit due to the unknown dissipation rate of the dust as individual grains disperse from the coma and tail and become undetectable above the sky background. Comparing with other MBCs, we nonetheless find that P/La Sagra has a dust production rate comparable to or larger than that of 238P (Hsieh et al. 2009b), and at least an order of magnitude larger than those of 133P and 176P (Hsieh et al. 2004, 2011a). For reference, we also use our measurements of the near-nucleus flux of the comet (using 5.(cid:48)(cid:48)0 photometry apertures) in each set of observations to estimate the dust contribution as parameterized by Af ρ (A'Hearn et al. 1984). These results are also shown in Table 3, and again 5 indicate significantly stronger activity for P/La Sagra as compared to 133P, for which Af ρ was also systematically tabulated (Hsieh et al. 2010b). We also attempt to reproduce our observed photom- etry using a sublimation model similar to that used by (Hsieh et al. 2011b) to analyze the activity of 238P. This effort was hampered, however, by our lack of observa- tional constraints on the start time of activity and the rate of its eventual decline. We do find that steadily in- creasing activity during the post-perihelion portion of P/La Sagra's orbit cannot be modeled by surface ice sublimation, indicating that the ice must instead reside at some depth below the surface of the nucleus, similar to what was found for 238P (Hsieh et al. 2011b). Un- fortunately, the numerous free parameters that can be modified in our model mean that this finding is the only meaningful result that can be derived at this time. Bet- ter sampling of the activity over a longer time baseline will be required to gain additional insight into the na- ture of P/La Sagra's activity via this method, and as such, should be considered an observational priority dur- ing its next expected period of activity (likely beginning in mid-2015). We note that if P/La Sagra exhibits activity over a period of ∼ 1 year during each of its perihelion passages (similar to 133P; Hsieh et al. 2004), assuming an average Md = 0.2 kg s−1, a total mass loss mass loss rate of of ∼ 107 kg per orbit would be expected. Assuming a nucleus size of rN = 0.7 km (Section 3.2.1) and ρ = 1300 kg m−3, such a mass loss rate should consume the entire nucleus over the next ∼ 105 orbits, or < 106 yr. However, activity in MBCs (even over just a portion of their orbits) is expected to be transient (cf. Hsieh et al. 2004), perhaps amounting to just 1% of an object's total lifetime (Hsieh 2009). As such, we expect a timescale for nucleus disintegration due to cometary activity closer to ∼ 108 yr (Hsieh 2009), or approximately comparable to the ∼ 108-109 yr collisional lifetime expected for a main- belt asteroid the size of P/La Sagra's nucleus (Bottke et al. 2005). 3.2.3. Comparison to Other MBCs To further understand P/La Sagra's place among the population of known MBCs, we plot Ad/AN for the first five known MBCs as functions of true anomaly, ν (Figure 5a). We note a few interesting features of this plot. First, we find quantitative confirmation that for their nucleus sizes, 238P and P/Garradd exhibit far more dust when active than 133P and 176P (also see Table 1). This fact has of course always been qualita- tively apparent. The observed morphologies of 238P and P/Garradd, when active, resemble those of highly active comets (Hsieh et al. 2009b, 2011b; Jewitt et al. 2010), whereas the nuclei of 133P and 176P dominate the sur- face brightness profiles of those objects, even at the peaks of their active periods. Only comparatively faint dust tails and small amounts of photometrically-detected un- resolved near-nucleus dust have attested to the cometary natures of these latter MBCs (Hsieh et al. 2004, 2010b, 2011a). As for P/La Sagra, we find that it exceeds even 238P and P/Garradd in the amount of dust it produces for the size of its nucleus, producing ∼ 3 times as much dust relative to its nucleus size at its peak as 238P and 6 P/Garradd produced at their peaks, which in turn were ∼ 50 times as productive relative to their nucleus sizes as 133P and 176P. We caution that given the different geocentric dis- tances at which the observations in question were made, the near-nucleus photometry performed for each MBC measures dust contained within different physical dis- tances from the nucleus. As such, considering that other physical characteristics of the dust emission (e.g., grain sizes, bulk grain densities, and ejection velocities) for each MBC are unknown and likely also vary from ob- ject to object, these comparisons are best interpreted in terms of orders of magnitude. We therefore find 238P, P/Garradd, and P/La Sagra to have similar levels of activity which are 1-2 orders of magnitude higher than those of 133P and 176P, implying similar ratios of frac- tional active surface areas for these objects. To quantify the total amount of ejected dust exhib- ited by each comet, we also multiply the Ad/AN values for each comet by estimates of AN for each object (de- rived from nucleus radii listed in Table 1) to obtain ab- solute values for Ad. We plot the results as functions of true anomaly in Figure 5b. Interestingly, we find that despite the high Ad/AN values measured for 238P and P/Garradd, the total absolute amount of dust in those comets is approximately comparable to the amounts of dust present for 133P and 176P, which visually appear far less active in comparison. P/La Sagra is seen to pro- duce roughly an order of magnitude more dust in abso- lute terms than any of the other MBCs (assuming similar particle size distributions). We therefore find the interesting result that within an order of magnitude, all five MBCs considered here ex- hibit similar total amounts of visible dust in their comae and tails, despite the wide variation in their observed morphologies. Determining whether this finding is phys- ically significant will require similar dust mass measure- ments for newly-discovered MBCs to investigate whether this trend is maintained as the sample of known objects grows. We suggest that future research investigating the validity and implications of this possible trend may pro- vide valuable insights into the origin and evolution of MBC activity. 3.3. Spectroscopic Analysis 3.3.1. Data Reduction In the optical region of the electromagnetic spectrum, the emission band of the CN radical at 3880 A is among the strongest among cometary gaseous species and there- fore is considered to be the most sensitive probe for de- tecting gas resulting from sublimating ice. We search for evidence of the presence of the CN band in a one dimensional spectrum extracted from the spectral image using a 1.(cid:48)(cid:48)0 × 7.(cid:48)(cid:48)5 rectangular aperture centered on the continuum. We use a relatively wide extraction aper- ture to incorporate as much light as possible from the surrounding coma. The sky background is reconstructed and subtracted using flanking regions extending from 9(cid:48)(cid:48) to 15(cid:48)(cid:48) from the nucleus. In Figure 6, we present the spectrum of P/La Sagra in the relevant spectral region from ∼ 3700 A to ∼ 4000 A. The continuum of the comet consists of scattered sun- light from dust particles in the inner coma and shows several absorption features in the Solar spectrum, such as the prominent Ca H and K absorption lines at 3933 A and 3966 A (Figure 6). The shaded spectral region from 3830 A to 3905 A indicates where we expect to detect CN emission if any is present (Cochran et al. 1992). In order to correct for solar spectral features, the spec- trum of a G2V solar analog star is needed. As discussed in Section 2, we were not able to observe a G2V star at the time of our P/La Sagra observations due to poor weather conditions. We instead adopt a flux-calibrated spectrum of G2V star HD 91163 which was obtained us- ing the same instrument but was recorded a few months later on UT 2010 December 17 (Hsieh et al. 2012a). This spectrum was observed using the 600/4000 grism and thus has a higher spectral resolution than our P/La Sagra data, however, so the additional step of binning the HD 91163 spectrum to the same spectral resolution as our P/La Sagra spectrum is required. The scaled spec- trum of HD 91163 is shown as the blue dashed line in Figure 6. We note that the spectra of P/La Sagra and HD 91163 are quite similar, particularly in the shaded region. However, minor discrepancies are observed at wavelengths λ < 3800 A and λ > 3980 A. These discrep- ancies may be due to the time and airmass differences between the comet observations and standard star ob- servations. We find no evidence of CN emission in the shaded wavelength region. 3.3.2. Upper Limits on Gas Production Rates The continuum-removed spectrum of P/La Sagra is shown in Figure 7. In principle, this spectrum shows emission from atoms and molecules in the inner coma. We measure the standard errors in the residual spectrum in three wavelength regions 70 A in width: 3760 A to 3830 A, 3830 A to 3900 A, and 3900 A to 3970 A. The standard errors on either side of the expected CN band are 5.1 × 10−18 and 2.8 × 10−18 erg cm−2 s−1 A , re- spectively, shown as red error bars in Figure 7. The stan- dard error in the region of the expected CN band is 4.2× 10−18 erg cm−2 s−1 A . We choose the largest standard error of 5.1 × 10−18 erg cm−2 s−1 A as a conservative estimate of our observational uncertainty in the region of the expected CN emission band. We therefore estimate a 3-σ upper limit of 1.5 × 10−17 erg cm−2 s−1 A −1 −1 −1 −1 . We calculate the integrated CN band flux by summing the estimated emission flux in the shaded CN region, obtaining fCN = 3.7 × 10−16 erg cm−2 s−1. We then convert the integrated flux to the total number of CN molecules using LCN = 4π∆2fCN NCN = LCN g(R) (5) (6) where NCN is the number of CN molecules, ∆ and R are the geocentric and heliocentric distances, re- spectively, and g(R) is the resonance fluorescence ef- ficiency, which describes the number of photons scat- tered per second per radical for an optically thin coma, in erg s−1 molecule−1 at a heliocentric distance of R. At the time of our observations, P/La Sagra had a ra- dial velocity of for which g(R = R = 1.1 km s−1, 1 AU) = 2.86 × 10−13 (Schleicher 2010). Substituting fCN = 3.7 × 10−16 erg cm−2 s−1, R = 2.66 AU, and ∆ = 1.82 AU, we find an upper limit of N = 8.3 × 1025 CN molecules in the extraction aperture. A simple Haser model (Haser 1957) was used to de- rive the CN production rate based on the number of CN molecules. The details of the model are described in Hsieh et al. (2012a). We find an upper limit to the CN production rate of QCN < 6 × 1023 mol s−1. Taking average ratios measured for previously observed comets (log[QCN/QOH] = −2.5; QOH/QH2O = 90%) (A'Hearn et al. 1995), we estimate an upper limit on the water pro- duction rate of QH2O < 1026 mol s−1. Incorporating the Af ρ value calculated from data taken at the same time as these spectra (Section 3.2.2; Table 3), we find a lower- limit dust-to-gas ratio of log[Af ρ/QCN] (cid:38) −22, simi- lar to values measured for other Jupiter-family comets (A'Hearn et al. 1995). As this result is only an lower limit, however, it is not possible to ascertain from these data the degree to which P/La Sagra is depleted in volatiles relative to other classical comets from its ap- parently long residence time in the inner solar system (Section 3.4.1). 3.3.3. Estimation of Effective Active Area We can estimate the effective active area of the nucleus of P/La Sagra by comparing the observed mass loss rate with theoretical rates obtained from applying a simple thermal model. To calculate the sublimation rate of sur- face ice, we solve the energy balance equation for the equilibrium temperature on the object's surface: S(cid:12)(1 − p) r2 J(θ) = σT 4 e + Lψ(Te) (7) where S(cid:12) = 1360 W s−1 is the solar constant, p = 0.05 is the albedo, R = 2.66 AU, J(θ) is a function of the so- lar zenith angle θ which describes the ratio between the absorbing and radiating area,  = 0.9 is the emissivity, σ is the Stefan-Boltzmann constant, L = 2.68× 106 J kg−1 is the latent heat and Te is the effective surface temper- ature. The specific sublimation rate, ψ (in kg m−2 s−1), of water is given by (cid:114) m 2πkTe ψ = P (Te) (8) where m is the mass of a water molecule, k is Boltz- mann's constant, and P (Te) is the saturated vapor pres- sure (Fanale & Salvail 1984). In the fast rotator approximation (which yields the coldest surface temperature), the nucleus is considered isothermal and J = (cid:104)cos θ(cid:105) = 1/4. The hottest surface temperatures are derived in the sub-solar point (i.e., slow rotator) approximation, where the absorbing and radiat- ing areas are identical and therefore J = 1. We ob- tain sublimation rates of 1.3 × 10−6 kg m−2 s−1 and 6.1 × 10−5 kg m−2 s−1 for the cold and hot scenar- ios, respectively. Assuming an average mass loss rate Md ∼ 0.2 kg s−1 (Section 3.2.2), we estimate active of areas of 1.5 × 105 km2 and 3.3 × 103 km2, correspond- ing to total active surface fractions of ∼3% and ∼0.05% (likely precise to an order of magnitude, at best), for the cold and hot cases, assuming dust-to-gas mass ratios of 7 unity. These derived active surface fractions are compa- rable to those of other Jupiter-family comets measured by A'Hearn et al. (1995). For comparison, both 133P and 176P are estimated to have active areas of ∼ 102 - 103 m2 (Hsieh et al. 2004; Hsieh 2009), corresponding to active surface fractions of 0.0002 - 0.002%, several orders of magnitude smaller than for P/La Sagra. 3.4. Dynamical Analysis 3.4.1. Stability Analysis To better understand the likely origin of P/La Sagra, we also consider its dynamical properties. Jewitt et al. (2009) raised the possibility that MBC P/Garradd could have originated in a different part of the asteroid belt from where we see it today. As such, an important issue to consider is whether P/La Sagra is dynamically stable at its present location, or whether it is possible that it may have also originated elsewhere. To assess P/La Sagra's dynamical stability, we ran- domly generate three sets of 100 test particles, each set with Gaussian distributions in orbital element space, cen- tered on P/La Sagra's JPL-tabulated osculating orbital elements, where each set is characterized by a σ value equal to the JPL-tabulated uncertainties (as of 2011 Jan- uary 1) for each orbital element. To investigate the sta- bility of the region (in orbital element space) surrounding P/La Sagra, we also generate three additional sets of 100 Gaussian-distributed test particles each, where each set is again centered on P/La Sagra's osculating orbital el- ements but is characterized by a σ value equal to 100× the JPL-tabulated uncertainties. We then use the N- body integration package, Mercury (Chambers 1999), to integrate the orbits of each set of test particles forward in time for 100 Myr (limitations on available comput- ing resources unfortunately prevent us from conducting these simulations over significantly longer time periods, e.g., 1-2 Gyr, in a reasonable amount of time), where we treat the eight major planets as massive particles and all test particles are considered to be massless. Performing these integrations for all six sets of test particles, we find that 7% of the objects in the three 1- σ test particle sets are ejected (i.e., reach a heliocentric distance of > 50 AU) from the asteroid belt over the course of 100 Myrs (Figure 8), and 11% of the objects in the three 100-σ test particle sets are ejected (Figure 9). P/La Sagra itself is found to be stable over the 100 Myr test period, consistent with its Lyapunov time (Tlyap), estimated by us to be ∼200 kyr (computed using the pro- cedure described in Tsiganis et al. 2003), where a body is generally considered to be stable for Tlyap > 100 kyr. The 1-σ test particles that are ejected are not localized to par- ticular regions of orbital element space near P/La Sagra, suggesting that their ejection may be a consequence of stochastic encounters with a weak mean-motion reso- nance (or multiple resonances), rather than the effect of a single dominant resonance. Nearby resonances in- clude the "moderate-order" 13:6 mean-motion resonance with Jupiter at a = 3.1063 AU and the (3,-2,-1) three- body mean-motion resonance (with Jupiter and Saturn) at a = 3.0790 AU (Nesvorn´y & Morbidelli 1998). Both resonances are more than 1-σ away in a, given the uncer- tainties for P/La Sagra's osculating elements, but could conceivably affect the object as its osculating elements 8 fluctuate over time (Figure 10). In our 100-σ test particle simulations, the effect of the (3,-2,-1) three-body resonance is clear, with ∼70% of the ejected test particles in those simulations having initial semimajor axes within 0.01 AU of the resonance. The effect of the 13:6 resonance is comparatively less pronounced, with only ∼20% of the ejected particles in these simulations starting within 0.01 AU of the reso- nance, despite the fact that this resonance is closer to P/La Sagra than the (3,-2,-1) resonance and therefore lies in a more highly populated region of our initial test particle distribution. P/La Sagra itself has an initial os- culating semimajor axis 0.06 AU of the 13:6 resonance, and even occasionally crosses the resonance during our simulations (Figure 10), yet is found to be stable over the 100 Myr test period considered here, suggesting that this resonance only weakly affects objects in this region of orbital element space. We conclude from this analysis that the region in or- bital element space occupied by P/La Sagra is largely stable, implying that the object is likely to be native to its present-day location. As such, its physical prop- erties may be reflective of other objects in this region of the asteroid belt. However, between the nearby 13:6 mean-motion resonance with Jupiter and (3,-2,-1) three- body mean-motion resonance with Jupiter and Saturn, enough dynamical instability is imparted to the region that a non-negligible fraction of simulated test particles escape in less than 100 Myr. Due to the non-negligible instability of this region, we therefore cannot definitively conclude that P/La Sagra formed in situ (and did not dynamically evolve onto its present-day orbit from else- where in the main belt, or even elsewhere in the so- lar system), particularly over timescales longer than the 100 Myr test period considered in this analysis. We further note that our analysis does not consider non- gravitational effects such as the Yarkovsky effect or the effects of asymmetric mass loss such as that associated with cometary jets. As such, in future investigations of this object, particularly those which pertain to the com- parison of its physical properties to those of other MBCs, the possibility that P/La Sagra may not be native to its current location cannot be discounted. 3.4.2. Family Search Having investigated the dynamical stability of P/La Sagra and its surroundings, we seek to further character- ize the object's dynamical environment by searching for evidence of an associated asteroid family, i.e., a grouping of dynamically linked asteroids likely to be the result of either the catastrophic fragmentation of a larger parent asteroid (Hirayama 1918), or a significant cratering event on a still-existing asteroid (e.g., Vesta; Asphaug 1997). Such a search is particularly crucial for this object be- cause of studies suggesting that MBCs may preferentially be found among families (Hsieh et al. 2004; Haghighipour 2009), and possibly specifically among extremely young families (Nesvorn´y et al. 2008; Hsieh 2009). The lat- ter hypothesis is based on the facts that MBC 133P ap- pears to be a member of the < 10-Myr-old Beagle family (Nesvorn´y et al. 2008), and collisional devolatilization of near-surface ice of km-scale MBCs (like 133P and 176P) is expected to occur at rates incompatible with the ex- istence of present-day activity, unless those MBCs are recently-produced fragments of larger asteroids with sig- nificant interior ice reservoirs (Hsieh 2009). Due to P/La Sagra's high inclination, we cannot search for an associated dynamical asteroid family using analytically-determined proper orbital elements (Milani & Knezevi´c 1994), as can be done for objects at lower inclinations. The complication stems from the fact that analytically-computed proper elements are significantly less accurate for objects with high eccentricities or in- clinations. The development of numerical methods for computing so-called synthetic proper elements for char- acterizing the long-term behavior of high-inclination and high-eccentricity objects resolved this problem (Knezevi´c & Milani 2000), an advance that was employed by Gil- Hutton (2006) and Novakovi´c et al. (2011) to conduct systematic searches for families among high-inclination asteroids. For our analysis of P/La Sagra, we first compute the object's own synthetic proper elements -- proper semi- major axis (ap), proper eccentricity (ep), and proper inclination (ip) -- using the procedure described in Knezevi´c & Milani (2000) and osculating orbital elements from the JPL Small-Body Database Browser. We list these results in Table 4. We then plot these newly com- puted synthetic proper elements in ap − ep and ap − ip space along with the synthetic elements of all of the fam- ilies (as well as smaller or less-significant groupings des- ignated as clumps or clusters) identified by Gil-Hutton (2006) and Novakovi´c et al. (2011) to check for plausible close associations of P/La Sagra with any of these previ- ously identified groupings (Figure 11). We find no such associations. Finding no link between P/La Sagra and any currently known asteroid families, we turn to our own search for dynamically associated asteroids to determine whether a clustering specific to P/La Sagra may have missed simply because this newly discovered object was not considered in previous family searches. To do so, we add our com- puted proper elements for P/La Sagra to an updated list of synthetic proper elements for high-inclination aster- oids (Novakovi´c et al. 2011), and then employ the Hier- archical Clustering Method (HCM; Zappal`a et al. 1990, 1994) to search for dynamically associated asteroids. Re- sults from this analysis are plotted in Figure 12 where we show the number of asteroids associated with P/La Sagra as a function of cutoff "distance" (which, despite its name, has units of velocity). We find just six objects in P/La Sagra's dynamical vicinity (including P/La Sagra itself) within a cutoff dis- tance of 120 m s−1 (Table 4). Confirming our search (described above) for associations of P/La Sagra with previously identified groupings, none of these objects be- long to any of the families, clumps, or clusters found by Gil-Hutton (2006) or Novakovi´c et al. (2011). According to the criteria used by Novakovi´c et al. for the outer belt region, this group of objects is neither compact nor large enough to be classified as a family or clump. At best, it can be classified as a cluster, defined by Novakovi´c et al. (2011) as a compact group of objects that could share a common physical origin, but does not formally satisfy criteria for being a family or clump, usually because the number of members that it contains is too small. To obtain a quantitative assessment of the likelihood that members of the P/La Sagra group share a common physical origin, we use a method employed by Nesvorn´y & Vokrouhlick´y (2006) to estimate the probability that a certain cluster of objects is the result of random chance. Confining our analysis to the high-inclination region of the outer main belt (2.825 < ap < 3.278 AU, ep < 0.4, 0.3 < sin(ip) < 0.55), we generate 10,000 random distri- butions of 9569 synthetic objects (the same number of ob- jects currently known to populate this region) in proper orbital element space. We then perform HCM analy- sis on these synthetic populations to search for possible groupings with at least 6 objects within cutoff distances of 95 m s−1. The results of this analysis show that there is a significant probability (∼ 30%) that the P/La Sagra cluster could in fact be due to chance. While these re- sults still leave a ∼ 70% chance that this grouping is not random and therefore could have a physical origin, more information is certainly needed to establish whether or not this clustering can be considered real. We compute Lyapunov times for members of the P/La Sagra group (Table 4), finding the orbits of all six bodies to be stable (Tlyap >100 kyr). This fact provides an opportunity to apply the backward integration method (BIM; Nesvorn´y et al. 2002) to this group. BIM is a means for estimating the age of a <20-Myr-old family that is based on the premise that immediately after the disruption of a parent body, the orbits of the resulting fragments must have been nearly the same, and has been successfully applied to estimate the ages of the Karin cluster and Veritas family (Nesvorn´y et al. 2003). In the case of the group of asteroids surrounding P/La Sagra, the purpose of BIM analysis is twofold. If a sus- picious clustering in the longitude of the ascending node (Ω) or the longitude of perihelion () can be found, this analysis would not only allow us to estimate the age of the group, but would also provide support for the hy- pothesis that these objects share a physical origin in the first place. In the outer belt, however, application of this method is generally complicated by large gra- dients of secular frequencies with ap, particularly when considering  in the vicinity of the 2:1 resonance with Jupiter (Nesvorn´y et al. 2003). Although the P/La Sagra group is not particularly close to this resonance, the gradient dg/dap, where g is the frequency of the lon- gitude of perihelion, is still relatively large in this region (∼0.1◦ yr−1 AU−1). As such, for the purpose of our study we only consider Ω. To investigate whether or not a clustering in Ω ex- ists, we numerically integrate the orbits of all six known members of the P/La Sagra group for 20 Myr backward in time. We then calculate the average difference, (cid:104)∆Ω(cid:105), calculated for all possible combinations of objects, where ∆Ω = Ωi − Ωj, and i and j denote two bodies under consideration. Plotting the results of this analysis (Fig- ure 13), we find three clusterings within 60 degrees. The clusterings at about 4.9 and 18.2 Myr ago are less sig- nificant than the one occurring about 6.2 Myr ago. This implies that the age of the cluster is likely close to 6.2 Myr, but due to the small number of asteroids involved, it is not possible to draw a definitive conclusion about the physical significance of this cluster. The results shown in Figure 13 are obtained in a purely gravitational model. Including other forces (e.g., the Yarkovsky effect) could affect the positions and depths of 9 various minima (cf. Novakovi´c 2010). We believe, how- ever, that such an analysis should be deferred pending the discovery of more members of the cluster so that more statistically meaningful results can be obtained than are currently achievable with just six possible cluster mem- bers. In the meantime, the determination of colors or acquisition of spectra of the currently suspected cluster members could provide some preliminary insights into the likelihood that these objects are physically related. 4. DISCUSSION To date, the only published analysis of P/La Sagra prior to this work was presented by Moreno et al. (2011), who focused on numerical modeling of object's dust emis- sion. They found that the activity can be plausibly mod- eled by an episode of continuous dust emission, consis- tent with a sublimation-driven dust ejection process and inconsistent with impulsive dust ejection as would be ex- pected from an impact, beginning shortly before the ob- ject's perihelion passage and persisting over >7 months. Furthermore, they found that the dust cloud's morphol- ogy was best reproduced by anistropic dust emission and indicated that their modeling suggests that P/La Sagra's activity could be seasonally modulated in a man- ner similar to that of other MBCs (e.g., Hsieh et al. 2004, 2011a). In our photometric analysis of the dust cloud (Section 3.2.2), we found steadily increasing dust mass between August 2011 and December 2011, indicative of ongoing dust production, and consistent with the con- clusion by Moreno et al. (2011) that P/La Sagra's dust emission appears to have persisted over an extended pe- riod of time, and therefore was likely to be cometary (i.e., sublimation-driven) in nature. While the conclusion that P/La Sagra's dust emission is driven by the sublimation of volatile ices appears to be at odds with the results of our spectroscopic analysis (Section 3.3), we note that the extreme difficulty of de- tecting sublimation products in a cometary coma at the distance of the main belt means that the non-detection of gas emission in such observations cannot be consid- ered conclusive evidence against the presence of gas (cf. Hsieh et al. 2012a). Rather, such non-detections should only be interpreted as evidence that the levels of gas emission for all MBCs for which spectroscopy has been obtained to date (133P, P/Garradd, and P/La Sagra; Li- candro et al. 2011; Jewitt et al. 2009, and this work) have simply been below the sensitivity levels of those observa- tions. However, the aforementioned observations were all obtained using large ground-based telescopes (the Euro- pean Southern Observatory's 8 m Very Large Telescope in Chile, and the 10 m Keck I telescope in Hawaii), and so future attempts to detect gas in MBCs may need to focus on significantly more active MBCs (i.e., with higher gas emission levels) since significantly more powerful observ- ing facilities are not currently available. It may also be possible that given the large distances and low activity levels (relative to other comets observed at much smaller heliocentric and geocentric distances) of the MBCs, con- clusive evidence of gas emission may only ever be ob- tainable from a close-approaching spacecraft equipped to make such measurements. Despite the lack of a successful gas detection (Sec- tion 3.3), we conclude from the numerical modeling of Moreno et al. (2011) and the morphological and pho- 10 tometric analyses presented in this work (Sections 3.1 and 3.2) that P/La Sagra's activity is likely to be sublimation-driven. As such, the object is likely to be a bona fide MBC like 133P, 176P, 238P, and P/Garradd, where dust is ejected via gas drag from sublimating ice, and not as ejecta from a recent impact as is suspected in the cases of disrupted asteroids, P/2010 A2 and (596) Scheila (e.g., Jewitt et al. 2010, 2011; Snodgrass et al. 2010; Bodewits et al. 2011). Definitive criteria for identifying the true nature of comet-like activity remain elusive, however (cf. Hsieh et al. 2012a). In the case of P/La Sagra, Moreno et al. (2011) show that comet-like dust emission can plausi- bly explain the appearance of P/La Sagra's dust tail, but by no means show that their solution is unique, a common weakness in virtually all dust modeling efforts where models are underconstrained by available observa- tions. Given the case of P/2010 A2's long-lived dust tail (Jewitt et al. 2010; Snodgrass et al. 2010), the longevity of P/La Sagra's activity is likewise not incontrovertible evidence of sublimation-driven dust emission. Even the apparently steadily increasing dust mass in P/La Sagra's coma and tail (Section 3.2.2) could be considered incon- clusive evidence of sublimation given the potential that fragmenting dust particles ejected in an impulsive event could mimic ongoing dust production as in the case of the 2007 outburst of comet 17P/Holmes (Hsieh et al. 2010a). At this time, recurrent activity, separated by interven- ing periods of inactivity, appears to be the only reliably observable indicator of sublimation-driven dust emission that cannot also be plausibly explained by a scenario in- volving impact-driven dust ejection (Hsieh et al. 2011c). Repeated impacts on individual asteroids are unexpected from either theoretical calculations or empirical obser- vations. The tendency of episodes of repeated activity to occur over similar arcs of each object's orbit makes such behavior even more difficult to explain as the ac- tion of random impacts rather than the effect of periodic cometary sublimation. Two of the six currently known MBCs have been observed to exhibit recurrent activ- ity to date (133P and 238P; Hsieh et al. 2004, 2010b, 2011b), while 176P is expected to exhibit renewed activ- ity shortly (Hsieh et al. 2011a). Confirmation of renewed activity in 176P will be crucial for supporting the hypoth- esis that its activity is cometary in nature. The remain- ing three MBCs (P/Garradd, P/La Sagra, and P/2006 VW139), on the other hand, were actually discovered so recently that they have not yet completed full orbits since being discovered, but as they do (P/Garradd in early 2013, P/La Sagra in mid-2015, and P/2006 VW139 in mid-2016), observations to search for recurrent activity are highly encouraged. 5. SUMMARY & CONCLUSIONS We present a wide-ranging study of the recently discov- ered comet P/2010 R2 (La Sagra) including photometric and spectroscopic observations and dynamical analyses of the object. Key results are as follows: 1. A year-long observational monitoring campaign of P/La Sagra from August 2010 to August 2011 show that the morphology of its dust cloud undergoes substantial evolution, exhibiting a strong coma and a single tail from August 2010 to November 2010, and a coma and two tails from December 2010 to February 2011, before diminishing to a faint, de- tached dust trail by August 2011. A long-lived an- tisolar dust tail and the steadily increasing scatter- ing surface area measured for the dust cloud over much of this period suggests that dust production was ongoing over a period of several months, be- havior consistent with dust ejection via the subli- mation of volatile ice and inconsistent with dust ejection via the impulsive action of an impact on the surface of P/La Sagra. 2. We estimate that the nucleus of P/La Sagra has an absolute magnitude of HR = 17.9 ± 0.3 mag, corresponding to an approximate effective radius of rN = 0.7 ± 0.1 km. The B − R color of the nucleus (B − R = 0.9 ± 0.1 mag) is approximately solar, as are the B − R colors of the active comet measured on two separate occasions in 2010. 3. Optical spectroscopic observations of the active P/La Sagra 20 days after it was discovered reveal no evidence of CN emission. Based on the sensitiv- ity of our observations, we find upper-limit CN and H2O production rates of QCN < 6 × 1023 mol s−1 and QH2O < 1026 mol s−1, respectively. 4. Numerical integrations show that P/La Sagra is largely dynamically stable, indicating it is likely native to its current location in the main belt. The surrounding region in orbital element space, how- ever, is moderately unstable with two resonances, the 13:6 mean-motion resonance with Jupiter and the (3,-2,-1) three-body mean-motion resonance with Jupiter and Saturn, present nearby. 5. We find that P/La Sagra is not associated with any known asteroid families, though find a small cluster of asteroids in its dynamical vicinity. Due to the small number of objects in this cluster at the current time, however, we cannot confirm whether members of this cluster in fact resulted from the recent fragmentation of a larger parent asteroid, or if they are simply clustered in orbital element space by chance. Despite our unsuccessful attempt to detect gas emission from P/La Sagra, we conclude from the long-lasting and steadily evolving morphology of the object's dust cloud that its activity is likely to be sublimation-driven, and unlikely to be the result of an impulsive impact. As such, it is likely to be a genuine comet, making it just the fifth MBC discovered (out of a total of six known to date). Numerical simulations indicating that P/La Sagra is likely to be native to its current location in the main belt mean that its physical composition is likely to be rep- resentative of other objects in the same region. A search for other MBCs in this region of orbital element space could prove fruitful. It has been hypothesized, however, that present-day MBC activity may require triggering by small impacts that excavate near-surface ice and ex- pose it to the Sun (Hsieh et al. 2004; Hsieh 2009). As such, we caution that the lower rate of potential activity- triggering impacts at high inclinations (Farinella & Davis 1992) means that, overall, the rate of currently active 11 tiva (Argentina). Observations at the Danish 1.54m telescope were supported by the Danish Natural Sci- ence Research Council (FNU). Funding for SDSS-III has been provided by the Alfred P. Sloan Foundation, the Participating Institutions, the National Science Founda- tion, and the U.S. Department of Energy Office of Sci- ence. The SDSS-III web site is http://www.sdss3.org/. SDSS-III is managed by the Astrophysical Research Con- sortium for the Participating Institutions of the SDSS-III Collaboration. MBCs in this region is likely to be lower than at lower inclinations. In any event, the presence of P/La Sagra in a com- pletely disparate region of orbital element space from the other MBCs underscores the caveat by Hsieh (2009) that the currently known population is far too small to make any conclusions about the abundance and distribution of such objects in the asteroid belt. Current and upcoming systematic searches of the entire asteroid belt for MBCs, now starting to see success in the case of PS1 (Hsieh et al. 2011c, 2012b), will be critical for ascertaining the true abundance and distribution of icy objects in the inner so- lar system. Support for this work was provided by NASA to H.H.H. through Hubble Fellowship grant HF-51274.01 awarded by the Space Telescope Science Institute, which is operated by the Association of Universities for Re- search in Astronomy, Inc., for NASA, under contract NAS 5-26555. B.Y. and N.H. acknowledge support through the NASA Astrobiology Institute under Cooper- ative Agreement No. NNA08DA77A issued through the Office of Space Science. B.N. is supported by the Min- istry of Education and Science of Serbia, under Project 176011, while M.G. is funded by the Academy of Fin- land grant #137853. C.S. additionally acknowledges funding from the European Union Seventh Framework Programme (FP7/2007-2013) under grant agreement no. 268421. We are grateful to John Dvorak and Richard Morriarty, Greg Wirth, Heather Hershley, Scott Dahm, Marc Kassis, Luca Rizzi, and Julie Renaud-Kim, and Jay Rhee, Jonathan Kemp, and Chad Trujillo for their assis- tance in obtaining observations on the UH 2.2 m, Keck, and Gemini telescopes, respectively, and Ilona Busenben- der and Eduardo Solares for technical assistance at the INT. We also thank Alberto Cellino for valuable discus- sions about this work. Acquisition of some of the data presented was enabled using the PS1 System operated by the PS1 Science Consortium (PS1SC) and its mem- ber institutions. The PS1 Survey has been made possible through contributions of the Institute for Astronomy, the University of Hawaii, the Pan-STARRS Project Office, the Max-Planck Society and its participating institutes, the Max Planck Institute for Astronomy, Heidelberg and the Max Planck Institute for Extraterrestrial Physics, Garching, The Johns Hopkins University, Durham Uni- versity, the University of Edinburgh, Queen's Univer- sity Belfast, the Harvard-Smithsonian Center for Astro- physics, and the Las Cumbres Observatory Global Tele- scope Network, Incorporated, the National Central Uni- versity of Taiwan, and the National Aeronautics and Space Administration under Grant No. NNX08AR22G issued through the Planetary Science Division of the NASA Science Mission Directorate. The Gemini Obser- vatory is operated by the Association of Universities for Research in Astronomy, Inc., under a cooperative agree- ment with the NSF on behalf of the Gemini partnership: the National Science Foundation (United States), the Science and Technology Facilities Council (United King- dom), the National Research Council (Canada), CON- ICYT (Chile), the Australian Research Council (Aus- tralia), Minist´erio da Ciencia e Tecnologia (Brazil) and Ministerio de Ciencia, Tecnolog´ıa e Innovaci´on Produc- 12 A'Hearn, M. F. 1982, IAU Colloq. 61: Comet Discoveries, Statistics, and Observational Selection, 433 Hsieh, H. H., et al. 2012b, ApJ, submitted Ipatov, S. I., & Hahn, G. J. 1997, in Lunar and Planetary Science AHearn, M. F., Schleicher, D. G., Millis, R. L., Feldman, P. D., XXVIII (Houston: Lunar Planet. Inst.), 619 Thompson, D. T. 1984, AJ, 89, 579 Ishiguro, M., Sarugaku, Y., Ueno, M., Miura, N., Usui, F., Chun, A'Hearn, M. F., Millis, R. L., Schleicher, D. G., Osip, D. J., & M.-Y., & Kwon, S. M. 2007, Icarus, 189, 169 REFERENCES Birch, P. V. 1995, Icarus, 118, 223 Asphaug, E. 1997, Meteor. Planet. Sci., 32, 965-980. Birtwhistle, P., Ryan, W. H., Sato, H., Beshore, E. C., & Kadota, K. 2010, IAU Circ., 9105, 1 Bodewits, D., Kelley, M. S., Li, J.-Y., Landsman, W. B., Besse, S., & A'Hearn, M. F. 2011, ApJ, 733, L3 Bottke, W. F., Durda, D. D., Nesvorn´y, D., Jedicke, R., Morbidelli, A., Vokrouhlick´y, D., & Levison, H. F. 2005, Icarus, 179, 63 Burns, J. A., Lamy, P. L., & Soter, S. 1979, Icarus, 40, 1 Buzzoni, B., Delabre, B., Dekker, H., Dodorico, S., Enard, D., Focardi, P., Gustafsson, B., Nees, W., Paureau, J., & Reiss, R. 1984, The ESO Messenger, 38, 9 Chambers, J. E. 1999, MNRAS, 304, 793 Cheng, A. F. 2004, Icarus, 169 357 Cochran, A. L., Barker, E. S., Ramseyer, T. F., & Storrs, A. D. 1992, Icarus, 98, 151 Elst, E. W., Pizarro, O., Pollas, C., Ticha, J., Tichy, M., Moravec, Z., Offutt, W., & Marsden, B. G. 1996, IAUC 6496 Epifani, E., Rotundi, A., Foster, M. J., Green, S. F., Colangeli, L., Fulle, M., Mennella, V., & Palumbo, P. 1999, Planet. Space Sci., 47, 765 Fanale, F. P., & Salvail, J. R. 1984, Icarus, 60, 476 Farinella, P. & Davis, D. R. 1992, Icarus, 97, 111 Fern´andez, J. A., Gallardo, T., & Brunini, A. 2002, Icarus, 159, 358 Finson, M. L., & Probstein, R. F. 1968, ApJ, 154, 327-380. Florczak, M., Lazzaro, D., Moth´e-Diniz, T., Angeli, C. A., & Betzler, A. S. 1999, A&AS, 134, 463 Fujiwara, A., et al. 2006, Science, 312, 1330-1334. Fukugita, M., Ichikawa, T., Gunn, J. E., Doi, M., Shimasaku, K., & Schneider, D. P. 1996, AJ, 111, 1748 Fulle, M. 2004, Comets II, 565 Garradd, G. J., Sostero, G., Camilleri, P., et al. 2008, IAU Circ., 8969, 1 Gil-Hutton, R. 2006, Icarus 183, 93. Haghighipour, N. 2009, Meteoritics & Planetary Science, 44, 1863-1869. Hainaut, O. R. et al. 2012, A&A, in press Hanner, M. S., Giese, R. H., Weiss, K., & Zerull, R. 1981, A&A, 104, 42 Hardorp, J. 1980, A&A, 88, 334 Hartmann, W., Cruikshank, D., & Degewij, J. 1982, Icarus, 52, 377 Hartmann, W., Tholen, D., Meech, K., & Cruikshank, D. 1990, Icarus, 83, 1 Haser, L. 1957, Bulletin de la Societe Royale des Sciences de Liege, 43, 740 Hirayama, K. 1918, AJ, 31, 185-188. Hook, I. M., Jørgensen, I., Allington-Smith, J. R., Davies, R. L., Metcalfe, N., Murowinski, R. G., & Crampton, D. 2004, PASP, 116, 425 Hsieh, H. H., Jewitt, D., & Fern´andez, Y. R. 2004, AJ, 127, 2997 Hsieh, H. H., & Jewitt, D. 2006, Science, 312, 561 Hsieh, H. H., Jewitt, D., & Pittichov´a, J. 2006, IAUC 8704, 1 Hsieh, H. H. 2009, A&A, 505, 1297 Hsieh, H. H., Jewitt, D., & Fern´andez, Y. R. 2009a, ApJ, 694, L111 Hsieh, H. H., Jewitt, D., & Ishiguro, M. 2009b, AJ, 137, 157 Hsieh, H. H., Fitzsimmons, A., Joshi, Y., Christian, D., & Pollacco, D. L. 2010a, MNRAS, 407, 1784-1800 Hsieh, H. H., Jewitt, D., Lacerda, P., Lowry, S. C., & Snodgrass, C. 2010b, MNRAS, 403, 363 Ishiguro, M., Hanayama, H., Hasegawa, S., Sarugaku, Y., Watanabe, J., Fujiwara, H., Terada, H., Hsieh, H. H., Vaubaillon, J. J., Kawai, N., Yanagisawa, K., Kuroda, D., Miyaji, T., Fukushima, H., Ohta, K., Hamanowa, H., Kim, J., Pyo, J., & Nakamura, A. M. 2011a, ApJ, 740, L11 Ishiguro, M., Hanayama, H., Hasegawa, S., Sarugaku, Y., Watanabe, J., Fujiwara, H., Terada, H., Hsieh, H. H., Vaubaillon, J. J., Kawai, N., Yanagisawa, K., Kuroda, D., Miyaji, T., Fukushima, H., Ohta, K., Hamanowa, H., Kim, J., Pyo, J., & Nakamura, A. M. 2011b, ApJ, 741, L24 Jewitt, D. C. 2002, AJ, 123, 1039 Jewitt, D., Sheppard, S., & Fern´andez, Y. 2003, AJ, 125, 3366 Jewitt, D., Lacerda, P., & Peixinho, N. 2007, IAU Circ., 8847, 1 Jewitt, D., Yang, B., & Haghighipour, N. 2009, AJ, 137, 4313 Jewitt, D., Weaver, H., Agarwal, J., Mutchler, M., & Drahus, M. 2010, Nature, 467, 817 Jewitt, D., Weaver, H., Mutchler, M., Larson, S., & Agarwal, J. 2011, ApJ, 733, L4 Jutzi, M., Benz, W., & Michel, P. 2008, Icarus, 198, 242-255. Kimura, H., Okamoto, H., & Mukai, T. 2002, Icarus, 157, 349 Kitazato, K., Abe, M., Ishiguro, M., & Ip, W.-H. 2007, A&A, 474, L5-L8. Knezevi´c, Z., & Milani, A. 2000, Celest. Mech. 78, 17 Kres´ak, L. 1980, Moon and Planets, 22, 83 Landolt, A. U. 1992, AJ, 104, 340 Larson, S. M. 2010, IAU Circ., 9188, 1 Levison, H. F., Terrell, D., Wiegert, P. A., Dones, L., & Duncan, M. J. 2006, Icarus, 182, 161 Licandro, J., Campins, H., Tozzi, G. P., de Le´on, J., Pinilla-Alonso, N., Boehnhardt, H., & Hainaut, O. R. 2011, A&A, 532, A65. Lisse, C. M., A'Hearn, M. F., Hauser, M. G., Kelsall, T., Lien, D. J., Moseley, S. H., Reach, W. T., & Silverberg, R. F. 1998, ApJ, 496, 971-991. Lisse, C. M., Fern´andez, Y. R., A'Hearn, M. F., Grun, E., Kaufl, H. U., Osip, D. J., Lien, D. J., Kostiuk, T., Peschke, S. B., & Walker, R. G. 2004, Icarus, 171, 444-462. Lowry, S. C., Fitzsimmons, A., & Collander-Brown, S. 2003, A&A, 397, 329 Lowry, S. C., & Weissman, P. R. 2007, Icarus, 188, 212 Maclennan, E. M., & Hsieh, H. H. 2011, ApJ, submitted Magnier, E. 2006, Proceedings of The Advanced Maui Optical and Space Surveillance Technologies Conference, Ed.: S. Ryan, The Maui Economic Development Board, p.E5 Nomen, J., et al. 2010, IAU Circ., 9169, 1 Marzari, F., Davis, D., & Vanzani V. 1995, Icarus, 113, 168 Milani, A., & Knezevi´c, Z. 1994, Icarus, 107, 219 Moreno, F., Lara, L. M., Licandro, J., Ortiz, J. L., De L´eon, J., Al´ı-Lagoa, V., Agis-Gonz´alez, B., & Molina, A. 2011, ApJ, 738, L16. Nesvorn´y, D., & Morbidelli, A. 1998, AJ, 116, 3029-3037 Nesvorn´y, D., Bottke, W. F., Jr., Dones, L., & Levison, H. F. 2002, Nature, 417, 720 Nesvorn´y, D., Bottke, W. F., Levison, H. F., & Dones, L. 2003, ApJ, 591, 486 Nesvorn´y, D., & Vokrouhlick´y, D. 2006, AJ, 132, 1950 Nesvorn´y, D., Bottke, W. F., Vokrouhlick´y, D., Sykes, M., Lien, D. J., & Stansberry, J. 2008, ApJ, 679, L143 Novakovi´c, B. 2010, MNRAS, 407, 1477 Novakovi´c, B., Cellino, A., Knezevi´c, Z. 2011, Icarus, 216, 69 Oke, J. B., Cohen, J. G., Carr, M., Cromer, J., Dingizian, A., Harris, F. H., Labrecque, S., Lucinio, R., Schaal, W., Epps, H., & Miller, J. 1995, PASP, 107, 375 Hsieh, H. H., Ishiguro, M., Lacerda, P., & Jewitt, D. 2011a, AJ, Reach, W. T., Sykes, M. V., Lien, D., & Davies, J. K. 2000, 142, 29 Icarus, 148, 80-94. Hsieh, H. H., Meech, K. J., & Pittichov´a, J. 2011b, ApJ, 736, L18 Hsieh, H. H., Denneau, L., Wainscoat, R. J., Fitzsimmons, A., Armstrong, J. D., Yang, B., Hergenrother, C. W. 2011c, CBET 2920 Read, M. T., Bressi, T. H., Gehrels, T., Scotti, J. V., & Christensen, E. J. 2005, IAU Circ., 8624, 1 Russell, H. N. 1916, ApJ, 43, 173 Sarugaku, Y., Ishiguro, M., Pyo, J., Miura, N., Nakada, Y., Usui, Hsieh, H. H., Yang, B., & Haghighipour, N. 2012a, ApJ, 744, 9 F., Ueno, M. 2007, Publ. Astron. Soc. Japan 59, L25 TABLE 1 Comet-Like Main-Belt Asteroids Name Typea ab ec id 133P/Elst-Pizarro 238P/Read 176P/LINEAR P/2008 R1 (Garradd) P/2010 A2 (LINEAR) P/2010 R2 (La Sagra) (596) Scheila P/2006 VW139 MBC 3.160 MBC 3.165 MBC 3.194 MBC 2.726 DA 2.291 MBC 3.099 DA 2.928 MBC 3.052 0.162 0.253 0.194 0.342 0.124 0.154 0.165 0.201 1.386 1.266 0.238 15.903 5.255 21.395 14.661 2.438 e TJ 3.184 3.153 3.166 3.217 3.583 3.099 3.209 3.203 P f 5.62 5.63 5.71 4.50 3.47 5.46 5.01 5.33 g,h HR 15.49 ± 0.05 [1] 19.05 ± 0.05 [2] 15.10 ± 0.05 [1] 20.3 ± 0.1 [3] 21.9 ± 0.1 [4] 17.9 ± 0.2 [5] 8.54 [6] 15.9 [7] 13 i rn 1.9 0.4 2.0 0.2 0.06 0.7 56.67 1.7 (Ad/AN )max j,h Discovery Datek 0.7 [8] 21 [2] 0.3 [9] 28 [3] -- 60 [5] 2.2 [10] 1.1 [11] 1996 Aug 7 [12] 2005 Oct 24 [13] 2005 Nov 26 [14] 2008 Sep 2 [15] 2010 Jan 6 [16] 2010 Sep 14 [17] 2010 Dec 10 [18] 2011 Nov 05 [19] a Object classification as main-belt comet (MBC) or disrupted asteroid (DA). b Osculating semimajor axis, in AU. c Osculating eccentricity. d Osculating inclination, in degrees. e Tisserand parameter. f Orbital period, in years. g Absolute R-band magnitude of nucleus, assuming solar colors, in the IAU H, G system. h References: [1] Hsieh et al. (2009a); [2] Hsieh et al. (2011b); [3] Maclennan & Hsieh (2012); [4] Jewitt et al. (2010); [5] this work; [6] Tedesco et al. (2004); [7] JPL Small Body Database (http://ssd.jpl.nasa.gov/sbdb.cgi); [8] Hsieh et al. (2010b); [9] Hsieh et al. (2011a); [10] Jewitt et al. (2011); [11] Hsieh et al. (2012b); [12] Elst et al. (1996); [13] Read et al. (2005); [14] Hsieh et al. (2006); [15] Garradd et al. (2008); [16] Birtwhistle et al. (2010); [17] Nomen et al. (2010); [18] Larson (2010); [19] Hsieh et al. (2011c). i Estimated effective radius of nucleus, from same works referenced for HR, in km. j Peak observed dust-to-nucleus scattering surface area ratio. k Discovery date of comet-like activity. Schleicher, D. G. 2010, AJ, 140, 973 Schwarz, H. E., et al. 2004, Proc. SPIE, 5492, 564 Snodgrass, C., Tubiana, C., Vincent, J.-B., Sierks, H., Hviid, S., Moissi, R., Boehnhardt, H., Barbieri, C., Koschny, D., Lamy, P., Rickman, H., Rodrigo, R., Carry, B., Lowry, S. C., Laird, R. J. M., Weissman, P. R., Fitzsimmons, A., Marchi, S., & the OSIRIS Team. 2010, Nature, 467, 814 Stellingwerf, R. F. 1978, ApJ, 224, 953 Stubbs, C. W., Doherty, P., Cramer, C., Narayan, G., Brown, Y. J., Lykke, K. R., Woodward, J. T., & Tonry, J. L. 2010, ApJS, 191, 376 Tedesco, E. F., Noah, P. V., Noah, M., & Price, S. D. 2004, IRAS Minor Planet Survey, IRAS-A-FPA-3-RDR-IMPS-V6.0. NASA Planetary Data System. Tsiganis, K., Varvoglis, H., & Morbidelli, A. 2003, Icarus, 166, 131 Vaghi, S. 1973, A&A, 24, 107. Warner, B. D., Harris, A. W., & Pravec, P. 2009, Icarus, 202, 134-146. Yang, B., & Hsieh, H. H. 2011, ApJ, 737, L39 York, D. G., et al. 2000, AJ, 120, 1579-1587. Zappal`a, V., Celliono, A., Farinella, P., & Knezevi´c, Z. 1990, AJ, 100, 2030 Zappal`a, V., Cellino, A., Farinella, P., & Milani, A. 1994, AJ, 107, 772. 14 TABLE 2 Observation Log UT Date Telescopea Moonb Seeingc Nd te Filter Rf ∆g αh νi P.A.−(cid:12)j P.A.−v k l αpl N -- N+6 PS1 PS1 Perihelion INT Keck Keck -- 2 1 20 5 3 2 4 3 6 21 4 2 4 6 -- -- r(cid:48) r(cid:48) R R R BR Spec. -- 1.0 0.9 1.4 1.6 1.5 0.9 0.9 1.1 1.8 0.8 1.1 0.8 1.0 0.8 -- -- Dk1.54 UH2.2 FTN Keck Keck UH2.2 UH2.2 UH2.2 N+11 N+11 N−9 N−2 N−2 N+11 N−9 N+6 N−3 N N−3 N+2 2.228 1.793 1.739 1.756 1.756 1.790 1.824 1.824 1.922 2.343 2.535 2.793 3.210 3.289 3.231 2.758 2.946 2010 Jun 26 2010 Aug 16 2010 Sep 08 2010 Sep 19 (1) 2010 Sep 19 (2) 2010 Sep 29 2010 Oct 05 2010 Oct 05 2010 Oct 19 2010 Nov 26 2010 Dec 12 2011 Dec 31 2011 Feb 03 2011 Aug 26 2011 Aug 31 2013 Mar 13 2015 Nov 30 -- -- 2.623 2.631 80 40 2.641 2.649 2400 2.649 1500 2.652 450 2.659 240 2.659 3600 900 2.666 2.698 1800 2.713 6300 2.733 1200 2.772 360 3.066 1200 1800 3.073 -- -- 3.570 -- -- 2.620 -0.5 -9.4 -11.7 -12.0 -12.0 -11.7 -11.3 -11.3 -9.8 -4.0 -1.5 1.2 4.4 -5.7 -6.0 7.5 -2.8 a Telescope used (PS1: 1.8-m Pan-STARRS 1 telescope; Dk1.54: Danish 1.54-m telescope; FTN: 2.0-m Faulkes Telescope North; Gemini: 8-m Gemini-North telescope; Keck: Keck I 10-m telescope; UH2.2: University of Hawaii 2.2-m telescope; INT: 2.5-m Isaac Newton Telescope) b Phase of moon expressed in offset from new moon ("N") in days. c Approximate average seeing (FWHM) in arcsec. d Number of images. e Total effective exposure time in seconds. f Heliocentric distance in AU. g Geocentric distance in AU. h Solar phase angle (Sun - P/La Sagra - Earth) in degrees. i True anomaly in degrees. j Position angle of the antisolar vector, as projected in the plane of the sky, in degrees east of north. k Position angle of the negative velocity vector, as projected in the plane of the sky, in degrees east of north. l Orbit-plane angle (between observer and object orbit plane as seen from object) in degrees. 241.2 203.3 162.1 138.8 138.0 119.9 110.7 110.7 94.5 71.3 65.7 60.5 53.4 270.9 270.8 76.6 67.0 242.8 243.7 241.1 239.7 239.7 238.7 238.2 238.2 237.7 239.6 241.4 243.9 248.7 291.6 292.5 302.8 238.0 0.0 12.9 18.5 21.3 21.3 23.7 25.2 25.2 28.5 37.8 41.3 45.9 53.5 95.1 96.0 180.0 0.0 22.3 15.0 12.0 12.3 12.3 13.4 14.4 14.4 16.8 21.1 21.3 20.5 17.0 17.9 18.2 10.5 19.3 Gemini Aphelion Perihelion R R R R BR BR r(cid:48) -- -- TABLE 3 Photometry Results f e c a g mtot mavg UT Date 18.3 ± 0.1 18.0 ± 0.1 17.9 ± 0.1 17.8 ± 0.1 18.0 ± 0.1 18.0 ± 0.1 18.1 ± 0.1 18.4 ± 0.1 18.4 ± 0.1 18.2 ± 0.2 19.1 ± 0.1 18.49 ± 0.05 18.26 ± 0.05 18.28 ± 0.02 18.23 ± 0.02 18.39 ± 0.02 18.34 ± 0.03 18.47 ± 0.03 18.89 ± 0.03 18.78 ± 0.02 18.93 ± 0.04 19.53 ± 0.08 24.1 ± 0.1 23.4 ± 0.1 mavg(1, 1, 0)b 14.25 ± 0.15 14.20 ± 0.15 14.15 ± 0.15 14.10 ± 0.15 14.19 ± 0.15 14.05 ± 0.20 13.97 ± 0.20 13.80 ± 0.20 13.50 ± 0.20 13.45 ± 0.20 13.85 ± 0.20 18.1 ± 0.3 17.4 ± 0.3 mtot(1, 1, 0)d Ad/AN 30 ± 10 14.0 ± 0.2 13.9 ± 0.2 30 ± 10 13.8 ± 0.2 30 ± 10 13.7 ± 0.2 30 ± 10 30 ± 10 13.8 ± 0.2 35 ± 10 13.7 ± 0.2 35 ± 10 13.6 ± 0.2 13.3 ± 0.2 45 ± 15 13.1 ± 0.2 60 ± 20 60 ± 20 12.7 ± 0.2 13.4 ± 0.2 40 ± 10 2010 Aug 16 2010 Sep 08 2010 Sep 19 (1) 2010 Sep 19 (2) 2010 Sep 29 2010 Oct 05 2010 Oct 19 2010 Nov 26 2010 Dec 12 2010 Dec 31 2011 Feb 03 2011 Aug 26 2011 Aug 31 a Mean apparent magnitude inside a 5.(cid:48)(cid:48)0 photometry aperture. b Mean magnitude, normalized to R = ∆ = 1 AU and α = 0◦, assuming −0.03 < GR < 0.26. c Total apparent magnitude inside rectangular aperture enclosing entire comet (nucleus and dust tail). d Total magnitude, normalized to R = ∆ = 1 AU and α = 0◦, assuming −0.03 < GR < 0.26. e Inferred ratio of scattering surface area of dust to nucleus scattering surface area. f Inferred scattering surface area of dust, in 106 m2, using rN = 0.6 km. g Estimated dust mass, in 105 kg, assuming 10 µm-radius grains and ρ = 1300 kg m−3. h Dust contribution (computed using a 5.(cid:48)(cid:48)0 photometry aperture), as parameterized by A'Hearn et al. (1984), in cm. i Computed from data in which comet was trailed by 4.(cid:48)(cid:48)1, and as such, represents a lower limit. Ad 40 ± 15 45 ± 15 45 ± 15 45 ± 15 45 ± 15 50 ± 20 55 ± 20 65 ± 20 85 ± 30 85 ± 30 60 ± 20 Md 7 ± 2 7 ± 2 8 ± 2 8 ± 2 8 ± 2 8 ± 3 9 ± 3 11 ± 4 14 ± 5 15 ± 5 10 ± 3 -- -- -- -- -- -- -- -- -- -- Af ρh 42 ± 6 45 ± 6 46 ± 6 48 ± 6 44 ± 6 49 ± 7 50 ± 8 49 ± 9 60 ± 11 57 ± 10i 34 ± 6 0.7 ± 0.1 1.2 ± 0.2 15 TABLE 4 P/La Sagra Cluster Members Object a ap b ep sin(ip)c HV d Tlyap e 0.1223 0.1146 0.1215 0.1177 0.1150 0.1148 0.3791 0.3812 0.3824 0.3831 0.3824 0.3817 12.1 14.4 14.5 14.2 15.4 18.3 683 480 781 999 900 270 18901 (2000 MR5) 106020 (2000 SS294) 106064 (2000 SA323) 131501 (2001 SX272) (2002 RQ261) P/2010 R2 (La Sagra) 3.0942 3.0928 3.0931 3.0889 3.0887 3.1001 a Proper semimajor axis, in AU b Proper eccentricity c Sine of proper inclination d Absolute magnitude, in V -band e Lyapunov time, in kyr Fig. 1. -- Plots of eccentricity (upper panel) and inclination (lower panel) versus semimajor axis showing the distributions in orbital element space of main-belt asteroids (black dots), MBCs (red circles), and likely disrupted asteroids (blue circles). Also marked with dotted lines are the semimajor axes of Mars (aMars) and Jupiter (aJup), the semimajor axis of the 2:1 mean-motion resonance with Jupiter, and the loci of Mars-crossing orbits and Jupiter-crossing orbits. 16 Fig. 2. -- Composite images of P/La Sagra (at the center of each panel) constructed from data obtained on (a) 2010 August 16 (80 s of effective exposure time on PS1 in r(cid:48)-band), (b) 2010 September 08 (40 s on PS1 in r(cid:48)-band), (c) 2010 September 19 (2400 s on the Danish 1.54 m in R-band), (d) 2010 September 19 (1500 s on the UH 2.2 m in R-band), (e) 2010 September 29 (450 s on FTN in R-band), (f) 2010 October 05 (240 s on Keck I in R-band), (g) 2010 October 19 (900 s on the UH 2.2 m in R-band), (h) 2010 November 26 (1800 s on the UH 2.2 m in R-band), (i) 2010 December 12 (5400 s on the UH 2.2 m in R-band), (j) 2010 December 31 (1200 s on the INT in R-band), and (k) 2011 February 03 (360 s on Keck I in R-band). All panels are 60(cid:48)(cid:48) × 60(cid:48)(cid:48) in size with north (N), east (E), the antisolar direction (−(cid:12)), and the negative heliocentric velocity vector (−v), as projected on the sky, marked. Fig. 3. -- Composite images of P/La Sagra (point-like object in the lower left of each panel) constructed from data obtained on (a) 2011 August 26 (1200 s on Keck I in R-band) and (b) 2011 August 31 (1800 s on Gemini in r(cid:48)-band). Panels are 120(cid:48)(cid:48) × 60(cid:48)(cid:48), with north (N), east (E), the antisolar direction (−(cid:12)), and the negative heliocentric velocity vector (−v), as projected on the sky, marked. 17 Fig. 4. -- Orbital position plot of P/La Sagra observations detailed in Table 2. The Sun is shown at the center as a solid dot, with the orbits of Mercury, Venus, Earth, Mars, P/La Sagra, and Jupiter (from the center of the plot outwards) shown as black lines. Solid squares mark positions where P/La Sagra was observed to be active in 2010 and early 2011, while an open square marks the position where P/La Sagra was observed to be inactive in August 2011. Perihelion (P) and aphelion (A) positions are marked with crosses. Observations plotted were obtained on (a) 2010 August 16, (b) 2010 September 8-19, (c) 2010 October 5, (d) 2010 November 25-28, (e) 2010 December 12, (f) 2010 December 31, (g) 2011 February 3, and (h) 2011 August 26-31. 18 Fig. 5. -- Comparisons of (a) dust-to-nucleus ratios (by scattering cross-section) and (b) total dust scattering cross-sections, in 106 km2, as functions of true anomaly measured for the five known MBCs: 133P (red symbols; Hsieh et al. 2010b), 176P (green symbols; Hsieh et al. 2011a), 238P (orange symbols; Hsieh et al. 2011b), P/Garradd (purple symbols; Jewitt et al. 2009; Maclennan & Hsieh 2012), and P/La Sagra (blue symbols; this work). Observations where activity was detected either visually or from photometry are marked with circular symbols, while observations where no activity was detected are marked with squares. For points where no error bars are visible, the amount of uncertainty is equal to or less than the size of the plotted symbol. 19 Fig. 6. -- Plots of the sky-background-subtracted and flux-calibrated spectrum of P/La Sagra (solid black line), and the scaled spectrum of the solar analog, the G2V star HD 91163 (blue dashed line). The shaded region indicates the wavelength region where the CN emission band is expected. We note that the spectrum of P/La Sagra closely resembles that of HD 91163 and thus no emission features are observed within the uncertainties of our measurements. Fig. 7. -- The spectrum of P/La Sagra with the underlying solar continuum removed. The three red error bars show the 1-σ uncertainties in the three wavelength regions (3760-3830A, 3830-3900A, and 3900-3970A). 20 Fig. 8. -- Plots of semimajor axis versus eccentricity (left) and inclination (right) showing initial osculating elements of test particles in three sets of 100 Gaussian-distributed particles each of which are subjected to a 100 Myr dynamical integration (Section 3.4.1). Each set of test particles is centered on the current osculating orbital elements of P/La Sagra (red circle) and characterized by a σ value equal to the object's JPL-tabulated uncertainties. Particles ejected in less than 20 Myr, between 20 Myr and 50 Myr, and between 50 Myr and 100 Myr are plotted with orange X symbols, yellow asterisk symbols, and green triangles, respectively, while particles that are not ejected after the 100 Myr test period used in our simulations are marked with blue squares. Fig. 9. -- Same as Figure 8 for three sets of 100 Gaussian-distributed particles each, where each set is characterized by a σ value equal to 100× P/La Sagra's JPL-tabulated uncertainties. The locations of the (3,-2,-1) three-body mean-motion resonance (with Jupiter and Saturn) at a = 3.0790 AU and the 13:6 mean-motion resonance (with Jupiter) at a = 3.1063 AU are plotted with dashed lines. 21 Fig. 10. -- Plot of P/La Sagra's semimajor axis (AU) as a function of time (Myr), marked by a black solid line, as simulated in the numerical integrations described in Section 3.4.1, where the semimajor axes of the 13:6 mean-motion resonance with Jupiter and the (3,-2,-1) three-body mean-motion resonance with Jupiter and Saturn are marked with red dashed lines. 22 Fig. 11. -- Synthetic orbital element plots of eccentricity (upper panel) and sine of inclination (lower panel) versus semimajor axis of outer main-belt asteroids (small black dots), where the central bodies of families and clumps identified by Gil-Hutton (2006) are marked with blue and green triangles, respectively, and central bodies of families, clumps, and clusters identified by Novakovi´c et al. (2011) are marked with blue, green, and orange circles, respectively. The location of P/La Sagra in synthetic proper element space is plotted with a red circle. 23 Fig. 12. -- Number of objects associated with the P/La Sagra cluster. The six members determined to belong to this cluster are all linked within 95 m s−1, with no more linkages found to other asteroids until 125 m s−1 when the group merges with the local background population. Our choice for the nominal cutoff distance for the cluster of 110 m s−1 is marked. Fig. 13. -- Average differences (cid:104)∆Ω(cid:105) of the longitudes of the ascending nodes for members of the P/La Sagra cluster. Three significant clusterings at 4.9, 6.2 and 18.2 Myr in the past are marked with arrows.
1609.06718
1
1609
2016-09-21T20:00:01
Exocometary gas structure, origin and physical properties around $\beta$ Pictoris through ALMA CO multi-transition observations
[ "astro-ph.EP" ]
Recent ALMA observations unveiled the structure of CO gas in the 23 Myr-old $\beta$ Pictoris planetary system, a component that has been discovered in many similarly young debris disks. We here present ALMA CO J=2-1 observations, at an improved spectro-spatial resolution and sensitivity compared to previous CO J=3-2 observations. We find that 1) the CO clump is radially broad, favouring the resonant migration over the giant impact scenario for its dynamical origin, 2) the CO disk is vertically tilted compared to the main dust disk, at an angle consistent with the scattered light warp. We then use position-velocity diagrams to trace Keplerian radii in the orbital plane of the disk. Assuming a perfectly edge-on geometry, this shows a CO scale height increasing with radius as $R^{0.75}$, and an electron density (derived from CO line ratios through NLTE analysis) in agreement with thermodynamical models. Furthermore, we show how observations of optically thin line ratios can solve the primordial versus secondary origin dichotomy in gas-bearing debris disks. As shown for $\beta$ Pictoris, subthermal (NLTE) CO excitation is symptomatic of H$_2$ densities that are insufficient to shield CO from photodissociation over the system's lifetime. This means that replenishment from exocometary volatiles must be taking place, proving the secondary origin of the disk. In this scenario, assuming steady state production/destruction of CO gas, we derive the CO+CO$_2$ ice abundance by mass in $\beta$ Pic's exocomets to be at most $\sim$6%, consistent with comets in our own Solar System and in the coeval HD181327 system.
astro-ph.EP
astro-ph
MNRAS 000, 1 -- 21 (2016) Preprint 24 September 2018 Compiled using MNRAS LATEX style file v3.0 Exocometary gas structure, origin and physical properties around β Pictoris through ALMA CO multi-transition observations L. Matr`a1,2(cid:63), W. R. F. Dent3, M. C. Wyatt1, Q. Kral1, D. J. Wilner4, O. Pani´c1, A. M. Hughes5, I. de Gregorio-Monsalvo3, A. Hales3, J.-C. Augereau6,7, J. Greaves8, A. Roberge9 1Institute of Astronomy, University of Cambridge, Madingley Road, Cambridge CB3 0HA, UK 2European Southern Observatory, Alonso de C´ordova 3107, Vitacura, Santiago, Chile 3ALMA SCO, Alonso de C´ordova 3107, Vitacura, Santiago, Chile 4Harvard-Smithsonian Center for Astrophysics, 60 Garden Street, Cambridge, MA 02138, USA 5Department of Astronomy, Van Vleck Observatory, Wesleyan University, 96 Foss Hill Dr., Middletown, CT 06459, USA 6Univ. Grenoble Alpes, Institut de Plan´etologie et d'Astrophysique de Grenoble (IPAG, UMR 5274), 38000 Grenoble, France 7CNRS, Institut de Plan´etologie et d'Astrophysique de Grenoble (IPAG, UMR 5274), 38000 Grenoble, France 8School of Physics & Astronomy, Cardiff University, 4 The Parade, Cardiff CF24 3AA, UK 9Exoplanets and Stellar Astrophysics Laboratory, NASA Goddard Space Flight Center, Greenbelt, MD, USA Accepted 2016 September 20. Received 2016 September 01; in original form 2016 July 18. ABSTRACT Recent ALMA observations unveiled the structure of CO gas in the 23 Myr-old β Pic- toris planetary system, a component that has been discovered in many similarly young debris disks. We here present ALMA CO J=2-1 observations, at an improved spectro- spatial resolution and sensitivity compared to previous CO J=3-2 observations. We find that 1) the CO clump is radially broad, favouring the resonant migration over the giant impact scenario for its dynamical origin, 2) the CO disk is vertically tilted compared to the main dust disk, at an angle consistent with the scattered light warp. We then use position-velocity diagrams to trace Keplerian radii in the orbital plane of the disk. Assuming a perfectly edge-on geometry, this shows a CO scale height increasing with radius as R0.75, and an electron density (derived from CO line ratios through NLTE analysis) in agreement with thermodynamical models. Furthermore, we show how observations of optically thin line ratios can solve the primordial versus secondary origin dichotomy in gas-bearing debris disks. As shown for β Pictoris, sub- thermal (NLTE) CO excitation is symptomatic of H2 densities that are insufficient to shield CO from photodissociation over the system's lifetime. This means that replen- ishment from exocometary volatiles must be taking place, proving the secondary origin of the disk. In this scenario, assuming steady state production/destruction of CO gas, we derive the CO+CO2 ice abundance by mass in β Pic's exocomets to be at most ∼6%, consistent with comets in our own Solar System and in the coeval HD181327 system. Key words: matter -- comets: general -- molecular processes -- stars: individual: β Pictoris. submillimetre: planetary systems -- planetary systems -- circumstellar 6 1 0 2 p e S 1 2 . ] P E h p - o r t s a [ 1 v 8 1 7 6 0 . 9 0 6 1 : v i X r a 1 INTRODUCTION The circumstellar environment of β Pictoris, a nearby (19.44 ± 0.05 pc, van Leeuwen 2007), young (23 ± 3 Myr, Mamajek & Bell 2014) main sequence A6 star (Gray et al. 2006), has been a continuous source of discoveries since (cid:63) E-mail: [email protected] c(cid:13) 2016 The Authors more than three decades ago, when first circumstellar gas through optical absorption lines (Slettebak 1975) and then dust through an infrared excess from IRAS (Aumann 1985) were discovered. Its edge-on geometry was unveiled soon af- ter through the first resolved image of a circumstellar dust disk (Smith & Terrile 1984), and explained the observed cir- cumstellar gas absorption lines (Hobbs et al. 1985). Later, it was understood that to maintain the low levels of dust 2 L. Matr`a et al. present in many nearby main sequence stars such as β Pic- toris, a replenishment mechanism is needed (Backman & Paresce 1993). This need for replenishment of second gener- ation dust sets the physical definition of a debris disk, and represents the fundamental difference with primordial pro- toplanetary disks, where the presence of large amounts of gas means the dust does not need replenishment (e.g. Wy- att et al. 2015). Considered one of the archetypes of debris disks, the presence of gas has given β Pictoris particular attention. The observed atomic absorption lines were seen to have both a stable component, at a radial velocity similar to that of the star, and a variable component, seen at different ra- dial velocities (e.g. Slettebak & Carpenter 1983; Kondo & Bruhweiler 1985; Lagrange et al. 1987). The latter feature was attributed to evaporating cometary bodies approach- ing the star on eccentric orbits (Ferlet et al. 1987; Beust et al. 1990; Kiefer et al. 2014). In addition, Hubble Space Telescope (HST) observations showed that the gas compo- sition is rather peculiar, with an extreme overabundance of carbon compared to other elements. This carbon acts as a braking agent and explains how the observed metallic gas levels can be maintained through braking against stellar ra- diation pressure (Roberge et al. 2006; Fern´andez et al. 2006). Though generally absorption studies against the stellar con- tinuum are the most sensitive, they are limited in that they only probe the gas column along the line of sight to the star. Recent studies have therefore been focusing on emis- sion lines, which on the other hand can be used to trace the overall morphology of the gas disk. Firstly resolved obser- vations of metallic atoms in the optical/UV (Olofsson et al. 2001; Brandeker et al. 2004; Nilsson et al. 2012), and sub- sequently observations of ionised carbon and oxygen in the far-IR (Cataldi et al. 2014; Kral et al. 2016) showed that the bulk of the atomic gas does not originate from the infalling cometary bodies at a few stellar radii, but from a more ex- tended gas disk in Keplerian rotation at several tens of AU around the central star. Molecular gas has been more difficult to detect, with only upper limits on H2 and OH (Martin-Zaıdi et al. 2008; Vidal-Madjar et al. 1994). The presence of CO gas along the line of sight to the star was first marginally detected in ab- sorption by the International Ultraviolet Explorer (Deleuil et al. 1993) and then confirmed through HST observations (e.g. Vidal-Madjar et al. 1994; Roberge et al. 2000). How- ever, detection of its rotational transitions at millimetre wavelengths proved impossible with single dish telescopes, despite very long integration times (Liseau & Artymowicz 1998). It is only through recent interferometric observations with the Atacama Large Millimeter/sub-millimeter Array (ALMA), bringing a drastic improvement in sensitivity and angular resolution, that detection of the J=3-2 transition at 345 GHz has been made possible (Dent et al. 2014). The data revealed for the first time the spatial distribution of CO, pre- senting a clump of emission at 85 AU on the SW side of the disk, which is co-located with both a radial peak in the dust millimetre emission and a SW dust clump similarly observed at mid-IR wavelengths (Telesco et al. 2005). This was inter- preted as evidence for a common production location for both second-generation debris dust and second-generation CO. The clumpy azimuthal structure was then attributed to enhanced collision rates between icy bodies at specific azimuthal locations, which could be the dynamical evidence of either a giant impact between Mars-sized bodies (Jackson et al. 2014) or the migration of a yet unseen planet sweeping icy planetesimals into resonance (Wyatt 2003, 2006). We here present new ALMA follow-up observations of the β Pictoris system. In this work, we focus on CO J=2-1 emission at 230 GHz, observed at an improved sensitivity and spatial resolution of ∼ 5.5 AU, and analyse it together with archival 345 GHz ALMA observations of the J=3-2 transition. We postpone the analysis of the dust continuum and upper limits on SiO emission as the subject of forth- coming work. In Section 2, we describe the observations, calibration and imaging procedures, whereas in Section 3 we analyse the radial, vertical and azimuthal structure of both CO transitions as well as the ratio between the two. In Section 4, we model the resolved line ratios using the non-local thermodynamic equilibrium methods developed in Matr`a et al. (2015), showing how they can be used to probe undetectable species in the disk such as electrons and H2, and measure the mass and optical thickness of CO gas. Fi- nally, in Section 5, we discuss how our analysis impacts the current understanding of gas in the β Pictoris system itself as well as other gas-bearing debris disks. 2 OBSERVATIONS 2.1 ALMA Band 6 We observed the β Pictoris disk with ALMA during its Cy- cle 2 (project code 2012.1.00142.S) using band 6 receivers. Observations were performed using both the 12-m array and the Atacama Compact Array (ACA). The 12-m observations were carried out in two antenna configurations; for the most compact one, observed in December 2013, a mosaic strategy was used with two pointings at ±5(cid:48)(cid:48) from the stellar loca- tion along the disk midplane, whereas for the most extended one, observed in August 2015, a single-pointing strategy was adopted. ACA observations were also carried out in single- pointing mode during October 2013. The on-source times were 28, 114 and 50 minutes respectively for the 12-m (com- pact and extended) and ACA observations. The spectral setup of the correlator consisted of four spectral windows; of these, two were centred around 218.5 and 232.5 GHz and set in time division mode (TDM) to achieve maximum band- width (∼2 GHz each) for continuum observations. The re- maining two were set in frequency division mode (FDM) to target the 12CO J=2-1 line (at rest frequency 230.538 GHz) and the SiO J=5-4 line (at 217.105 GHz) with a channel spacing of 244.141 and 488.281 kHz, respectively. The corre- sponding velocity resolutions are 0.64 and 1.35 km/s (after Hanning smoothing). Across all observations, either Uranus, Ganymede or J0519-454 were used as amplitude calibrators, whereas J0519-4546 was used as phase calibrator and either J0538-4405 or J0334-4008 as bandpass calibrators. The data calibration was carried out using the CASA software version 4.3.0, including appropriate weighting be- tween different observations and/or array configurations. We note that the flux calibration is estimated to be accurate within 10% (ALMA Cycle 2 Technical Handbook). The cali- brated visibilities from the ACA and the 12-m datasets (cov- ering baselines from 9 to 1574 m) were then concatenated, MNRAS 000, 1 -- 21 (2016) ALMA observations of exocometary CO lines in β Pic 3 achieving a u-v coverage which gives us sensitivity to struc- ture on scales between 0.(cid:48)(cid:48)3 and 27(cid:48)(cid:48). To image the CO line, we first subtracted the continuum from the combined visibil- ity dataset using the uvcontsub task within CASA. Then, we imaged velocity channels within ±10 km/s from the radial velocity of the star (20.0±0.7 km/s in the heliocentric refer- ence frame, Gontcharov 2006) using the CLEAN algorithm (Hogbom 1974). Natural weighting of the visibilities was ap- plied, resulting in a synthesised beam of size 0.(cid:48)(cid:48)30×0.(cid:48)(cid:48)26 and position angle (PA, East of North) of -83.◦9, corresponding to 5.8×5.1 AU at the known distance to the system. As one of the main goals of this work is to directly com- pare this new Band 6 CO J=2-1 dataset with the archival CO J=3-2 Band 7 dataset, we also produced a final CO J=2- 1 data cube at the degraded spatial and spectral resolution of the archival J=3-2 dataset; details of the procedure are described in Appendix A. 2.2 ALMA Band 7 In addition to the new Band 6 observations, we retrieved archival ALMA Band 7 CO J=3-2 data obtained within Cy- cle 0 (project code 2011.0.00087.S), which are described in Dent et al. (2014). In summary, a two-point mosaic strat- egy analogous to the Band 6 compact configuration observa- tions was applied, although without use of the ACA (base- lines ranging from 16 to 382 m). For consistency, we re- peated the data reduction and imaging procedure the same way as done for the Band 6 data, using the same version of CASA. This resulted in a final synthesised beam of size 0.(cid:48)(cid:48)69×0.(cid:48)(cid:48)55 (13.4×10.7 AU) and PA 65.◦4, with a channel width of 488.281 kHz (corresponding to a 0.85 km/s res- olution at 345.796 GHz). As for the Band 6 dataset, the uncertainty in the flux calibration is estimated to be 10% (ALMA Cycle 0 Technical Handbook). 3 RESULTS Fig. 1 shows CO J=2-1 and J=3-2 moment-0 CLEAN images of the β Pictoris disk at their original spatial resolution, ob- tained by spectrally integrating over heliocentric velocities between 14 and 26 km/s (±6 km/s in the reference frame of the star). The 1-σ noise levels on the moment-0 images are 2.7 and 24 ×10−23 W m−2 beam−1 (or 3.5 and 21 mJy km/s), respectively, yielding a peak detection of CO inte- grated line emission at a SNR per beam of 17 and 21 for the 2-1 and 3-2 transitions, respectively. The total line flux was measured on the primary-beam-corrected moment-0 im- ages by spatially integrating over a region that contains all significant disk emission, but small enough to avoid signif- icant noise contamination. The integrated fluxes measured are (3.5± 0.4)× 10−20 W m−2 for the J=2-1 transition, and (6.7 ± 0.7) × 10−20 W m−2 for the J=3-2 transition, indi- cating an integrated 3-2/2-1 line ratio of 1.9± 0.3. Absolute errors on integrated line fluxes take into account both the noise level in the moment-0 images and the flux calibration uncertainty; we then summed the relative errors on both fluxes in quadrature (under the assumption that they are Gaussian in nature) to obtain the uncertainty on the line ratio. MNRAS 000, 1 -- 21 (2016) 3.1 Spectrally-integrated radial structure In Fig. 2 we present the radial profile of CO emission along the β Pic disk midplane (xsky axis in Fig. 1). This was de- rived from the moment-0 images by vertically integrating disk emission from pixels within ±20 AU from the midplane (i.e. with ysky < 20 AU). The J=2-1 dataset confirms the presence of a clump of CO on the SW side of the disk (previ- ously discovered in J=3-2 by Dent et al. 2014) at a projected separation of 60 AU from the star, and faint J=2-1 emission detected out to larger distances on the NE side than J=3-2 emission, which will become even clearer when looking at the CO velocity structure (Sect. 3.3). This is attributable to both the lower sensitivity of the Band 7 dataset, but also to an intrinsic decrease in the J=3-2/J=2-1 line ratio at in- creasing disk radii. 3.2 Spectrally-integrated vertical structure In order to probe the vertical structure in the disk, we anal- yse its spine, which we define as the locus of the centres of best-fit vertical Gaussians measured at different locations along the disk midplane. We remind the reader that the disk images shown in Fig. 1 have been rotated by the posi- tion angle (PA) of the main disk observed in scattered light, corresponding to 29.◦3+0.2−0.3 (Lagrange et al. 2012). In scat- tered light, this differs from the PA of the inner part of the disk, which appears tilted when compared to the main disk observed in the outer regions (e.g. Apai et al. 2015). Fig. 3 (top row) illustrates the procedure employed. We apply vertical cuts in the ysky direction for each location xsky along the midplane where emission greater than 5σ is detected. The measured (normalised) flux versus ysky is shown as red and blue solid lines. The green dashed Gaus- sian profiles have width equal to the FWHM of the restoring beam projected onto the ysky direction (11.4 and 5.7 AU for J=3-2 and J=2-1, respectively), and indicate that the disk is resolved vertically for both CO lines. We then employ 1D Gaussians to fit the vertical profiles and obtain the observed best-fit vertical location of the Gaussian peak, which we de- fine as yobs (see fitting procedure and error determination in Appendix B). Repeating the process at several midplane locations yields the locus of vertical Gaussian centres yobs at different xsky locations, i.e. the disk spine. This appears significantly tilted with respect to the PA of the main disk midplane, presenting an extra anticlockwise rotation (as was noted by Dent et al. 2014) by a tilt angle dPA. The latter is similar to the tilt angle observed for the scattered light inner disk (∼4◦, Apai et al. 2015), and is most pronounced at the location of the clump. Accurate measurement of the true warp angle from this sky-projected tilt is challenging. To begin with, the disk spines derived from both CO and scattered light are not well represented by a straight line, meaning that measure- ment of the sky-projected tilt itself is highly sensitive to the disk radii between which the straight line is drawn. Secondly, even assuming a perfectly edge-on disk, there is no reason to believe that the warp axis lies in the plane of the sky; any azimuthal displacement of the warp axis in the orbital plane would cause a substantial difference between the true warp angle and the observed projected sky tilt. To further investigate any potential azimuthal dependence on this tilt, 4 L. Matr`a et al. Figure 1. CO J=2-1 and J=3-2 spectrally integrated (moment-0) CLEAN images of the β Pic disk, at their original synthesized resolution (natural weighting). The images have been rotated by the position angle of the main disk observed in scattered light (29.3◦). This way, we can define the xsky axis as the direction along the disk on-sky midplane (positive towards the SW), and the ysky axis as the direction perpendicular to it (positive towards the NW). The origin of the axis is set to the stellar location, and the restoring beam is shown as the yellow ellipse in the bottom left corner. obs − σ2 H =(cid:112)σ2 standard deviation defined as the scale height H), we can derive H by simply deconvolving σobs by the observed in- strumental resolution projected along the vertical axis, i.e. res. The resulting scale height as a function of midplane location is shown in the bottom row of Fig. 3. Given the near edge-on geometry of the disk, its rather flat appearance is unsurprising, as the scale height observed on- sky at any midplane location will tend to trace the scale height at the disk's orbital outer radius. As such, in Sect. 3.3.1 we add the velocity information from the data cube to retrieve the scale height dependence on the orbital radius and explore its relation to the disk temperature. Figure 2. Radial distribution of CO line flux obtained from spa- tially integrating the moment-0 images in Fig. 1 along the perpen- dicular to the main disk midplane ysky, where(cid:12)(cid:12)ysky (cid:12)(cid:12) < 20 AU. In order to highlight differences in the projected radial structure be- tween the two datasets, the profile for each image was normalised to its maximum. we repeat the same analysis as a function of radial velocity in Sect. 3.3.1. As well as the vertical displacement from the midplane, the same Gaussian fits on the moment-0 images yield ver- tical standard deviations σobs as a function of midplane location. Assuming that the disk is edge-on and that its true vertical structure can be represented by a Gaussian (of 3.3 Velocity information In order to understand the 3D gas disk kinematics and ex- tract its vertical and azimuthal structure, we analyse the CO dataset through position-velocity (PV) diagrams, i.e. maps of different quantities as a function of both position along the disk midplane (xsky axis in Fig. 1) and radial velocity vrad. In particular, we aim to link the (xsky, vrad) PV lo- cation in these diagrams to a (x, y) location in the orbital plane of the disk. This can be done for a given inclination i if we assume that the gas disk is infinitely thin vertically and in Keplerian rotation (see Appendix C for details). In Sect. 3.3.1, we begin by analysing the sky-projected vertical MNRAS 000, 1 -- 21 (2016) -200-1000100200-60-40-2002040600123410-22 W m-2 beam-1CO J=2-1NE-200-1000100200Stellocentric distance (AU)-60-40-200204060Stellocentric distance (AU)01234510-21 W m-2 beam-1CO J=3-2NE3002001000100200300Distance along midplane (AU)20020406080100Integrated line flux (percentage of maximum)CO J=2-1CO J=3-2 ALMA observations of exocometary CO lines in β Pic 5 Figure 3. Vertical structure of the CO disk for the J=2-1 line (blue, left column) and the J=3-2 line (red, right column), from the moment-0 images (Fig. 1). Top: Fitting procedure to the disk's vertical emission profiles as a function of distance along the midplane. Coloured lines represent the measured flux as a function of distance along the vertical (ysky) axis for different midplane locations xsky. The black dashed lines are the best-fits derived, whereas the green dashed lines (as well as the filled ellipses) represent the instrumental resolution, centred on the best-fit Gaussian peaks yobs. Centre: Disk spine, i.e. the vertical offset of Gaussian fits yobs from the top panel as a function of midplane location xsky. The green crosses represent the same spine derived for the dust emission from HST observations (Apai et al. 2015). Bottom: Scale height H as a function of distance along the midplane, derived from the standard deviation σobs of best-fit Gaussians in the top panel. structure along ysky in PV space to derive constraints on the disk vertical and azimuthal structure, assuming a per- fectly edge-on disk (i = 90◦). In Sect. 3.3.2, on the other hand, we interpret the same sky-projected vertical structure along ysky in PV space purely as azimuthal structure for a disk close to, but not perfectly edge-on (i < 90◦). Finally, in Sect. 3.3.3 we carry out a PV diagram comparison between the new CO J=2-1 and the archival J=3-2 dataset. 3.3.1 CO 3D structure in the edge-on assumption In Fig. 4, we show PV diagrams of CO J=3-2 and J=2- 1 line flux obtained by integrating emission vertically (i.e. along the ysky axis in Fig. 1) within 20 AU above and be- low the disk midplane, including all significant disk emis- sion. The spectro-spatial resolution is given by the projec- MNRAS 000, 1 -- 21 (2016) tion of the restoring beam along the disk midplane xsky, combined with the velocity resolution along vrad. The two CO transitions show similar PV structure, consistent with a near edge-on gas disk in Keplerian rotation around a 1.75 M(cid:12) star (Crifo et al. 1997). Diagonal lines on the diagrams represent different radii R in the orbital plane of the disk (Appendix C), which we here assume to be perfectly edge- on; we can use this to constrain the disk's radial extent in the orbital plane to between ∼50 and ∼220 AU. For both transitions, two flux enhancements are observed; one in the resolved SW clump, approaching us at a velocity vrad be- tween 3 and 4 km/s and sky-projected midplane location xsky between 60 and 90 AU. The other enhancement, less pronounced, is found around the projected stellar location and radial velocity (xsky, vrad) ∼ (0, 0). As the latter is not observed towards the star in the moment-0 images, it is likely -150-100-50050100150-50-40-30-20-1001020304050-150-100-50050100150Normalised flux at location along the midplane (AU)-50-40-30-20-1001020304050Vertical distance from midplane (AU)-150-100-50050100150-50-40-30-20-1001020304050-150-100-50050100150Normalised flux at location along the midplane (AU)-50-40-30-20-1001020304050Vertical distance from midplane (AU)-10-50510-10-50510Vertical distance from midplane (AU)12CO J=2-1-10-50510-10-50510Vertical distance from midplane (AU)12CO J=3-2-150-100-5005010015005101520-150-100-50050100150Distance along midplane (AU)05101520Scale height (AU)-150-100-5005010015005101520-150-100-50050100150Distance along midplane (AU)05101520Scale height (AU) 6 L. Matr`a et al. Figure 4. Position-velocity (PV) diagrams of the β Pic disk, showing CO intensity as a function of position along the disk on-sky midplane xsky and radial velocity vrad for each of the two transitions observed. The black asterisk represents the stellar position, while the solid white curves are the maximum radial ve- locity observable in an edge-on Keplerian disk around a 1.75 M(cid:12) star. The yellow rectangle in the bottom left corner represents the spectro-spatial resolution. The dashed white lines represent dif- ferent radii R in the orbital plane of the disk, assuming Keplerian rotation (see Appendix C). due to the larger disk volume per velocity channel probed at low compared to higher radial velocities. Another difference lies in the NE side of the disk, where there is a clear deficit of J=3-2 compared to J=2-1 emission at large radii and low radial velocities (Fig. 4). This suggests that J=2-1 emission is detected extending further out in the disk on the NE side compared to J=3-2 emission. As mentioned above (again, see Appendix C for details), for a disk in circular Keplerian rotation, each PV location (i.e. each (xsky, vrad) spaxel ) in the diagram corresponds to a radius R and hence an (x,±y) location in the orbital plane of the disk. The degeneracy in the sign of y, however, means that each spaxel traces two orbital locations, (x, +y) and (x,−y). Thus, we do not know how much flux belongs to either of the two locations; we only know that the sum of the two fluxes must equal that of the original spaxel. This leads to an infinite number of possible deprojections; Fig. 5 shows the two that are most likely physically plausible, as justified by dynamical models (see Sect. 5.1 and Dent et al. 2014). We obtained these by placing all of the spaxel emission in the NE side (negative xsky) at −y (i.e. in front of the sky plane) and that in the SW (positive xsky) at ei- ther +y (i.e. behind the sky plane, left column in Fig. 5) or −y (right column in Fig. 5). The direction of rotation is known to be clockwise from the sign of the observed ra- dial velocity. The central ±30 AU from the star are masked as our spectro-spatial resolution is not sufficient to deter- mine orbital locations accurately. These deprojected images of CO J=3-2 and J=2-1 line emission reveal the azimuthal structure of the disk; on the SW side, the clump peaks at φ ∼ ±32◦ (measured from the positive xsky direction), with a tail of emission extending to either the front or the back of the star along the line of sight direction. Depending on the system configuration, emission on the NE side can be interpreted either as the continuation of the tail originating from the SW clump (right column), or as a separate dimmer clump at φ ∼ +212◦ with its own tail also extending into the line of sight to the star (left column). In Section 3.2 we analysed the spectrally-integrated disk vertical structure along ysky as a function of position along the disk midplane xsky. Given our velocity information, how- ever, it is more insightful to measure the disk spine and width separately for each of the radial velocities vrad at which the disk is detected, using the same procedure as out- lined in Appendix B. This yields a measurement of the sky- projected scale height H and of the disk vertical offset from the main disk midplane yobs as a function of position along the midplane xsky and radial velocity vrad. In other words, we obtain a PV diagram (shown for the J=2-1 transition in Fig. 6). The anti-clockwise tilt (dPA=PAobs-PAmain disk) ob- served in the moment-0 images (Fig. 3) is also present throughout the radial velocity channels, with CO at neg- ative radial velocities (SW) being vertically displaced above CO at positive velocities (NE). Some substructure is ob- served, with an enhanced vertical displacement at low radial velocities (hence large orbital radii) on the NE side, and a decreased displacement at high radial velocities (hence small orbital radii) on the SW side. In addition, a positive vertical offset of disk emission along the line of sight to the star (en- hanced at positive velocities) is present. These local features are marginally significant at the typical 0.5-1 AU 1σ uncer- tainty on vertical offsets in each spaxel; their interpretation in terms of a putative 3D warp structure requires detailed dynamical modelling, and is beyond the scope of this work. The scale height also presents significant PV structure in the form of a gradient from low values at high radial ve- locities to high values at lower radial velocities (Fig. 6, bot- tom). For the assumed perfectly edge-on configuration, and given that assuming Keplerian rotation different diagonal lines represent different orbital radii, this gradient is repre- sentative of an increase in the disk scale height as a function of orbital radius. Indeed, if we assign an orbital radius R to each PV location (Appendix C) and assume the disk scale height to be azimuthally symmetric, in Fig. 7 (top) we show that the scale height H scales as (cid:18) R (cid:19)0.75+0.02−0.02 85AU H = 7.0+0.6−0.6 × . (1) The error bars (1σ) in Fig. 7 (top) represent the uncertainty in the derivation of both the scale height H and the orbital radius R. The error on the scale height was calculated from MCMC fits as described in Appendix B, whereas the error on the orbital radius R was propagated from the uncertainty on the PV location (assumed to be equal to the size of a spaxel). Assuming the CO disk to be vertically isothermal and in hydrostatic equilibrium, the scale height can then be used to trace the disk temperature at a certain orbital radius MNRAS 000, 1 -- 21 (2016) -200-1000100200Distance along midplane (AU)-6-4-20246Radial velocity (km/s)0.00.51.01.52.02.53.010-1 Jy beam-1 channel-150AU220AUCO J=2-1-200-1000100200Distance along midplane (AU)-6-4-20246Radial velocity (km/s)024681010-1 Jy beam-1 channel-150AU220AUCO J=3-2 ALMA observations of exocometary CO lines in β Pic 7 Figure 5. CO emission in the (x, y) orbital plane of the disk, derived from the(cid:0)xsky, vrad (cid:1) information in the PV diagrams for the J=2-1 transition (top) and for the J=3-2 transition (bottom), under the assumption of a perfectly edge-on disk. The two columns represent two plausible deprojections of the PV diagrams, justified by the dynamical scenarios proposed in Dent et al. (2014). The asterisk represents the location of the star, with the inner and outer dashed circles representing deprojected radii of 50 and 220 AU. The 'fishbone' appearance is caused by the finite velocity resolution of the dataset, with any radial velocity tracing a characteristic curve in the displayed (x, y) space. The central ±30 AU have been masked due to all y locations approaching zero radial velocity along the line of sight to the star, rendering the deprojection highly degenerate. through T = GM∗µ H2 R3 , where G is the gravitational con- kNA stant, M∗ is the mass of the star, µ is the mean molecular mass of the gas, k is Boltzmann's constant, and NA is Avo- gadro's number. We here assume the gas to be dominated by the carbon and oxygen atoms released from CO pho- todissociation, giving a mean molecular mass µ = 14 (see Sect. 5.3). We can then use our measured dependence of the scale height on radius to derive the radial dependence of the gas temperature in the β Pic disk (Fig. 7, bottom). This decreases as a function of radius, scaling as (cid:18) R (cid:19)−1.58 T =(cid:0)2.1+0.4−0.4 × 102(cid:1) × 85AU (2) where T is in K and R is in AU. No temperature in- crease/decrease is seen at the clump location compared to MNRAS 000, 1 -- 21 (2016) the rest of the disk, which is in line with the expectation that the CO, after release, should quickly collide and couple to the rest of the atomic disk (again, see discussion in Sect. 5.3). 3.3.2 CO 3D structure for a non-edge-on configuration In the previous section, we neglected a potential deviation of the disk inclination from perfectly edge-on, which has been proposed before through modelling observations of op- tical and near-IR scattered light from the dust disk (e.g. Ahmic et al. 2009; Milli et al. 2014). If the disk is inclined from edge-on, the vertical structure becomes coupled with the azimuthal structure. On one hand, this means that our derivation of the disk scale height and temperature in the -200-1000100200Distance along midplane (AU)-200-1000100200Distance along l.o.s. (AU)0.51.01.52.02.53.0S2-1 (Jy beam-1 channel-1)12CO J=2-1-200-1000100200Distance along midplane (AU)-200-1000100200Distance along l.o.s. (AU)0.51.01.52.02.53.0S2-1 (Jy beam-1 channel-1)12CO J=2-1-200-1000100200Distance along midplane (AU)-200-1000100200Distance along l.o.s. (AU)246S3-2 (Jy beam-1 channel-1)12CO J=3-2-200-1000100200Distance along midplane (AU)-200-1000100200Distance along l.o.s. (AU)246S3-2 (Jy beam-1 channel-1)12CO J=3-2 8 L. Matr`a et al. Figure 6. Position-velocity diagrams of both the CO J=2-1 disk vertical offset yobs with respect to the main dust disk (top) and of the disk scale height H (bottom), for an assumed perfectly edge- on disk. These measurements were obtained through Gaussian fitting of measured fluxes along cuts perpendicular to the disk midplane (as displayed in Fig. 3) repeated for each radial velocity channel. previous section may be in part biased by the edge-on as- sumption. On the other hand, we can interpret the vertical displacement yobs of the disk from the midplane (i.e. the disk spine) purely as an effect of azimuthal structure seen at an inclination i < 90◦. Using the observed vertical displacement yobs measured at each PV location (xsky, vrad) (see Fig. 6 top) we can de- termine whether CO emission originates in front or behind the plane of the sky (i.e. at + or −y in Fig. 5). This is because for a given orbit, characterised here by its inclina- tion from face-on i and its on-sky tilt angle dPA, an on-sky location (xsky,orb, +ysky,orb) at radial velocity vrad will cor- respond to either orbital location (x, +y) behind the sky plane or (x,−y) in front of the sky plane. Then, in the pres- ence of a ±y asymmetry in the orbital plane of the disk, the flux contribution from the two on-sky vertical locations +ysky,orb and −ysky,orb will differ and the vertical centroid yobs of the CO emission will be displaced from the mid- plane. As described in Appendix D, we use this vertical dis- placement to infer the level of CO emission originating from (xsky,orb, +ysky,orb) as opposed to (xsky,orb,−ysky,orb) in the sky plane. Figure 7. Top: Measured dependence of the CO disk scale height on orbital radius, derived from the PV diagrams in Fig. 6 (bot- tom) assuming Keplerian rotation. Each point corresponds to a spaxel in the J=2-1 PV diagram. Bottom: Radial dependence of the temperature derived from the scale height under the assump- tion of a vertically isothermal disk. In both plots, the red lines are constructed by randomly picking 1000 values of our fitting parameters (intercept and slope in log-log space) from their joint posterior probability distribution. The best-fit power law coeffi- cients are displayed in the top right of the diagrams, with their respective 1σ uncertainties. The blue lines represent model pre- dictions by Kral et al. (2016). A degeneracy remains in that we do not know which be- tween +ysky,orb and −ysky,orb is in front or behind the plane of the sky. This may be solved using scattered light imaging, since material in front of the star will appear brighter than behind the star, owing to the phase function of the grains be- ing peaked towards and hence favouring forward-scattering angles (e.g. Milli et al. 2015). For β Pictoris, scattered light emission above the midplane is seen to be brighter than be- low the midplane (e.g. Golimowski et al. 2006); that sug- gests that dust at +ysky lies preferentially in front of the star, whereas dust at −ysky is located behind the star. How- ever, this does not take into account that in the presence of an azimuthal asymmetry there may be an excess of material above the midplane, which would affect the above argument by leading us into interpreting the observed flux excess as a phase function effect. As the method produces a model data cube for a given orbit, we can compare different orbits and hence find how well different (i, dPA) combinations fit the data (see Ap- pendix D and χ2 map in Fig. D2). We note once again that this approach does not take into account any intrinsic disk vertical structure (and in particular its dependence on or- bital radius) nor any uncertainty in the determination of the MNRAS 000, 1 -- 21 (2016) ALMA observations of exocometary CO lines in β Pic 9 • The majority of the flux from the CO clump on the SW side of the disk is located above the sky-projected disk mid- plane for any reasonable choice of i (≥ 86) and dPA (< 5◦). This confirms that the flux in the sky-projected clump does originate from a physical clump, located either behind or ahead of the sky plane (due to the aforementioned degener- acy).• The disk deprojection obtained for a lower inclination of 86◦ and no dPA qualitatively reproduces well the resonance sweeping scenario proposed in Dent et al. (2014) and based on previous work by Wyatt (2003, 2006), with two clumps on opposite sides of the star and their respective trailing tails (see discussion in Sect. 5.1). Further 3D modelling taking into account at the same time the disk vertical and azimuthal structure as well as its geometry is needed to rule out or confirm this possibility. 3.3.3 CO 3-2/2-1 line ratio In order to better investigate morphological differences be- tween the two transitions, we combined the J=2-1 and J=3-2 PV diagrams to obtain line ratios in each position-velocity spaxel. As such ratios can be extreme in noise-dominated re- gions, we facilitate visualisation by only displaying ratios in spaxels where either CO J=3-2 and J=2-1 emission is over 4σ (where σ is the RMS noise level measured in each PV dia- gram). A detection at 3-2 but only an upper limit at 2-1 will then lead to a lower limit on the line ratio, and conversely a detection at 2-1 but not at 3-2 will lead to an upper limit on the ratio. Such line ratio PV diagram is shown in Fig. 9 (left), with the corresponding uncertainty diagram in Fig. 9 (right). At each spaxel, the uncertainty is the quadratic sum of the relative errors on the flux for the two original PV dia- grams, where these are given by the RMS levels. We here do not take into account the 10% flux calibration error on each dataset, since we are here most interested in the presence of gradients rather than in the absolute value of the ratios. The line ratio PV diagram shows a noticeably different structure to the flux PV diagrams for the individual transi- tions in Fig. 4. In particular, the line ratio shows no evidence of the CO clump nor the NE/SW asymmetry. On the other hand, it shows a significant gradient where the line ratio is seen increasing from low values at low radial velocities to high values at high radial velocities, for any given on-sky midplane location. On the NE side, where individual line fluxes are generally dimmer than in the SW side (see Fig. 4), this causes a non-detection of the 3-2 line at low veloci- ties (as illustrated through the upper limits on the ratios). This then explains the differences between the two lines that were already apparent from the individual moment-0 images and PV diagrams (Fig. 1 and 4), with 2-1 emission lying at larger projected radii than the 3-2 line on the NE side of the disk. The physical significance of this gradient is explored through modelling in Sect. 4 below. 4 LINE RATIO MODELLING Through detailed analysis of resolved ALMA CO J=3-2 and CO J=2-1 line observations, we unveiled the radial, vertical and azimuthal structure of CO line emission in the β Pictoris Figure 8. Face-on deprojection of CO disk emission, derived from the velocity information in the PV diagrams for the J=2-1 tran- sition, for two cases of disk inclination i and on-sky tilt angle dPA. Flux at each x location has been divided between + and -y locations in the orbital plane displayed here using the vertical displacement yobs of disk emission in the plane of the sky (Fig. 6, top), using the method described in Appendix D. Though the (i, dPA) choice for the deprojection in the top panel is formally the best-fitting in the framework of our simple model (see main text for details), the image in the bottom panel reproduces well the expected morphology from resonant trapping of planetesimals due to an outward migrating planet. vertical displacement yobs from the on-sky midplane. Keep- ing this in mind, we find that the formal best-fit (i.e. the χ2 minimum) is at i ∼ 88◦ and dPA ∼ 3◦, though we consider all inclinations ≥ 86◦ and on-sky tilt angles ≤ 5◦ to provide reasonable fits. Of course, for any choice of (i, dPA), we can use our method to produce a map of emission in the orbital plane of the disk, as we now have determined what fraction of emission belongs to +y and to −y for a given x. Fig. 8 shows two examples, for our formal best-fit orbit (i = 88◦, dPA = 3◦) and in the case of a lower inclination (i = 86◦) and no on-sky tilt (dPA = 0◦). We report two main findings: MNRAS 000, 1 -- 21 (2016) 10 L. Matr`a et al. Figure 9. Left: CO J=3-2/J=2-1 line ratio position-velocity diagram of the β Pic disk. Only spaxels where either CO transition is detected (at ≥4 σ) are shown; downward arrows indicate spaxels where only the J=2-1 transition was detected. The white asterisk represents the stellar position and velocity, while the solid white curves are the maximum radial velocity observable in an edge-on Keplerian disk around a 1.75 M(cid:12) star. The yellow rectangle in the bottom left corner represents the spectro-spatial resolution. Right: Percentage error map of the line ratio PV diagram, where this can only be quantified in spaxels where both transitions are detected. disk. Under the assumption that the system is viewed per- fectly edge-on, we interpreted the scale height observed as a function of orbital radius to give us an estimate for the tem- perature of the gas and its radial dependence. We then re- laxed the edge-on assumption and interpreted the observed vertical structure as solely caused by azimuthal structure, in order to break the degeneracy in deprojecting PV dia- grams to the disk orbital plane. Finally, we analysed the CO 3-2/2-1 line ratio PV diagram and discovered a signifi- cant gradient with the ratio increasing with radial velocity at any given midplane location. In this section, we aim to use the non-local thermodynamic equilibrium (NLTE) anal- ysis of CO excitation developed in Matr`a et al. (2015) to interpret these line observations (particularly resolved line ratios) in terms of the physical conditions, origin and mass of CO in the β Pic disk. 4.1 Second generation: NLTE line ratios as a probe for electron densities We begin by assuming CO line emission to be optically thin at the wavelengths of the observed transitions, and double- check this assumption later in Sect. 4.3. In this regime, CO excitation and hence line ratios are only dependent on the kinetic temperature Tk and the collisional partner density ncoll, for the known external millimetre radiation field Jν at the frequency of the transitions in question. In a scenario where the gas is of second-generation origin, the dominant collisional partners for which collisional rates are available are electrons, which originate from the photoionisation of atoms created by rapid CO photodissociation. In the ab- sence of H2, electrons are more efficient colliders than H2O (Matr`a et al. 2015), but also H i, since the latter, despite be- ing potentially ∼6 times more abundant than electrons (Kral et al. 2016), is on average a much less efficient collider. For example, collisional de-excitation of a CO molecule from ro- tational level 3 to 2, for the given H i/e− abundance of 6 happens at a rate that is (γ3−2,H inH i)/(γ3−2,e− ne− ) ∼ 50 times faster for electron collisions than for H i collisions1. Nonetheless, it is useful to keep in mind that other less domi- nant species may contribute to observed CO collisional exci- tation, which implies that the electron densities probed here should strictly speaking be considered upper limits. Fig. 10 (top) shows the J=3-2 / J=2-1 line ratios ex- pected for CO that is excited by collisions with electrons and by millimetre radiation from the cosmic microwave back- ground (CMB, which dominates over the dust continuum for low CO transitions), as a function of the electron density ne− and the gas kinetic temperature Tk. Each ratio traces a line in (Tk, ne− ) space, since these are the only free parame- ters. For ne− , we explore the parameter space between 10−1 and 106 cm−3, encompassing the transition between the two limiting excitation regimes (LTE and radiation-dominated, Matr`a et al. 2015). For the kinetic temperature of the gas, we probe a broad range of Tk between 5 and 2000 K to illustrate the behaviour of the line ratios. Its true value depends on the detailed balance between local heating and cooling pro- cesses in the gas disk. Our temperature estimates from the CO scale height (Sect. 3.3.1) suggest values between 40 and a few hundred K between 50 and 200 AU, whereas detailed thermodynamical modelling of the disk suggests values be- tween 30 and 80K at the same radii (Kral et al. 2016). In any case, we expect the gas kinetic temperature to be above a few tens of K throughout the disk. These predictions can then be compared with our mea- surements, tracing lines overplotted on Fig. 10 (top). Above 1 Collisional rate coefficients γ were obtained from Dickinson & Richards (1975) for electrons and from Walker et al. (2015) for H i MNRAS 000, 1 -- 21 (2016) -200-1000100200Distance along midplane (AU)-6-4-20246Radial velocity (km/s)12345678950AU220AU12CO 3-2/2-1-200-1000100200Distance along midplane (AU)-6-4-20246Radial velocity (km/s)05101520253050AU220AU12CO 3-2/2-1 % Error ALMA observations of exocometary CO lines in β Pic 11 ne− is only obtainable with two (or more) line ratios, our measurements can confidently constrain the electron density in the disk to around 102-103 cm−3. This allows us to attribute the gradient of increasing CO line ratio with radial velocity observed in the PV dia- grams (Fig. 9, left) to the electron density distribution in the disk. Since we want to measure the electron density as robustly as possible, we mask the highest 10% of the line ratios (i.e. those above 3.80), which arise preferentially from high-noise regions of the PV diagram, and have a stronger kinetic temperature dependence. For the remaining line ra- tios, this temperature dependence is much weaker, so any assumption we make is not going to affect the result signif- icantly. As such, we assume the temperature radial profile predicted by Kral et al. (2016), i.e. a power law with slope of −0.8, normalised to a value of 80 K at 50 AU. Using the temperature radial profile derived from our scale height measurement in Fig. 7 (bottom) yields analogous results. Each location in the PV diagram traces a certain or- bital radius within the disk (Appendix C). This way, we can solve for the radial dependence of the electron density, which is shown in Fig. 11, left. The uncertainty on ne− was propagated from that of the line ratio, assuming a perfectly known Tk from the models; the uncertainty on the radius, on the other hand, was calculated from the uncertainty in the (xsky, vrad) location within a spaxel in the PV diagram. Since we are most interested in the radial dependence rather than on the absolute value of the electron density, we have not included sources of systematic error that may shift all of the observational points in our plot (e.g. an error on M∗ or i may cause an overall shift in the radial direction, whereas the ALMA flux calibration systematic may cause an overall shift in the ne− direction). The electron density in the disk shows a power law ra- dial dependence between 40 and 200 AU (i.e. where the CO is detected), with a slope of γ ∼ −1; this is in line with the prediction by the Kral et al. (2016) model, which we will discuss in more detail in Sect. 5.3. 4.2 Primordial origin: H2 densities from line ratios In the previous section, we derived constraints on the density of electrons using the CO line ratios, where we assumed a gas disk of secondary origin, as already suggested by the CO morphology and short survival timescale (Dent et al. 2014). However, we note that the disk could still contain large amounts of unseen H2, which may itself shield CO and allow it to survive since the protoplanetary phase of evolution. This would imply a primordial origin for the gas disk. In such scenario, H2 dominates the disk gas mass and will act as the dominant collider species for CO excitation. Then, we can use the same line ratio analysis as in the previous subsection to estimate the H2 density. Fig. 10, bot- tom shows the dependence of the 3-2/2-1 line ratios mea- sured on Tk and nH2 . Though once again for the highest line ratios the temperature dependence becomes stronger and we cannot make an accurate prediction, the large majority of the measured ratios have a weak temperature dependence. We can therefore use them to estimate the H2 density, and use the PV diagram to once again obtain the radial depen- dence of nH2 , shown in Fig. 11, right. We find that if H2 were Figure 10. Colour maps of the J=3-2/2-1 line ratio expected in an optically thin disk as a function of the density of the dominant collider species and the kinetic temperature of the gas. The dom- inant collider species is electrons in a secondary origin scenario (top), and H2 in a primordial origin scenario (bottom). Contours represent minimum and 90th percentile values observed in our ratio PV diagram (Fig. 9) kinetic temperatures of ∼30 K, the vast majority (up to ∼90th percentile) of the line ratios observed in the β Pic disk (from spaxel values in Fig. 9) are very good probes of the electron density, being largely independent of the choice of kinetic temperature. On the other hand, if the electron densities were high enough for CO to reach local thermody- namic equilibrium (LTE, right edge of the colour map), the line ratios would be higher and become solely dependent on the kinetic temperature. While a unique solution for Tk and MNRAS 000, 1 -- 21 (2016) 10-110010110210310410510610100100010-1100101102103104105106ne- (cm-3)101001000Tk (K)10-11001011021031041051061010010000.51.01.52.02.53.03.54.04.55.05.56.06.57.07.5FJ=3-2/FJ=2-1min = 1.0490% = 3.80CO sublimation temperature100101102103104105106107101001000100101102103104105106107nH2 (cm-3)101001000Tk (K)1001011021031041051061071010010000.51.01.52.02.53.03.54.04.55.05.56.06.57.07.58.08.59.0FJ=3-2/FJ=2-1min = 1.0490% = 3.80CO sublimation temperature 12 L. Matr`a et al. Figure 11. Electron density (left) and H2 density (right) as a function of deprojected radius derived from the PV ratio diagram in Fig. 9 using the NLTE analysis as displayed in Fig. 10, and assuming temperatures predicted by the hydrodynamical model of Kral et al. (2016). Line ratios higher than 3.8 were masked due to their strong dependence on kinetic temperature, which only adds noise to the diagram. For the remaining line ratios, our temperature choice does not influence the resulting electron densities strongly. Error bars in both directions were derived from uncertainties in the line ratio and deprojected radial location. to make up the bulk of the gas disk, its density would be between 103 and 105 cm−3 at radii between 50 and 200 AU. Strictly speaking, these values should be considered upper limits, as the presence of other collider species, here unac- counted for, would lower the derived H2 densities necessary to maintain the observed level of CO excitation. The H2 density levels we derive are much below values observed and expected from theoretical models for protoplanetary disks (e.g. Boneberg et al. 2016), proving that the gas content of the β Pic debris disk is intrinsically different in molecular hydrogen content to a typical primordial disk, and is insuf- ficient to allow CO survival over the system's lifetime (see discussion in Sect. 5.4). 4.3 Optical thickness and total CO mass Analysing the 3-2/2-1 line ratio also allows us to set im- proved constraints on the total CO mass in the system. Dent et al. (2014), having only information on the J=3-2 integrated line flux, obtained a value within a factor 2 of 2.85×10−5 M⊕, assuming optically thin emission and exci- tation temperatures Texc between 20 and 85 K. We hereby test both these assumptions and their implications on the derived CO mass. For optically thin emission, the excitation temperature corresponds to the kinetic temperature only if LTE applies; in NLTE the discrepancy between the two can be significant, with Texc << Tk. In fact, our measured disk-integrated line ratio of 1.9 ± 0.3 corresponds to an excitation temperature Texc = 12 ± 4 K, which may appear low, but is entirely con- sistent with the expectation of NLTE. In order to calculate the CO mass from a given CO integrated line flux (we here choose the J=2-1), we need to know the fractional popula- tion of the upper level of the transition, where this can be obtained for each of the (Tk, ne− ) pairs traced by the con- tour of our disk-integrated line ratio in Fig. 10 (top). The intrinsic (Tk, ne− ) degeneracy implies that a range of CO masses will be possible. We estimate the range of possible CO masses us- ing Monte-Carlo methods. We assume both integrated line fluxes to follow a Gaussian probability distribution with standard deviation equal to the quoted uncertainty (which includes the absolute flux systematic added in quadrature to the noise level from the cube). For each integrated line flux, we sample this Gaussian 104 times and for each sample cal- culate the line ratio, which will itself be sampled 108 times. For each line ratio we calculate possible (Tk, ne− ) pairs (i.e. draw a contour on Fig. 10, top) and randomly draw one of these pairs, assuming a uniform underlying (Tk, ne− ) distri- bution. This pair will yield a value for the fractional popu- lation of the upper level of the transition in question (x2), which then, combined with a sample of the integrated line flux (F2−1), yields a sample for the CO mass (through Eq. 2 in Matr`a et al. 2015). Repeating this procedure for the large number of samples drawn from the measured distributions of integrated line fluxes yields a probability distribution for the total CO mass in the disk. The CO mass range obtained is 3.4+0.5−0.4 × 10−5 M⊕, representing the (50+34−34)th percentiles of this distribution. Low line ratios and hence excitation temperatures, such as observed in β Pic, can also be symptomatic of optical thickness of the CO line (e.g. Flaherty et al. 2016). Opti- cal thickness can be tested through observations of CO iso- topologues such as 13CO and C18O, assuming interstellar isotopic ratios (e.g. K´osp´al et al. 2013); there is however no direct measurement of these at millimetre wavelengths in β Pic. Another possibility is to estimate it directly through its definition, τν = hν 4π∆ν (xlBlu − xuBul)N, (3) where h is Planck's constant, ν is the frequency of the tran- sition in Hz, ∆ν is the line width in Hz (a rectangular line MNRAS 000, 1 -- 21 (2016) 100102103104100Radial distance from star (AU)102103104ne- (cm-3)Kral et al. 2016ALMA observations100103104105100Radial distance from star (AU)103104105nH2 (cm-3) ALMA observations of exocometary CO lines in β Pic 13 profile is assumed), xu and xl are the fractional populations of the upper and lower energy level of the transition, respec- tively, N is the CO column density along the line of sight in m−2, and Blu and Bul are the Einstein B coefficients for the upward and downward transition. In our case, we measure xu, xl and N under the optically thin assumption, and use them to verify that the optical thickness is indeed < 1. We follow this procedure in the PV spaxel where the CO flux is highest, and hence where we would expect the optical thick- ness to be highest. This corresponds to the location of the clump at a radial velocity of −3.5 km/s and midplane loca- tion of 75 AU to the SW of the star. The line ratio measured here is 2.3 ± 0.3; from this, we can derive possible (Tk, ne− ) combinations, and corresponding combinations of fractional populations x3 and x2. The column density in the spaxel can then be estimated through a modified version of Eq. 2 in Matr`a et al. (2015), N = 4πd2 hνu,lAu,l∆A Fu,l xu , (4) where d is the distance to the star, Au,l is the Einstein A coefficient of the transition, Fu,l is the integrated line flux observed, and ∆A is the on-sky area of the line-of-sight col- umn. Applying this column density to Eq. 3, we obtain an optical thickness τ3−2 = 0.27 ± 0.05 for the 3-2 line at the clump location, where the error takes into account both the intrinsic (Tk, ne− ) degeneracy and the observational uncer- tainties on the observed fluxes F3−2 and F2−1. This value of the optical depth τ3−2 derived at the clump location can be interpreted as an upper limit to the optical depth, which is likely lower in the rest of the disk, showing that our assump- tion of optically thin CO emission is a good approximation to the physical conditions in the β Pic disk. This is also sup- ported by the absence of the SW clump in the PV diagram of the CO line ratio, as opposed to the CO J=3-2 or J=2-1 PV diagrams; this indicates that CO excitation is indepen- dent of the CO column density and hence no optical depth effects are at play. It is valuable to note that increasing the clump CO den- sity by an order of magnitude would have resulted in CO millimetre emission being optically thick along the line of sight to Earth. A similar argument applies for the vertical optical thickness of CO to UV light, which allows CO to self-shield against photodissociation and hence prolong his survival timescale in the disk (see Sect. 5.2.2, and Visser et al. 2009). Despite the low levels observed, if the total CO mass or the clump CO density were only an order of mag- nitude higher, the CO would have survived longer than an orbital timescale at 85 AU, significantly reducing the az- imuthal asymmetry observed and casting significantly more uncertainty on its secondary nature. This suggests that sig- nificant optical thickness in a second generation gas disk can be easily attained for CO production rates slightly higher than observed in β Pictoris; as such, observations of opti- cally thick, azimuthally symmetric CO in debris disk sys- tems (e.g. HD141569A and HD21997, White et al. 2016; K´osp´al et al. 2013), should not be treated as sufficient proof for a primordial disk origin. In such cases, multi-transition observations of optically thin isotopologues are necessary to probe the H2 densities in the disk (Sect. 4.2), and discern between the primordial and secondary origin scenarios (Sect. 5.4). MNRAS 000, 1 -- 21 (2016) 5 DISCUSSION 5.1 Radially resolved CO clump: pointing towards the resonant migration scenario Dent et al. (2014) brought forward two possible dynamical scenarios to explain the origin of the clumpy morphology in the CO disk. The first (explaining an azimuthal morphol- ogy like shown in Fig. 5, right column) is a recent impact between planetary bodies, releasing dust and planetesimals that once created follow a range of orbits around the star. Crucially, all these orbits have to cross at the original im- pact location, where they cause enhanced collisional activity (Jackson et al. 2014; Kral et al. 2015), which is very confined spatially. This location is fixed in time, so dust and/or gas density enhancement produced should also remain station- ary. The second scenario (explaining an azimuthal morphol- ogy like shown in Fig. 5, left column, or Fig. 8, bottom) con- sists of a planet migrating outwards and trapping planetes- imals in its 2:1 and 3:2 resonances (Wyatt 2003, 2006). The accumulation, or outward sweeping, of these planetesimals at the resonance locations causes once again an increase in collisional activity, causing enhanced dust and/or cometary gas production. Due to the shape of the resonant orbits of the planetesimals, the azimuthal structure predicted consists of two clumps of unequal brightness at ±90◦ azimuth with respect to planet, which in Fig. 8 (bottom) would be located at ∼+302◦ anticlockwise with respect to the positive x axis. These clumps can be quite extended radially, depending on the eccentricities of the planetesimals that are trapped into resonance. The higher these eccentricities, which correspond to a larger extent of planet migration, the broader the re- gion of enhanced collisional activity (see e.g. Fig. 6 in Wyatt 2003). In addition, the clumps are expected to move, rotat- ing with the orbital period of the planet. The improved sensitivity and resolution of the new Band 6 dataset allowed us to detect CO J=2-1 emission be- tween ∼50 and ∼220 AU. At the azimuthal location of the brightest SW clump spaxel (Fig. 5), CO emission is detected over a range of at least ∼100 AU in radius, where this ex- tent is clearly resolved at the resolution of the observations (∼5.5 AU along the x axis). This can mean two things: ei- ther CO is released over a range of radii, favouring eccentric planetesimals trapped in resonance in the planet migration scenario, or CO is released at a point-like location in the giant impact scenario and can spread radially in a timescale much shorter than the orbital timescale. In order to understand the dynamics of CO, we need to understand its surrounding environment. Molecules re- leased from cometary ices (e.g. CO itself, see next section) are short lived due to the short photodissociation timescales, and go on to form an accretion disk that is dominated by their long-lived atomic photodissociation products (see Sect. 5.3). Therefore, production of fresh CO will happen within the atomic gas disk that is already in place. As such, re- gardless of the dynamics of its release, any CO produced will rapidly collide and couple to other species in the disk, losing the dynamical 'memory' of its production and follow- ing the atomic gas in Keplerian rotation around the star. To verify this, we estimate the collision timescale between CO and C ii, which is created via photoionisation of C i pro- duced by CO photodissociation. This can be estimated as 14 L. Matr`a et al. τCO−C ii = (σnv)−1, where n is the number density of the C ii gas, v is the relative velocity of the collision, and σ is the cross-section of the ion-neutral collision. We estimate the latter through (McDaniel 1964; Beust et al. 1989) (cid:115) σCO−C ii = πq2α 0µ(cid:48) 1 v (5) where q is the charge of the ion, α is the polarisability of CO, 0 is the vacuum permittivity, µ(cid:48) is the reduced mass of the system, and v is again the relative impact velocity. For the C ii number density, we adopt the value predicted at 85 AU by Kral et al. (2016) (∼150 cm−3), whereas the relative velocity cancels out through the cross-section expression. The timescale for CO-C ii collision obtained is 0.19 years, much shorter than both the orbital and photodissocia- tion timescales at 85 AU (∼600 and ∼300 years, respectively; see Sect. 5.2.2). Given that there are many more gas species present than just C ii, this should be taken as an upper limit on the collision timescale. This means that CO will quickly couple to the surrounding gas, proceeding in Keplerian ro- tation. It also means that to reproduce the observed radial width, if released at a point-like production location, CO must spread to a width of ∼ 100 AU before colliding with the surrounding atomic gas. For a CO molecule to travel ∼ 50 AU in less than 0.19 years, under the simplifying as- sumption of a constant expanding velocity, it would have to be released (and travel) at an implausible velocity of at least ∼1200 km/s. Since such high CO velocities are not observed, a point-like production location such as predicted in the gi- ant impact scenario is ruled out. In turn, this implies that CO needs to be produced at a wide range of radii, leaving resonant sweeping due to outward planet migration as the only dynamical scenario proposed to date that can viably explain the observed CO morphology. 5.2 The ice content of β Pic's exocomets 5.2.1 Secondary molecular gas production The augmented collisional activity produced by a dynam- ical scenario such as resonant trapping due to planet mi- gration favours release of volatiles from cometary ices, ei- ther through direct release from cometary break-up (Zucker- man & Song 2012) or UV photodesorption (Grigorieva et al. 2007). Thermal desorption (i.e. sublimation) of CO can be ruled out since we are far within the system's CO snow line, meaning that CO cannot last on the ice surface. On the other hand, as much as 50% of the original CO ice content can remain trapped in other ices, e.g. water ice (Collings et al. 2003), and can therefore be released with them. Then, in the presence of a collisional cascade, volatile release is favoured through the collisions themselves or through the exposure of fresh ice that is then subject to rapid UV photodesorption. Grigorieva et al. (2007) analyse the latter effect in detail, showing that in the β Pic disk the desorption timescale of water ice (and hence CO with it) is shorter than the col- lisional timescale only for particles of size below ∼20 µm. Hence, in the absence of direct collisional release, the ma- jority of the ice would be retained through the top part of the collisional cascade and released mostly at the bottom of it. Assuming water and CO ice to be well mixed (which should be a good approximation as long as gas release occurs from undifferentiated bodies of size below tens of km), they will be released into the gas phase at the same production rate through the processes discussed above. For a collisional cascade at steady state, we can estimate the rate at which solid mass (i.e. volatiles plus rock) is being lost from the cascade itself through Eq. 15 and 16 in Wyatt (2008). We take the same assumptions as in Sect. 4.4 of Marino et al. (2016), with the additional assumptions of a uniform colli- sion rate in a ring of planetesimals between 50 and 150 AU (Dent et al. 2014) with a fractional luminosity Ldust/L∗ of 2.1× 10−3 around a star of 8.7 L(cid:12) (from best-fit of the SED, Kennedy & Wyatt 2014). The resulting total mass loss rate from the planetesimal belt is 1.9 M⊕/Myr (or 1.2 × 1019 kg/year), where most of this release will happen in regions of higher planetesimal mass concentration, since the above loss rate scales with (Ldust/L∗)2 (and hence with M 2 belt). Using the fraction of this solid mass that is composed of CO ice we can obtain the release rate of CO gas, regardless of the choice of release mechanisms discussed above. We note that there is a considerable uncertainty on this number, since this calculation relies on assumptions about the unknown planetesimal strength, mean planetesimal ec- centricities, and particle velocities (e.g. Marino et al. 2016). In addition, the resonant scenario invoked to explain the clumpy CO morphology would imply a much higher mass loss rate than calculated above, particularly at the clump location. As such, we interpret the value of 1.9 M⊕/Myr as a lower limit estimate to the solid mass loss rate in the disk. 5.2.2 Secondary molecular gas destruction On the other hand, the dominant destruction mechanism is photodissociation from the interstellar UV radiation field (ISRF), which will dominate over the stellar UV light at all radii where CO is detected in the β Pic disk (Kamp & Bertoldi 2000). The photodissociation timescale in an envi- ronment that is optically thin to the UV ISRF (Draine 1978) is ∼120 years (as assumed for β Pic in Dent et al. 2014). It is worth noting that the Kral et al. (2016) model best fits the atomic gas observations for an ISRF 60 times stronger than this, in order to explain the observed high carbon ionisation fraction in the disk. However, a similar ionisation fraction could be attained through effects not taken into account by the model, such as X-ray emission and enhanced UV emis- sion from the central star. These may significantly affect the C i/C ii levels without altering the CO photodissociation timescale, since X-rays do not induce CO photodissociation and stellar UV will mostly affect the CO-poor but C-rich inner disk regions. As such, we here assume a standard level for the UV ISRF. The optically thin dust and the low H2 levels (see next section) are not sufficient to attenuate this UV radiation field. However, in the presence of CO column densities larger than 1013 cm−2, the molecule can self-shield leading to much longer photodestruction timescales (Visser et al. 2009). The vertical column density at the clump location (i.e. brightest spaxel in the PV diagram) can be estimated through Eq. 4, where now the area ∆A of the column in question is the area in the orbital plane (Fig. 5) corresponding to the spaxel size at that PV diagram location (Fig. 4). The CO column density at the clump location through half of the vertical MNRAS 000, 1 -- 21 (2016) ALMA observations of exocometary CO lines in β Pic 15 extent of the disk is ∼ (1.6 − 1.7) × 1015 cm−2, leading to a photodissociation timescale that is ∼2.5 times longer than for the unshielded case (see Table 6 in Visser et al. 2009). In addition to CO self-shielding, we consider UV atten- uation in the vertical direction by the C i ionisation contin- uum, which needs to be taken into account since it acts at the same UV wavelengths as CO photodissociation. We calculate its effect using Eq. 4 in Rollins & Rawlings (2012) and ioni- sation cross sections from van Dishoeck (1988). We estimate the C i vertical column density at the clump radial location from the Kral et al. (2016) thermodynamical model, yielding a value of 2.8×1015 cm−2. This results in a photodissociation rate that is ∼0.96 times the unshielded rate, showing that at 85 AU vertical shielding of CO by the C i ionisation con- tinuum is minor compared to CO self-shielding. Using these shielding factors, we estimate a photodissociation timescale in the clump midplane of ∼300 years, which together with our measured total mass of CO in the disk yields a CO mass loss rate of 0.11 M⊕/Myr (or 6.8 × 1017 kg/year). This is under the simplifying assumption of a photodis- sociation timescale that is constant across the disk. We note that in reality this timescale will decrease down to the un- shielded value as we vertically move away from the midplane. However, most CO mass loss will occur at low heights, where most of the CO mass lies (assuming a vertical Gaussian for the CO number density distribution). This means that in the region where most mass loss occurs, the vertically aver- aged CO mass loss rate is actually well approximated by the midplane value (< 10% decrease). As well as in the vertical direction, the photodissociation timescale will also decrease away from the clump in the azimuthal direction: if we assume that the factor ∼ 5 drop in flux from the peak of the clump to the end of the tail also represents a factor ∼ 5 drop in CO mass and vertical column density, this corresponds to a fac- tor ∼ 1.8 drop in photodissociation timescale. That means that the mass loss rate will decrease by a factor 5/1.8 ∼ 2.8 across the CO tail, meaning that the average rate would be ∼32% lower than that measured at the CO clump. The above considerations only take into account shield- ing of CO along the vertical direction. In reality, at a given disk location, shielding of interstellar UV radiation over the entire 4π solid angle should be considered. This is higher than shielding along the vertical direction alone, as for ex- ample the radial optical depth to UV photodissociating ra- diation is higher than the vertical one, due to a CO column that we can estimate to be ∼ 3 times longer (assuming a Gaussian radial distribution at the clump location), yield- ing a higher column density in the radial direction than in the vertical direction. This effect implies that the average shielding over the entire 4π solid angle is higher, yielding a longer disk-averaged photodissociation timescale, than cal- culated in the vertical direction alone. Overall, we expect the effect of a non spatially uniform photodissociation timescale and of the increased radial optical depth to change our mass loss rate estimate at most by a factor of a few below and above the quoted value, respectively. We can then relate our estimate of the photodestruction timescale to the observed azimuthal extent of the CO tail. In the resonant migration picture (e.g. Fig. 8, bottom), the two clumps on the SW and NE sides of the disk are rotating at the Keplerian angular velocity of the inner planet that is causing the resonance (assumed to be located at 60 AU, MNRAS 000, 1 -- 21 (2016) as in Dent et al. 2014), with the CO lagging because of its slower Keplerian angular velocity at 85 AU. The length of the tails can then be calculated as the azimuthal displace- ment undertaken in one photodissociation timescale at an angular velocity equal to the difference between that of the inner planet at 60 AU and that of gas in Keplerian rotation at 85 AU. Despite the uncertainties in the photodissociation timescale calculation and in the radial location of the planet, we obtain a value of 126◦, which is in good agreement with the deprojection in Fig. 8, bottom. Therefore, a planet lo- cated near the inner edge of the disk can feasibly explain the observed CO azimuthal structure for a disk inclination of ∼86◦ from face-on. 5.2.3 Steady state abundances and the impact of CO2 Unless we are witnessing the system shortly after a stochas- tic event, which is unlikely, we can assume CO gas to be in steady state, with its production rate matching the de- struction rate. Since the production rate of CO is the total mass loss rate from the collisional cascade (derived in Sect. 5.2.1) multiplied by the CO mass fraction in the ice, we can obtain the latter by dividing the destruction rate (derived in Sect. 5.2.2) by the production rate. Given that the produc- tion rate is a lower limit, we obtain an upper limit of 6% on the CO mass abundance in β Pic's exocomets. However, we note that additional CO can be produced through the CO2 molecule, which has a similar ice abundance as CO in Solar System comets (e.g. Mumma & Charnley 2011). Whether CO2 ice can survive as ice in the β Pic disk depends on the temperature of the solid bodies, which in turn depends on their size, composition and radial location. The timescale for a CO2 molecule to sublimate from a comet surface is shorter than an orbital timescale for temperatures >50-55 K (Sand- ford & Allamandola 1990), which matches with the equilib- rium temperature of a blackbody at the radial location of the clump (85 AU). Therefore, we predict that in the inner disk regions all surface CO2 will have disappeared, leaving some of it trapped in other ices (e.g. water ice). This, like CO, may then be released through collisions or H2O pho- todesorption. In the outer regions, on the other hand, CO2 ice can survive on the surface of grains, and may be released again through collisions or photodesorption of CO2 ice itself. Near the radial location of the clump, sublimation of surface CO2 itself may also be contributing to produce CO2 gas. CO is rapidly created via photodissociation of CO2 gas (e.g. van Dishoeck 1988) and/or photodesorption of CO2 ice (e.g. Oberg et al. 2009). The photodissociation timescale of CO2 gas is considerably faster than that of CO in β Pic. In an optically thin medium, using calculated cross sections (Hudson 1971; Lewis & Carver 1983), the interstellar radia- tion field as expressed in Eq. 3 in van Dishoeck (1988), and the stellar radiation field obtain from fit to the SED, we calculate a CO2 photodissociation timescale of 0.35 years at 85 AU, ∼100 times shorter than the CO photodissociation timescale. While we are not aware of whether CO2 can self- shield significantly, we suggest that to first approximation this can be ignored, since too high a column density would have to be produced in a very short timescale for CO2 to be able to shield itself. Additionally, UV photodesorption of CO2 ice can also directly contribute to CO release, since experiments show that 20-50% of CO2 ice will be desorbed 16 L. Matr`a et al. directly as CO ( Oberg et al. 2009). This strengthens our as- sumption that any CO2 gas released is rapidly turned into CO, meaning that the CO gas observed can be interpreted to derive a collective CO+CO2 cometary abundance, which we measure to be at most 6%. 5.2.4 Comparison with other cometary ices In the Solar System, an ice to rock ratio of ∼1 is generally as- sumed, and the (CO+CO2)/H2O abundance ratio observed ranges between 2.4 and 60% (Mumma & Charnley 2011), yielding a CO+CO2 mass abundance in the comets of 2.6 to 27.2%. Therefore, the 6% upper limit mass abundance of CO+CO2 in β Pictoris is consistent with Solar System cometary compositions, suggesting a broadly similar carbon reservoir within the cometary ices of both planetary systems, despite the differences in age and stellar type. On the other hand, compared to the only other extrasolar system where the presence of secondary gas has been confirmed (around the coeval star HD181327, Marino et al. 2016), β Pic's exo- comets have a CO+CO2 content that is at most ∼20 times higher (for the same assumptions in the calculation). While the exact spatial distribution of CO in HD181327 is un- known, the observations suggest that it should be co-located with the dust ring at a radius of ∼86 AU from the star, very similar to the radial location of the clump in the β Pic disk. As the two stars belong to the same moving group, we as- sume that they formed from the same molecular cloud, which may suggest they should retain similar molecular reservoirs (which this remains observationally unconstrained). Differ- ent abundances, however, may arise if different molecule for- mation and/or freeze-out mechanisms have been at play in the original protoplanetary disks, which may in turn relate to the different stellar hosts. However, any such conclusion is highly dependent on the assumptions taken in our cometary abundance calculation, as pointed out before. Though the same assumptions on unknown quantities have been made for both disks, their values may actually differ significantly between the two systems, altering our conclusions on the relative CO+CO2 ice abundance. Nonetheless, it is a valu- able result that the abundances obtained give reasonable estimates compared to the Solar System, and shows that a new era is beginning where we can start setting Solar Sys- tem comets into the wider context of extrasolar cometary bodies. 5.3 An atom-dominated secondary disk Given the short photodissociation timescales, molecules in low gas mass environments such as secondary gas-bearing debris disks are rapidly destroyed to form atoms. For the β Pic disk, a recent model has been put forward by Kral et al. (2016) that is consistent with all far-IR/sub-mm atomic line observations of the system. The model traces the thermody- namical evolution of the atoms released by CO photodissoci- ation, and shows how these rapidly spread viscously from the production point (i.e. the CO clump) to form an axisymmet- ric accretion disk. Because of the ionisation of carbon caused by the strong stellar and interstellar UV field, the disk is re- quired to have a relatively high viscosity α > 0.1, which may be explained by the onset of the magneto-rotational insta- bility (MRI, Kral & Latter 2016). Carbon ionisation is the dominant heating mechanism in the disk, and produces a considerable amount of electrons, which then act as an effi- cient collider species to excite the CO molecule and produce the line ratios observed by ALMA (see Sect. 4.1). Our results in Fig. 11 show that the radial dependence of the electron density predicted by this model is consistent with that derived in this work. While on the other hand the absolute value of the electron density might appear under- predicted by a factor of a few, we caution the reader that our derivation of the electron density from the observations assumes electrons to be the only collisional partners that contribute to the excitation of CO. Other species might con- tribute as well, but the lack of calculated cross sections for collisions between CO and any species that are abundant in the disk (e.g. atomic carbon and oxygen) prevents us from being able to make quantitative predictions (see discussion in Matr`a et al. 2015). In general, the influence of additional collisional partners would lower our measured ne− , meaning that the model is still consistent with our observations. Another test for the model is our estimate of the disk scale height (Sect. 3.3.1). The gas temperature at radii where the CO is detected was predicted by the model to have a value of ∼52 K at 85 AU and decrease as T ∼ R−0.8, which for a vertically isothermal gas dominated by C and O (µ=14) translates into a scale height of 3.6 AU at 85 AU, increasing as H ∼ R1.1. As estimated in Sect. 5.1, CO should rapidly collide and couple with the atomic gas, meaning that we would expect a similar scale height for the CO and atomic disks. However, our results show that CO has a scale height that is twice larger than predicted by the model (7 AU vs 3.6 AU at 85 AU radius) and that it has a shallower radial increase (as ∼ R0.75 vs ∼ R1.1). This difference cannot be explained by the CO disk being decoupled from the atomic disk, since the increase in molecular mass would make the discrepancy even larger. As well as a geometrical effect due to a non-negligible inclination of the disk to the line of sight, which is degenerate with the disk's vertical structure, two physical reasons can be invoked to explain this discrepancy: • The gas is not vertically isothermal. This is to be ex- pected if the gas becomes vertically optically thick to C i- photoionising UV radiation, which is the main heating pro- cess in the disk (Kral et al. 2016). Then, the upper disk lay- ers would be warmer than the disk midplane, which would naturally produce a wider vertical distribution than in the vertically isothermal case. This is also suggested by the ver- tical distribution of Fe i and Ca ii gas, which is in fact even broader than that inferred for CO here (the scale height value derived for Fe i is ∼16 AU at 85 AU radius, for a near exponential vertical profile; Nilsson et al. 2012). • The excitation of CO, and hence the electron density, is not vertically uniform. In our analysis, we assumed the observed vertical flux distribution (after deconvolution by the instrumental resolution) to be a good tracer of the un- derlying CO mass distribution. However, if a vertical gra- dient in electron density is present in the disk, which once again would be expected if C i photoionisation preferentially takes place in the upper layers, the flux vertical distribution observed would appear broader than the true CO mass dis- tribution, skewing our scale height measurement to higher values. Furthermore, if the vertical optical thickness to C i- MNRAS 000, 1 -- 21 (2016) ALMA observations of exocometary CO lines in β Pic 17 photoionising UV radiation decreases as a function of radius, the effects above would be stronger in the disk inner regions compared to the outer regions. This would then naturally explain the shallower radial dependence of the scale height measured here. 5.4 Solving the primordial vs secondary origin dichotomy in gas-bearing debris disks Just like β Pic, an increasing number of young, <100 Myr- old debris disks has recently been detected in molecular and/or atomic gas lines. This phenomenon had only been observed for disks around A stars until recently, when CO in the debris belt around the late-F star HD181327 (mem- ber of the same moving group, and hence coeval with, β Pictoris) was detected by ALMA (Marino et al. 2016). To explain these discoveries, a dichotomy has emerged between the primordial and secondary gas origin scenarios, where the disk is viewed, respectively, as protoplanetary disk where gas dispersal has been delayed (as proposed for HD21997 and HD141569, K´osp´al et al. 2013; Flaherty et al. 2016) or as a debris disk where second-generation gas is released by plan- etesimals through a steady-state collisional cascade and/or a relatively recent stochastic event. In Sect. 4.2 we showed how optically thin CO line ra- tios can be used to probe the density of collisional partners in a gas-bearing debris disk. As discussed in Matr`a et al. (2015), at least three optically thin lines are needed to com- pletely solve the collider density-kinetic temperature degen- eracy; these do not have to be from the same species, as long as the species are located in the same disk regions and therefore are subject to similar collider densities and tem- peratures. From Fig. 10 we can deduce that if the collider density is high enough (towards the right in the plots), LTE applies and the line ratios become dependent solely on the kinetic temperature, meaning we are unable to probe the collider density in the disk. Nonetheless, we can set a lower limit to it, which will correspond to few times the critical density ncrit of the colliders, since by definition LTE applies if ncoll (cid:29) ncrit. If however collider densities are lower, we are in the NLTE regime and we can estimate the collider density directly, either through observations of 3 lines or by making reasonable assumptions on the kinetic temperature, as we did in the case of β Pic. This method is powerful as it can be used to test the presence of otherwise undetectable collider species, such as H2 gas, in the system. Since this species dominates the disk mass in a protoplanetary disk, its presence should be domi- nant in gas-bearing debris disks of primordial origin. H2 is an extremely volatile molecule and unlike heavier species does not freeze-out onto grains (e.g. Sandford et al. 1993, and references therein), meaning that it cannot survive proto- planetary disk dispersal by freezing out on cometary bodies. Therefore, its presence is not expected at all in debris disk gas of secondary origin. However, if we assume a disk to be primordial, H2 will dominate collisions with CO, and in turn the CO line ratio analysis should consistently yield H2 densities typical of protoplanetary disks. In the case of β Pic, under this assumption, we derive H2 densities of order 103 − 105 cm−3. The only direct measurement of the bulk H2 content in a protoplanetary disk comes from the detection of HD in MNRAS 000, 1 -- 21 (2016) the TW Hydrae protoplanetary disk (Bergin et al. 2013). This detection, combined with detailed models of the disk's physical and chemical structure, yielded midplane H2 densi- ties of order 107 − 1010 cm−3. These are considerably higher than the values we derived for β Pic, and crucially are well in the LTE regime of CO excitation. Thus, detection of op- tically thin CO lines that are excited subthermally (i.e. are out of LTE) like in β Pic is already proof that we are look- ing at environments that are H2-depleted by several orders of magnitude. A simple estimate of the vertical column den- sity can then be obtained by assuming a uniform column of length 15 AU (equivalent to about twice the vertical scale height at the clump location in the β Pic disk, Sect. 3.3.1) and our upper limit of 105 cm−3 on the H2 density. This yields a column density of ∼1019 cm−2, which is not suffi- cient to shield the CO from UV photodestruction over the age of the system (Visser et al. 2009). Therefore, even in the presence of low H2 levels, CO cannot have survived since the protoplanetary phase of evolution, and must therefore be replenished through exocometary volatiles. This provides additional and now definitive confirmation of the secondary origin of gas in the β Pictoris disk. Furthermore, we here propose this analysis of multi-transition observations of op- tically thin, subthermally excited CO as a fundamental way to indirectly probe the H2 density in other gas-bearing debris disks, hence solving the primordial versus secondary origin dichotomy. 6 SUMMARY AND CONCLUSIONS In this work, we presented new ALMA Band 6 observations of the CO J=2-1 line in the β Pictoris debris disk, and com- bined them with archival Band 7 observations of the CO J=3-2 line to derive the 3D morphology, excitation condi- tions and total mass of the CO gas. We reach the following conclusions: • We confirm the presence of the CO clump discovered by Dent et al. (2014), peaking at radius of 85 AU and an azimuth of ±32◦ in the orbital plane of the disk (SW on- sky). At the 0.(cid:48)(cid:48)3 resolution of the new dataset, we conclude that the clump is radially extended, spanning ∼100 AU in orbital radius. Since CO should rapidly couple to the atomic gas disk already in place and quickly proceed in Keplerian rotation around the star, it does not have time to spread radially and must be produced at a broad range of radii. This rules out a giant impact between planetary bodies as a possible dynamical scenario to explain the clumpy morphol- ogy, leaving resonant trapping by outward planet migration as the only viable mechanism proposed so far. This scenario is also consistent with the observed azimuthal structure for an assumed disk inclination of ∼86◦, and can be tested by looking for the clump's predicted orbital motion. • The CO disk is vertically tilted, i.e. its PA presents an extra anticlockwise rotation compared to the main disk ob- served in scattered light. The degree of misalignment is in line with that inferred for the warp observed in scattered light imaging of the disk, suggesting a common origin be- tween the CO clump and the scattered light warp. • Under the assumption of a vertically isothermal disk, the scale height measured is a function of orbital radius (in- creasing as H ∼ R0.75 with a value of 7 AU at the clump 18 L. Matr`a et al. radial location). As observed for metallic atoms, this implies temperatures that are higher than predicted by thermody- namical models (Kral et al. 2016). Aside from a geometrical effect due to our edge-on assumption, this discrepancy can be attributed to the fact that the disk is likely not vertically isothermal and/or to a potential higher electron abundance in the disk surface layers, where both effects are expected if the disk is vertically optically thick to C-photoionising UV photons. • The CO J=3-2/J=2-1 line ratio shows a clear gradient with radial velocity in the PV diagrams, which is consis- tent with a power law decrease with orbital radius. Through NLTE modelling, we attribute this to a radial decrease in the electron density in the disk, in line with model pre- dictions. No enhancement at the clump location nor a sig- nificant NE/SW asymmetry are observed, confirming that CO lies in an azimuthally symmetric, atom-dominated sec- ondary disk. • Repeating the NLTE analysis above under the assump- tion of an H2-dominated primordial disk, we derive H2 densi- ties that are many orders of magnitude lower than expected for protoplanetary disks. These imply column densities that are too low to shield CO and allow it to survive over the age of the system, providing further proof that CO needs to be replenished and hence the gas disk must be of secondary origin. We propose such measurement of H2 densities from subthermally excited, optically thin CO lines as a funda- mental way to solve the primordial vs secondary dichotomy in other gas-bearing debris disks. • Using both CO transitions, we refine the total CO mass measurement to 3.4+0.5−0.4×10−5 M⊕. We find that the vertical CO and C i column densities are sufficient to partially shield CO from UV light, prolonging its survival timescale to ∼300 years, which is still much shorter than both the system age (hence requiring replenishment from cometary ices) and the orbital period (hence producing the observed clumpy mor- phology and NE/SW asymmetry). Assuming CO gas to be in steady state, being released from ices through the colli- sional cascade and then rapidly photodestroyed, we estimate the CO+CO2 mass abundance in the comets to be at most 6%. This is still consistent with cometary abundances mea- sured in both our own Solar System and in coeval HD181327 system. eration with the Republic of Chile. The Joint ALMA Obser- vatory is operated by ESO, AUI/NRAO and NAOJ. REFERENCES Ahmic M., Croll B., Artymowicz P., 2009, ApJ, 705, 529 Apai D., Schneider G., Grady C. A., Wyatt M. C., Lagrange A.- M., Kuchner M. J., Stark C. J., Lubow S. H., 2015, ApJ, 800, 136 Aumann H. H., 1985, PASP, 97, 885 Backman D. E., Paresce F., 1993, in Levy E. H., Lunine J. I., eds, Protostars and Planets III. pp 1253 -- 1304 Bergin E. A., et al., 2013, Nature, 493, 644 Beust H., Lagrange-Henri A. M., Vidal-Madjar A., Ferlet R., 1989, A&A, 223, 304 Beust H., Vidal-Madjar A., Ferlet R., Lagrange-Henri A. M., 1990, A&A, 236, 202 Boneberg D. M., Pani´c O., Haworth T. J., Clarke C. J., Min M., 2016, preprint, (arXiv:1605.07938) Brandeker A., Liseau R., Olofsson G., Fridlund M., 2004, A&A, 413, 681 Cataldi G., et al., 2014, A&A, 563, A66 Collings M. P., Dever J. W., Fraser H. J., McCoustra M. R. S., Williams D. A., 2003, ApJ, 583, 1058 Crifo F., Vidal-Madjar A., Lallement R., Ferlet R., Gerbaldi M., 1997, A&A, 320, L29 Deleuil M., et al., 1993, A&A, 267, 187 Dent W. R. F., et al., 2014, Science, 343, 1490 Dickinson A. S., Richards D., 1975, Journal of Physics B Atomic Molecular Physics, 8, 2846 Draine B. T., 1978, ApJS, 36, 595 Ferlet R., Vidal-Madjar A., Hobbs L. M., 1987, A&A, 185, 267 Fern´andez R., Brandeker A., Wu Y., 2006, ApJ, 643, 509 Flaherty K. M., et al., 2016, ApJ, 818, 97 Foreman-Mackey D., Hogg D. W., Lang D., Goodman J., 2013, PASP, 125, 306 Golimowski D. A., et al., 2006, AJ, 131, 3109 Gontcharov G. A., 2006, Astronomy Letters, 32, 759 Goodman J., Weare J., 2010, Commun. Appl. Math. Comput. Sci., 5, 65 Gray R. O., Corbally C. J., Garrison R. F., McFadden M. T., Bubar E. J., McGahee C. E., O'Donoghue A. A., Knox E. R., 2006, AJ, 132, 161 Grigorieva A., Th´ebault P., Artymowicz P., Brandeker A., 2007, A&A, 475, 755 Hobbs L. M., Vidal-Madjar A., Ferlet R., Albert C. E., Gry C., 1985, ApJ, 293, L29 Hogbom J. A., 1974, A&AS, 15, 417 Hudson R. D., 1971, Reviews of Geophysics, 9, 305 Jackson A. P., Wyatt M. C., Bonsor A., Veras D., 2014, MNRAS, ACKNOWLEDGEMENTS 440, 3757 The authors would like to acknowledge D´aniel Apai for pro- viding the disk spine from HST observations, and Grant Kennedy for providing the SED fitting parameters. LM ac- knowledges support by STFC and ESO through graduate studentships and, together with MCW and QK, by the Eu- ropean Union through ERC grant number 279973. Work of OP is funded by the Royal Society Dorothy Hodgkin Fel- lowship, and AMH gratefully acknowledges support from NSF grant AST-1412647. This paper makes use of the fol- lowing ALMA data: ADS/JAO.ALMA#2012.1.00142.S and ADS/JAO.ALMA#2011.0.00087.S. ALMA is a partnership of ESO (representing its member states), NSF (USA) and NINS (Japan), together with NRC (Canada), NSC and ASIAA (Taiwan), and KASI (Republic of Korea), in coop- Kamp I., Bertoldi F., 2000, A&A, 353, 276 Kennedy G. M., Wyatt M. C., 2014, MNRAS, 444, 3164 Kiefer F., Lecavelier des Etangs A., Augereau J.-C., Vidal-Madjar A., Lagrange A.-M., Beust H., 2014, A&A, 561, L10 Kondo Y., Bruhweiler F. C., 1985, ApJ, 291, L1 K´osp´al A., et al., 2013, The Astrophysical Journal, 776, 77 Kral Q., Latter H., 2016, MNRAS, 461, 1614 Kral Q., Th´ebault P., Augereau J.-C., Boccaletti A., Charnoz S., 2015, A&A, 573, A39 Kral Q., Wyatt M., Carswell R. F., Pringle J. E., Matr`a L., Juh´asz A., 2016, MNRAS, 461, 845 Lagrange A. M., Ferlet R., Vidal-Madjar A., 1987, A&A, 173, 289 Lagrange A.-M., et al., 2012, A&A, 542, A40 Lewis B. R., Carver J. H., 1983, J. Quant. Spectrosc. Radiative Transfer, 30, 297 Liseau R., Artymowicz P., 1998, A&A, 334, 935 MNRAS 000, 1 -- 21 (2016) ALMA observations of exocometary CO lines in β Pic 19 Mamajek E. E., Bell C. P. M., 2014, MNRAS, 445, 2169 Marino S., et al., 2016, MNRAS, 460, 2933 Martin-Zaıdi C., et al., 2008, A&A, 484, 225 Matr`a L., Pani´c O., Wyatt M. C., Dent W. R. F., 2015, MNRAS, 447, 3936 Matr`a L., et al., in prep. McDaniel E. W., 1964, Collision phenomena in ionized gases Milli J., et al., 2014, A&A, 566, A91 Milli J., et al., 2015, A&A, 577, A57 Mumma M. J., Charnley S. B., 2011, ARA&A, 49, 471 Nilsson R., Brandeker A., Olofsson G., Fathi K., Th´ebault P., Liseau R., 2012, A&A, 544, A134 Oberg K. I., van Dishoeck E. F., Linnartz H., 2009, A&A, 496, 281 Olofsson G., Liseau R., Brandeker A., 2001, ApJ, 563, L77 Roberge A., Feldman P. D., Lagrange A. M., Vidal-Madjar A., Ferlet R., Jolly A., Lemaire J. L., Rostas F., 2000, ApJ, 538, 904 Roberge A., Feldman P. D., Weinberger A. J., Deleuil M., Bouret J.-C., 2006, Nature, 441, 724 Rollins R. P., Rawlings J. M. C., 2012, MNRAS, 427, 2328 Sandford S. A., Allamandola L. J., 1990, Icarus, 87, 188 Sandford S., Allamandola L., Geballe T., 1993, Science, 262, 400 Slettebak A., 1975, ApJ, 197, 137 Slettebak A., Carpenter K. G., 1983, ApJS, 53, 869 Smith B. A., Terrile R. J., 1984, Science, 226, 1421 Telesco C. M., et al., 2005, Nature, 433, 133 Vidal-Madjar A., et al., 1994, A&A, 290 Visser R., van Dishoeck E. F., Black J. H., 2009, A&A, 503, 323 Walker K. M., Song L., Yang B. H., Groenenboom G. C., van der Avoird A., Balakrishnan N., Forrey R. C., Stancil P. C., 2015, ApJ, 811, 27 White J. A., Boley A. C., Hughes A. M., Flaherty K. M., Ford E., Wilner D., Corder S., Payne M., 2016, preprint, (arXiv:1606.00442) Wyatt M. C., 2003, ApJ, 598, 1321 Wyatt M. C., 2006, ApJ, 639, 1153 Wyatt M. C., 2008, Annual Review of Astronomy and Astro- physics, 46, 339 Wyatt M. C., Pani´c O., Kennedy G. M., Matr`a L., 2015, Ap&SS, 357, 103 Zuckerman B., Song I., 2012, ApJ, 758, 77 van Dishoeck E. F., 1988, in Millar T. J., Williams D. A., eds, Astrophysics and Space Science Library Vol. 146, Rate Coef- ficients in Astrochemistry. pp 49 -- 72 van Leeuwen F., 2007, A&A, 474, 653 APPENDIX A: COMPARING THE ALMA DATASETS In order to allow for direct comparison between the CO J=2- 1 and J=3-2 data cubes, both need to have matching pixel and channel size, as well as spatial and velocity resolution. We therefore re-CLEANed the Band 6 dataset to achieve the coarser pixel size, spatial beam and channel size of the Band 7 dataset. While the CLEAN task within CASA allows us to choose the channel size of the restored data cube, it does so through simple spectral rebinning, and as such does not change the velocity resolution of the dataset (or in other words, its spectral response function). The native spectral response function of ALMA data is a Hanning-smoothed sinc function of width (resolution) equal to twice the native chan- nel width. This function is well approximated by a Gaussian of equal FWHM (ALMA Helpdesk, private communication). MNRAS 000, 1 -- 21 (2016) Therefore, as well as rebinning of channels at the CLEAN- ing stage, convolution by a Gaussian of FWHM equal to the quadratic difference between the target resolution and the native resolution was necessary to allow spectral comparison of the datasets. The stellar position from Hipparcos was corrected for proper motion separately between datasets, and the final images rotated by the position angle of the main scattered light dust disk (29.3◦, Lagrange et al. 2012), aligning the disk midplane to the horizontal on-sky spatial axis xsky (see Fig. 1). The astrometric accuracy of an ALMA dataset is based on the phase tracking accuracy of the array; this de- pends on several factors, including weather, observing band, accuracy in the position of each antenna, distance of the source to the phase calibrator, and frequency of phase cal- ibration measurements. We quantify this as advised in the ALMA Knowledgebase2, and assume a 1σ RMS astrometric accuracy of ∼0.(cid:48)(cid:48)037 for the Band 6 and ∼0.(cid:48)(cid:48)029 for the Band 7 dataset. In order to remove potential systematics in the ALMA phase tracking, we recenter the data cubes using the loca- tion of the star, which is itself detected in both continuum datasets. Since emission at its location includes a contribu- tion from the disk itself, which may bias our procedure, we determined its position through global fitting of disk+star models in each dataset, whose detailed description will ap- pear in forthcoming work (Matr`a et al. prep). In the Band 6 dataset the best-fit stellar offset from the image centre was found to be (−0.(cid:48)(cid:48)0084, 0.(cid:48)(cid:48)034) along the midplane and per- pendicular to it, respectively; this offset was (0.(cid:48)(cid:48)24, 0.(cid:48)(cid:48)12) for the Band 7 dataset, which was likely affected by systematics during the Science Verification Cycle 0 campaign of ALMA. APPENDIX B: FITTING THE DISK VERTICAL STRUCTURE The Gaussian fitting procedure employed to derive the ver- tical structure of the disk takes into account the following uncertainties: pixel, i.e. the RMS of the images in Fig. 1. • The uncertainty ∆SCO on the observed line flux at each • The uncertainty ∆y(cid:48) sky on the the pixel location along the direction which is perpendicular to the disk midplane. Let us define the rotated pixel reference frame to be centred on the stellar position, with x(cid:48) sky along the disk midplane and y(cid:48) sky perpendicular to it, and with both having a positive sign sky towards the SW direction. The position in this (cid:0)x(cid:48) (cid:1) sky, y(cid:48) reference frame was obtained from the original unrotated reference frame (also centred on the star, with xsky positive towards East and ysky positive towards North), through an anticlockwise rotation by an angle of 90◦ − θPA. Thus, y(cid:48) sky can be expressed as a function of xsky, ysky and θPA simply through (cid:48) y sky = sin(90 ◦ − θPA)xsky + cos(90 ◦ − θPA)ysky (B1) 2 https://help.almascience.org/index. php?/Knowledgebase/Article/View/319/0/ what-is-the-astrometric-accuracy-of-alma 20 L. Matr`a et al. This means that uncertainties in the PA (29.◦3+0.2−0.3), xsky and ysky will propagate as an uncertainty in y' through (cid:48)2 sky = sin2(90 ∆y ◦ − θPA)(∆xsky)2 + cos2(90 ◦ − θPA) + yskysin(90 ◦ − θPA)(∆ysky)2+ ◦ − θPA))2(∆θPA)2 +(−xskycos(90 (B2) The uncertainties ∆xsky and ∆ysky are given by the as- trometric accuracy of the instrument (see Appendix A), summed in quadrature with the uncertainty on the star's proper motion (∼0.(cid:48)(cid:48)004). The fit was carried out using the affine invariant Markov- Chain Monte-Carlo (MCMC) Ensemble sampler from Good- man & Weare (2010), implemented through the Python emcee package (Foreman-Mackey et al. 2013). The like- lihood function of the data given a model 1D Gaus- sian was defined through the χ2 distribution, taking into account uncertainties in both flux and vertical position (∆SCO and ∆y(cid:48) sky). The Gaussian vertical centre yobs, FWHM and peak max(SCO,model) were left as free parame- ters, and uniform, uninformative priors set as yobs < 20 AU, 0 < max(SCO,model) < 2.0 × max(SCO,observed) and FWHMbeam <FWHM< 80AU, where FWHMbeam is the resolution of the data projected onto the perpendicular to the disk midplane. For each vertical cut, the posterior prob- ability distribution of the Gaussian parameters, given the data, was sampled using 300 walkers through 500 steps. The resulting best-fit Gaussians are shown in the top panel of Fig. 3 as the black dashed lines. Their best-fit vertical cen- tres yobs and FWHM are plotted as a function of midplane location in the central and bottom panel of Fig. 3. The error bars show the 1σ error obtained from the posterior proba- bility distribution of yobs as sampled through the MCMC. APPENDIX C: FROM PV DIAGRAMS TO ORBITAL PLANE (cid:113) GM∗ (cid:113) GM∗ Assuming gas in a flat circular disk in Keplerian rotation around the star, the magnitude of its orbital velocity will be given by R , where G is the gravitational constant, M∗ is the stellar mass, and R is the radial distance from the star; its direction will be perpendicular to the radius vector. In Cartesian coordinates (centred on the star), the orbital velocity along the y axis can then be expressed as R cos θ, where θ is the angle between the x axis vy = and the radius vector. If we then rotate the disk about the x axis by an angle i corresponding to its inclination to the plane of the sky, its velocity perpendicular to the plane to the sky (i.e. its radial velocity), will be given by R cos θ sin(i). By definition, in the orbital plane (cid:113) GM∗ vrad = (cid:113) GM∗ x = R cos θ, leading to vrad = R3 sin(i)x. Since x is the axis of rotation when transforming from the orbital to the sky plane, the x component of both position and velocity remains unchanged (i.e. x = xsky). This has two important consequences: (i) Since the PV diagram is by definition a map in (xsky, vrad) space, knowing the inclination angle i means that each PV location traces a certain disk radius R. In fact, for a given inclination, each orbital radius defines the slope of a unique straight line in PV space (see Fig. 4 and 9). This can be directly used as a probe of the radial extent of the disk. However, the accuracy in measuring R is limited by the finite spectro-spatial resolution of the instrument; this is particularly important at vrad ∼ 0 (i.e., along the line of sight to the star), where it becomes impossible to disentangle contributions from different orbital radii. sition can immediately be derived through y = ±√ (ii) As R and x in the orbital plane are now known, a y po- R2 − x2. Due to the sign, however, there remains a degeneracy; this can only be solved if the disk is not edge-on, as in that case we have an extra observable, the vertical on-sky loca- tion, that allows to break the degeneracy. As applies for R, derivation of y is severely limited near the line of sight to the star. Therefore, assuming a flat, circular Keplerian disk and an inclination angle, PV diagrams can be used to map disk emission in its orbital plane (see e.g. Fig. 5). APPENDIX D: BREAKING THE DEPROJECTION DEGENERACY USING THE OBSERVED VERTICAL STRUCTURE Combining the methods outlined in Appendix B and C allow us to obtain, for a choice of inclination i and on-sky tilt angle dPA, parameters of the observed best-fit sky-projected vertical Gaussian (centre yobs, standard deviation σobs and peak vertical flux Fpeak) at each (x,±y) location in the disk orbital plane. If the disk is perfectly edge-on (i = 90◦), the degeneracy in the sign of y cannot be broken, as emission coming from in front and behind the sky plane will lie at the same sky-projected vertical location +ysky,orb = −ysky,orb = 0. If the disk is sufficiently inclined to the line of sight (i < 90◦), however, emission coming from behind and in front the sky plane will lie at different vertical locations on the sky plane, on opposite sides of the midplane (i.e. +ysky,orb (cid:54)= −ysky,orb). If (i (cid:28) 90◦), the distance between the two will be large enough for them to be resolved, if both the disk physical vertical thickness and the instrumental resolution allow. If i → 90◦, as is the case for β Pictoris, the two sides will not be vertically resolved. However, if the disk is azimuthally asymmetric, the front and back of the disk will contribute to the observed emission in unequal amounts, causing a vertical shift in the centroid of the observed emission (red solid line in Fig. D1) and consequently also in the centre of the best-fit verti- cal Gaussian yobs (blue dashed line) towards either + or −ysky,orb, whichever contributes most to the emission. We here aim to use this shift to retrieve the fraction of the flux originating at +ysky,orb and -ysky,orb, and hence in front and behind the sky plane at ±y. We note that a degeneracy is still present in that we do not know whether CO above the on-sky midplane is in front or behind the sky plane, and viceversa for CO observed below the midplane. For simplicity, we assume that the two flux contribu- tions we are trying to find take the form of Gaussians with the same standard deviation σobs as that of the best-fit ver- tical Gaussian, but centred at +ysky,orb and -ysky,orb for a given orbit (black dotted lines in Fig. D1). We then extract their relative flux contribution by imposing that the integral MNRAS 000, 1 -- 21 (2016) ALMA observations of exocometary CO lines in β Pic 21 Figure D2. Reduced χ2 map obtained by fitting the double- Gaussian vertical profile model to the CO J=2-1 ALMA data cube (as shown by the green long-dashed and red solid lines in Fig. D1 for a specific sky-projected midplane location and radial velocity). The two free parameters for the orbit are the disk inclination i and on-sky tilt angle dPA. Figure D1. Graphical representation of the procedure employed to recover the flux fraction originating from above (at +ysky,orb) and below the disk on-sky midplane (at −ysky,orb). Here we show an example for an orbit inclined by 89◦ to the line of sight, for a disk that is tilted in the plane of the sky by 3◦, at a location of -19.4 AU along the disk sky-projected midplane, and radial velocity of +0.4 km/s with respect to the star. See main text for details of the procedure. under the sum of the two Gaussians (green area in Fig. D1) from -∞ to yobs is equal to exactly half of the integral be- tween ±∞ under the original best-fit Gaussian (blue dashed curve). This reduces to a simple analytical formula for the fraction f + of flux originating from the +ysky,orb location, (cid:17) , (D1) (cid:16) ysky,orb−yobs erf √ 2σobs f + = erf (cid:16) yobs+ysky,orb (cid:17) (cid:17) (cid:16) yobs+ysky,orb 2σobs √ + erf √ 2σobs where erf is the error function, and of course the fraction originating from the -ysky,orb will be f− = 1 − f +. We note that due to either noise in the ALMA cubes or a choice of orbit (i, dPA) that does not well represent the data, it may happen that at some PV locations (xsky, vrad) yobs does not fall in between ±ysky,orb. In such cases, we assign the entirety of the flux to the nearest of either the +ysky,orb or −ysky,orb locations. By finding the relative contributions of the two Gaus- sians, we are basically creating a double-Gaussian model (green long-dashed line in Fig. D1) that we can then compare to the data (red solid line), for each specific PV location. The free parameters of this model orbit are the disk inclination i and the on-sky tilt angle dPA. To attempt constraining these parameters, we therefore construct this model for a grid of i and dPA and in each case evaluate how well it fits the observed data cube. We do this by simply using the χ2 statistic, showing the resulting reduced χ2 map in Fig. D2. This shows that for any tilt angle dPA our simple model prefers disk inclinations ≥ 86◦, with a formal best-fit at i = 88◦ and dPA = 3◦. The results are discussed in more detail in Sect. 3.3.2. This paper has been typeset from a TEX/LATEX file prepared by the author. MNRAS 000, 1 -- 21 (2016) 3020100102030Vertical distance from sky-projected midplane (AU)0.20.00.20.40.60.81.0Flux (normalised to observed maximum)xsky = -19.4 AUvrad = 0.24 km/si=89◦dPA = 3◦yobs+ysky,orb−ysky,orb
1603.05006
1
1603
2016-03-16T09:45:21
Isotopic ratios of H, C, N, O, and S in comets C/2012 F6 (Lemmon) and C/2014 Q2 (Lovejoy)
[ "astro-ph.EP" ]
The apparition of bright comets C/2012 F6 (Lemmon) and C/2014 Q2 (Lovejoy) in March-April 2013 and January 2015, combined with the improved observational capabilities of submillimeter facilities, offered an opportunity to carry out sensitive compositional and isotopic studies of the volatiles in their coma. We observed comet Lovejoy with the IRAM 30m telescope between 13 and 26 January 2015, and with the Odin submillimeter space observatory on 29 January - 3 February 2015. We detected 22 molecules and several isotopologues. The H$_2^{16}$O and H$_2^{18}$O production rates measured with Odin follow a periodic pattern with a period of 0.94 days and an amplitude of ~25%. The inferred isotope ratios in comet Lovejoy are $^{16}$O/$^{18}$O = 499 $\pm$ 24 and D/H = 1.4 $\pm$ 0.4 $\times 10^{-4}$ in water, $^{32}$S/$^{34}$S = 24.7 $\pm$ 3.5 in CS, all compatible with terrestrial values. The ratio $^{12}$C/$^{13}$C = 109 $\pm$ 14 in HCN is marginally higher than terrestrial and $^{14}$N/$^{15}$N = 145 $\pm$ 12 in HCN is half the Earth ratio. Several upper limits for D/H or 12C/13C in other molecules are reported. From our observation of HDO in comet C/2014 Q2 (Lovejoy), we report the first D/H ratio in an Oort Cloud comet that is not larger than the terrestrial value. On the other hand, the observation of the same HDO line in the other Oort-cloud comet, C/2012 F6 (Lemmon), suggests a D/H value four times higher. Given the previous measurements of D/H in cometary water, this illustrates that a diversity in the D/H ratio and in the chemical composition, is present even within the same dynamical group of comets, suggesting that current dynamical groups contain comets formed at very different places or times in the early solar system.
astro-ph.EP
astro-ph
Astronomy&Astrophysicsmanuscript no. 28041final July 18, 2018 c(cid:13) ESO 2018 Isotopic ratios of H, C, N, O, and S in comets C/2012 F6 (Lemmon) and C/2014 Q2 (Lovejoy) ⋆ ⋆⋆ ⋆⋆⋆ N. Biver1, R. Moreno1, D. Bockel´ee-Morvan1, Aa. Sandqvist2, P. Colom1, J. Crovisier1, D.C. Lis3, J. Boissier4, V. Debout1, G. Paubert5, S. Milam6, A. Hjalmarson7, S. Lundin8, T. Karlsson8, M. Battelino8, U. Frisk9, D. Murtagh10, and the Odin team. 6 1 0 2 r a M 6 1 . ] P E h p - o r t s a [ 1 v 6 0 0 5 0 . 3 0 6 1 : v i X r a 1 LESIA, Observatoire de Paris, PSL Research University, CNRS, Sorbonne Universit´es, UPMC Univ. Paris 06, Univ. Paris Diderot, Sorbonne Paris Cit´e, 5 place Jules Janssen, F-92195 Meudon, France 2 Stockholm Observatory, AlbaNova University Center, SE-106 91 Stockholm, Sweden 3 LERMA, Observatoire de Paris, PSL Research University, CNRS, Sorbonne Universit´es, UPMC Univ. Paris 06, F-75014, Paris, France 4 IRAM, 300, rue de la Piscine, F-38406 Saint Martin d'H`eres, France 5 IRAM, Avd. Divina Pastora, 7, 18012 Granada, Spain 6 NASA Goddard Space Flight Center, Astrochemistry Laboratory, Code 691.0, Greenbelt, MD 20771, USA 7 Onsala Space Observatory, Chalmers University of Technology, SE-439 92 ONSALA, Sweden 8 OHB Sweden, P.O. Box 1269, SE-164 29 Kista, Sweden 9 Omnisys Instruments, August Barks Gata 6B, SE-421 32 Vastra Frolunda, Sweden 10 Dept. of Radio and Space Science, Chalmers Technical University, Gothenburg, Sweden July 18, 2018 ABSTRACT 2 O and H18 The apparition of bright comets C/2012 F6 (Lemmon) and C/2014 Q2 (Lovejoy) in March-April 2013 and January 2015, combined with the improved observational capabilities of submillimeter facilities, offered an opportunity to carry out sensitive compositional and isotopic studies of the volatiles in their coma. We observed comet Lovejoy with the IRAM 30m telescope between 13 and 26 January 2015, and with the Odin submillimeter space observatory on 29 January - 3 February 2015. We detected 22 molecules and several isotopologues. The H16 2 O production rates measured with Odin follow a periodic pattern with a period of 0.94 days and an amplitude of ∼25%. The inferred isotope ratios in comet Lovejoy are 16O/18O=499 ± 24 and D/H =1.4 ± 0.4 × 10−4 in water, 32S/34S = 24.7 ± 3.5 in CS, all compatible with terrestrial values. The ratio 12C/13C = 109 ± 14 in HCN is marginally higher than terrestrial and 14N/15N = 145 ± 12 in HCN is half the Earth ratio. Several upper limits for D/H or 12C/13C in other molecules are reported. From our observation of HDO in comet C/2014 Q2 (Lovejoy), we report the first D/H ratio in an Oort Cloud comet that is not larger than the terrestrial value. On the other hand, the observation of the same HDO line in the other Oort-cloud comet, C/2012 F6 (Lemmon), suggests a D/H value four times higher. Given the previous measurements of D/H in cometary water, this illustrates that a diversity in the D/H ratio and in the chemical composition, is present even within the same dynamical group of comets, suggesting that current dynamical groups contain comets formed at very different places or times in the early solar system. Key words. Comets: general – Comets: individual: C/2012 F6 (Lemmon), C/2014 Q2 (Lovejoy) – Radio lines: solar system – Submillimeter 1. Introduction Comets are the most pristine remnants of the formation of the solar system 4.6 billion years ago. Investigating the composition of cometary ices provides clues to the physical conditions and chemical processes at play in the primitive solar nebula. Comets may also have played a role in the delivery of water and organic ⋆ Based on observations carried out with the IRAM 30m telescope. IRAM is supported by INSU/CNRS (France), MPG (Germany) and IGN (Spain). ⋆⋆ Odin is a Swedish-led satellite project funded jointly by the Swedish National Space Board (SNSB), the Canadian Space Agency (CSA), the National Technology Agency of Finland (Tekes) and the Centre National d' ´Etudes Spatiales (CNES, France). The Swedish Space Corporation is the prime contractor, also responsible for Odin operations. ⋆⋆⋆ The spectra datset mous http://cdsweb.u-strasbg.fr/cgi-bin/qcat?J/A+A/ the CDS via anony- via is available at (130.79.128.5) ftp to cdsarc.u-strasbg.fr or material to the early Earth (see Hartogh et al., 2011, and refer- ences therein). The latest simulations of the early solar system evolution (Brasser & Morbidelli , 2013; O'Brien et al., 2014), suggest a more complex scenario. On the one hand, ice-rich bod- ies formed beyond Jupiter may have been implanted in the outer asteroid belt and participated in the supply of water to the Earth or, on the other hand, current comets coming from either the Oort Cloud or the scattered disk of the Kuiper belt may have formed in the same trans-Neptunian region sampling the same diversity of formation conditions. Understanding the diversity in compo- sition and isotopic ratios of the comet material is thus essential in order to assess such scenarios (Altwegg & Bockel´ee-Morvan, 2003; Bockel´ee-Morvan et al., 2015). The recent years have seen significant improvement in the sensitivity and spectral coverage of millimeter receivers. The EMIR receivers (Carter et al., 2012) at the Institut de radioas- tronomie millim´etrique (IRAM) are equipped with a fast Fourier transform spectrometer that offers a wide frequency coverage at 1 Biver et al.: Isotopic ratios of H, C, N, O, S in comets Lemmon and Lovejoy a high spectral resolution (0.2 MHz). The combination enables sensitive spectral surveys of cometary atmospheres and simul- taneous observations of several molecules, including isotopo- logues in brighter comets. We report here observations of iso- topic ratios with the IRAM 30m radio telescope in two very ac- tive (maximum water outgassing rate close to 1030 molec. s−1) Oort cloud comets, C/2012 F6 (Lemmon) and C/2014 Q2 (Lovejoy), carried out in 2013 and 2015. They are both dynam- ically old Oort cloud comets: an original orbital period of 9 800 for comet Lemmon, and years and an orbit inclination of 83◦ a period of 11 000 years and an inclination of 80◦ for comet Lovejoy. The analysis of the observation in terms of molecular abun- dances and detection of rare and new molecular species has been reported (Biver et al. , 2014, 2015). In the present article we con- centrate on the measurement of several isotopic ratios. We report the detection of HDO in both comets. The compounds H16 2 O and H18 2 O were detected in comet Lovejoy with the Odin submillime- ter space observatory, which helped to accurately determine the D/H and 16O/18O ratio in water in this comet. Comet C/2012 F6 (Lemmon) was observed in March and April 2013 (as already presented in Biver et al. , 2014). Perihelion was on 24.51 March 2013 UT at 0.731 AU from the Sun. We mostly used the same tunings as for comet Lovejoy with the EMIR 1mm receiver combined with the FTS and VESPA spectrometers. The best weather conditions were found on 14.5 and 18.5 March, but with a comet at low elevation (15-20◦), and on 6.5 April 2013. We covered the 240–272 GHz frequency range and obtained noisier data on the 85–93, 166–170, and 210– 242 GHz bands with the EMIR 3mm, 2mm, and 1mm receivers, respectively. A summary of the observational setups is given in Table 1. In this paper we focus only on the measurement of isotopic ratios, and spectra of isotopologues are presented in Figs. 3– 8. The chemical inventory of these comets based on these observa- tions with the IRAM telescope has been presented in Biver et al. (2014, 2015) and will be further detailed in a future paper. 2.2. OdinobservationsofcometC/2014Q2(Lovejoy) 2. Observations 2.1. IRAMObservations Comet C/2014 Q2 (Lovejoy) was observed with the IRAM 30m radio telescope during two periods: on 13.8, 15.8, and 16.8 January 2015 (geocentric distance ∆ = 0.496 − 0.528 AU) and on 23.7, 24.7, 25.7, and 26.7 January 2015 (∆ = 0.624 − 0.675 AU) under very good weather conditions. The heliocentric dis- tances were 1.31 and 1.29 AU, respectively. Perihelion was on 30.07 January UT, at 1.290 AU from the Sun. The 13–25 and 26 January observations were conducted with the EMIR 1mm and 3mm receivers, respectively. The main backend used is a Fourier transform spectrometer, which covers a frequency range of 2×8 GHz (two sidebands separated by 8 GHz, each in two linear polarizations) in a single setup, with a high spectral reso- lution of 200 kHz. The spectral resolution (0.3–0.2 km s−1when converted into Doppler velocity) allows the velocity profiles of the narrow cometary lines (∼ 2 km s−1) to be resolved. We also used the high resolution VESPA autocorrelator set to 40 kHz sampling on dedicated lines of interest (e.g., CH3OH (5+2 − 4+1E) and H15CN (3-2)), but VESPA units can only cover lines within a 0.5 GHz window centered at 6.25 GHz in the IF band (upper and lower side band). Using four different tunings we covered the 210–218, 225–233, and 240–272 GHz frequency ranges and 85–93 and 101–109 GHz. The half power beam width of the IRAM 30 m telescope in the 1 mm band ranges from 9.1′′ to 11.6′′, which corre- sponds to 3300 to 5400 km at the comet distance (22–28′′ at 3 mm corresponding to 11000–14000 km on 26.7 Jan.). Given the expansion velocity of the coma (∼ 0.8 km s−1) and helio- centric distance, this means that we are not very sensitive to the photo-dissociation rate of the molecules, provided that the molecule lifetime at 1 AU is larger than 4000 s. The pointing accuracy (1′′-2′′) was regularly checked on reference pointing sources and also on the comet with coarse maps. The time vari- ation of the activity of the comet could be monitored with the multiple strong CH3OH lines present in all setups. We did not see any variations significantly larger than ±20% during the ob- servations (QCH3OH = 1.2±0.2×1028 molec. s−1, 13–16 January, and 1.4 ± 0.2 × 1028 molec. s−1, 23–25 January), which justified the averaging of several days of observations. 2 Because comet C/2014 Q2 (Lovejoy) clearly became a very active comet, observations with the Odin 1.1m submillimeter satellite (Frisk et al. , 2003) were triggered on a late notice. Observations took place between 30.5 January and 3.4 February 2015. These observations were scheduled to map the emission of H16 2 O at 556.9 GHz, follow its temporal evolution, and detect the H18 2 O line at 547.7 GHz. The half power beamwidth is close to 2.1′ at these frequencies and the main beam efficiency is 0.89. Odin uses three spectrometers: a 1 GHz bandwidth acousto- optical spectrometer (AOS with 0.6 MHz sampling) and two autocorrelators (AC1 and AC2) used in their highest resolution mode (∼0.15 MHz with 100 MHz bandwidth). For the first setup, two receivers tuned to the H16 2 O line were used in parallel, cou- pled to the three spectrometers: AOS and AC2 on receiver 555- B2 and AC1 on 549-A1. For the second setup, the AOS was connected to the 572-B1 receiver tuned near the ammonia line at 572.5 GHz, but not frequency locked, while AC2 was connected to the receiver 555-B2 and AC1 to 549-A1, both tuned to the H18 2 O line. As Odin is in a helio-synchronous polar Earth orbit at 550 km altitude, the comet was only observable during ∼55 min of each 96 min orbit. The Earth was in the field of view dur- ing the remaining time, and the atmospheric lines were used to calibrate the frequency scale. Some of the observations failed to track the comet correctly, but half were successful and pro- vided good results. Spectra of the nucleus-centered H16 2 O and 2 O lines are shown in Fig. 3 and a map of the H16 H18 2 O line is shown in Fig. 1. 3. Modeling of the lines and model parameters 3.1. Velocityandoutgassingpattern The mean gas expansion velocity was derived from the shape of the lines with the highest signal-to-noise ratio. We deter- mined a value of 1.1 to 1.0 km s−1 for comet Lemmon and 0.8 km s−1for comet Lovejoy (Table 1). As discussed hereafter, comet Lovejoy was observed with the larger beam of Odin; a value of 0.85 km s−1 provided a better fit to the width of the wa- ter lines. This is expected, as acceleration of the gas expansion velocity in the coma is predicted and observed in the outer part of the comae of active comets. Biver et al.: Isotopic ratios of H, C, N, O, S in comets Lemmon and Lovejoy Table 1. Circumstances of IRAM observations and reference parameters. UT date < rh > (yyyy/mm/dd.d–dd.d) (AU) Comet C/2012 F6 (Lemmon): 0.758 2013/03/14.52–14.55 2013/03/14.57–14.59 0.758 0.753 2013/03/15.52–15.57 0.753 2013/03/15.58–15.59 0.741 2013/03/18.53–18.55 0.741 2013/03/18.55–18.59 2013/04/06.40–06.51 0.776 0.776 2013/04/06.53–06.60 0.783 2013/04/07.49–07.52 0.783 2013/04/07.53–07.59 0.790 2013/04/08.43–08.50 2013/04/08.51–08.54 0.790 2013/04/08.55–08.59 0.791 Comet C/2014 Q2 (Lovejoy): 2015/01/13.75–13.77 2015/01/13.79–13.84 2015/01/13.85–13.88 2015/01/15.82–15.83 2015/01/15.85–15.96 2015/01/16.76–16.78 2015/01/16.81–16.95 2015/01/23.72–23.75 2015/01/24.73–24.75 2015/01/25.71–25.75 2015/01/26.74–26.75 1.314 1.314 1.313 1.308 1.308 1.306 1.306 1.294 1.293 1.292 1.291 < ∆ > (AU) Integ. time (min)a Freq. range (GHz) vexp (km s−1) 1.348 1.349 1.362 1.362 1.400 1.400 1.592 1.592 1.600 1.600 1.606 1.606 1.607 0.496 0.497 0.497 0.516 0.516 0.526 0.527 0.624 0.641 0.658 0.675 26 20 21 14 17 20 69 54 20 58 62 28 36 14 29 29 13 105 14 120 36 14 42 6 249-256, 264-272 240-248, 256-264 249-256, 264-272 210-218, 226-234 249-256, 264-272 240-248, 256-264 249-256, 264-272 240-248, 256-264 249-256, 264-272 210-218, 226-234 85-93, 166-170 249-256, 264-272 218-226, 234-242 249-256, 264-272 210-218, 225-233 240-248, 256-264 249-256, 264-272 240-248, 256-264 249-256, 264-272 240-248, 256-264 240-248, 256-264 226-234, 241-249 210-218, 225-233 85-93, 101-109 1.10 1.10 1.10 1.10 1.10 1.10 1.00 1.00 1.00 1.00 1.00 1.00 1.00 0.80 0.80 0.80 0.80 0.80 0.80 0.80 0.80 0.80 0.80 0.80 Tkin (K) 110 110 110 110 110 110 100 100 100 100 100 100 100 73 73 73 73 73 73 73 73 73 73 73 Qb,c H2O (molec. s−1) QCH3OH (molec. s−1) ∼ 10 × 1029 ∼ 10 × 1029 ∼ 10 × 1029 ∼ 10 × 1029 ∼ 10 × 1029 ∼ 10 × 1029 ∼ 9 × 1029 ∼ 9 × 1029 ∼ 9 × 1029 ∼ 9 × 1029 ∼ 9 × 1029 ∼ 9 × 1029 ∼ 9 × 1029 5 × 1029 5 × 1029 5 × 1029 5 × 1029 5 × 1029 5 × 1029 5 × 1029 6 × 1029 6 × 1029 6 × 1029 7 × 1029 2.1 ± 0.1 × 1028 2.3 ± 0.1 × 1028 1.7 ± 0.2 × 1028 1.6 ± 0.2 × 1028 2.1 ± 0.1 × 1028 1.5 ± 0.1 × 1028 1.4 ± 0.1 × 1028 1.6 ± 0.2 × 1028 1.2 ± 0.4 × 1028 1.5 ± 0.1 × 1028 1.5 ± 0.1 × 1028 1.7 ± 0.2 × 1028 1.17 ± 0.03 × 1028 1.25 ± 0.07 × 1028 1.25 ± 0.02 × 1028 1.18 ± 0.04 × 1028 1.42 ± 0.02 × 1028 1.14 ± 0.01 × 1028 1.17 ± 0.02 × 1028 1.48 ± 0.03 × 1028 1.27 ± 0.06 × 1028 1.00 ± 0.04 × 1028 1.99 ± 0.20 × 1028 Notes. a On nucleus pointings only. b For comet Lemmon, Combi et al. (2014) provide QH2O values in the range 7–12×1029molec. s−1during this period. Contemporaneous SWAN measurements are slightly lower than quoted here, but they were not time-deconvolved and the average values measured by SWAN a few days later are within 10% of those given here. Given our measured QCH3OH (Col. 9), QH2O are also in agreement with the measurement of Paganini et al. (2014) as they yield the same QCH3OH/QH2O ratio. Nanc¸ay observations are also compatible with these QH2O although the quenching of the OH maser is poorly constrained and the uncertainty on QOH is very large. c For comet Lovejoy, QH2O are based on Odin and Nanc¸ay data (see text and Biver et al. (2015)). 3.2. Temperatureandcollisionalexcitation Except for H16 line intensities are proportional to the production rates. 2 O and HCN, lines are optically thin, meaning that For all molecules we used the latest spectroscopic data avail- able in the JPL (Pickett et al., 1998) and CDMS (Muller et al., 2005) databases, both for line identification and computa- tion of line strengths. We employed the code previously used for other comets (Biver et al., 2007; Hartogh et al., 2011; Bockel´ee-Morvan et al., 2012) to compute the excitation of the rotational levels of the molecules and the line intensities. We take into account the pumping of the rotational levels by the fluores- cence of the infrared vibrational bands, which is not expected to play a major role at the cometocentric distances ≤5000 km sampled in the case of comet Lovejoy. In our model molecules slowly evolve from the local thermal equilibrium (T = 100 − 110 and 73 K, for comets Lemmon and Lovejoy, respectively; Table 1) maintained by collisions with neutrals and electrons close to the nucleus to spontaneous decay of the rotational pop- ulation in the less dense parts of the coma (Biver et al., 2000, 2006). Within the accuracy of the detections, line intensity ratios (or rotational temperatures) for methanol, and also for most other species, are in agreement with the model (e.g., Trot(CH3CHO, Lovejoy)= 67 ± 15 K for 68 K predicted). The density is de- scribed with the Haser model, using lifetimes from Crovisier (1994); Bockel´ee-Morvan et al. (2000); Crovisier et al. (2004a); and Crovisier et al. (2004b). Line intensities are computed with a radiative transfer code that takes opacity effects into account. 4. Water production rate and time evolution 4.1. Odinmapsandexcitationofwaterinthecomaofcomet Lovejoy Several maps (e.g., Fig. 1) of the water emission were performed for comet Lovejoy with Odin in order to precisely derive the water production rate. The extension of the emission of opti- cally thick submillimeter water lines in cometary comae is in large part governed by the excitation mechanism of the wa- ter rotational levels. It depends both on the temperature pro- file and collision rates with electrons and neutrals in the in- ner part of the coma (Biver et al., 2007; Hartogh et al., 2011; Bockel´ee-Morvan et al., 2012). However, in the outer part (be- yond 150′′) the line intensity mostly depends on the water pro- duction rate (and assumed photo-dissociation lifetime), and very little on the assumed temperature profile. Dividing the gas tem- perature by a factor of 2 only increases the derived production rate by 10%. We used the same excitation parameters for the H16 2 O observations with Odin, as those derived for the molec- ular lines observed with the IRAM 30m instrument: a constant temperature of 73 K throughout the coma and an electron den- sity scaling factor xne = 0.5 (Zakharov et al., 2007; Biver et al., 2007). The expansion velocity inferred from the H16 2 O and 3 Biver et al.: Isotopic ratios of H, C, N, O, S in comets Lemmon and Lovejoy Fig. 1. Map of the H16 2 O (110 − 101) line obtained with Odin and its acousto-optical spectrometer in comet Lovejoy. The arrow provides the projected direction of the Sun (phase angle was 50◦), but the water emission does not exhibit any significant asymmetry. 2 O H18 line widths is vexp = 0.85 km s−1 , slightly larger than for the lines observed with the smaller beam of the IRAM 30m telescope. The fit to the radial profile of the line inten- sity from the map (Fig. 1) is satisfactory and corresponds to QH2O = 8.7 ± 0.7 × 1029 molec. s−1and the reduced χ2 ν = 2.4 when we include the 5% uncertainty due to the 10′′ (Frisk et al. , 2003) pointing uncertainty. Taking into account an increase in the velocity with distance from the nucleus does not significantly change the goodness of the fit and the derived production rate. 4.2. PeriodicvariationinH16 2 OandH18 2 Oproductionrates 2 O and H18 Since H16 2 O measurements with Odin were not si- multaneous, but interlaced, in order to derive an accurate H16 2 O /H18 2 O production rate ratio it is necessary to determine if significant short-time variation of the activity is present. Inspection of production rates does suggest that time variation is present in the data. Since the observations of H18 2 O constituted a series of ten successive Odin orbits, we did a time averaging of ≈ 2.5 h, corresponding to two consecutive orbits of observations, yielding a sufficient signal-to-noise ratio (5–10) to look for time variations. We fitted a sinusoidal time variation to the individual measurements of the production rates of H16 2 O , inde- pendently. We only selected the H16 2 O measurements with point- ing offsets less than ∼100′′. The results are provided in Table 2. For both molecules, the sinusoidal fit is better than assuming no 2 O and H18 4 2 O and H18 2 O than H18 time variation (χ2 ν reduced by a factor ∼2.5), and the inferred pe- riod is similar (T p=0.922-0.940 day). A simulation of periodic variations of the production rates anticipated a longer delay be- tween peak outgassing and peak signal for H16 2 O (0.25 versus 0.14 day), which we took into account. We found differ- ent reference times T0 for H16 2 O , but these are strongly correlated to the T p value. Formally, the T p-independent 1σ un- certainty δT0 is on the order of 0.4 days. We also tried to fit the data with the same time dependence for H16 2 O , us- ing a fixed mean period of T p = 0.935 days (weighted average) and a corresponding reference time 31.742 Jan. 2015 UT. The goodness of the fits is similar (Table 2). A larger damping of the apparent production rate amplitude was expected for H16 2 O due to optical depth effects. From the observations the effect is not so obvious: ∆ Q/Q0 is only slightly smaller for H16 2 O than for H18 2 O (0.19 versus 0.24). We did not investigate the possible variation of the gas temperature with the production rate, but this could explain a variation of the line intensity with time that was larger than anticipated. 2 O and H18 We then combined both QH16 2 O to derive our best estimate of the water production rate as a function of time during the period 30 January to 3 February 2015 (last line of Table 2). Figure 2 shows the time evolution of the production rates mea- sured for the two water isotopologues and best fit. 2 O and QH18 Biver et al.: Isotopic ratios of H, C, N, O, S in comets Lemmon and Lovejoy 2 O (left vertical scale) and H18 Fig. 2. H16 2 O (pink, right vertical scale) production rates in comet Lovejoy from Odin observa- tions. Horizontal scale is the time relative to perihelion (30.069 January 2015 UT) in days, minus the expected delay between peak outgassing and peak intensity due to beam dilution and opacity effect for a periodic outgassing with similar periodicity. Production rates are apparent production rates computed in the stationary regime assumption. For H16 2 O the larger error bar in- cludes a 5% uncertainty due to the 10′′ pointing uncertainty. H18 2 O uncertainty is dominated by statistical noise. The least square fit to the combined production rate (last line of Table 2) is overplotted. 4.3. Referencewaterproductionrates The reference water production rates were primarily derived from contemporaneous observations of the OH radical maser lines at 18cm with the Nanc¸ay radio telescope. During the 12.8– 16.8 January 2015 observing period of comet Lovejoy, both ob- servations centered on the comet and at 3.0′ offset position were used to better constrain the quenching of the maser inversion (Gerard et al., 1998) yielding an average water production rate QH2O = 5.0 ± 0.2 × 1029molec.s−1 (Biver et al. , 2015). For other periods and for comet Lemmon, the observations of the OH rad- ical could not be used to derive an accurate water production rate owing to significant uncertainties in the quenching of the maser inversion. We either used the water production rates reported elsewhere (Combi et al., 2014; Paganini et al., 2014) or made an interpolation from the Odin observations of comet Lovejoy at perihelion. Based on Nanc¸ay and Odin observations, we adopt a mean QH2O = 5, 5.5, 6, and 7 × 1029 molec. s−1 for the 13– 16 January, 13–25 January, 23–25 January, and 26 January–3 February 2015 time intervals, respectively, with a ±20% time variation possible within each time interval. Observations done on a short time interval might be affected by a periodic time vari- ation, but this was tracked from the monitoring of the methanol production lines, which are present in all IRAM observations settings (Table 1). 5. Isotopic ratios Table 3 provides the list of molecular lines (main species and iso- topologues) observed at IRAM in comets C/2012 F6 (Lemmon) in 2013 and in C/2014 Q2 (Lovejoy) in 2015 and considered in this study. Line intensities and production rates (or upper limits) derived using the models presented in Sect. 3 are given. The cor- responding isotopic ratios or 3 σ limits are presented in Table 4. 2 O (110−101) line at 556.936 GHz, and H18 Fig. 3. Comet C/2014 Q2 (Lovejoy): Average on-nucleus spectra of the H16 2 O (110−101) line at 547.676 GHz obtained with Odin between 30 January and 3 February 2015. The spectrum of the HDO (211 − 212) line at 241.561 GHz observed with the IRAM 30m telescope on 13- 24 January 2015 is shown below. While the H18 2 O and HDO lines are marginally blue-shifted (∆ v = −0.07km s−1 in both cases) the H16 2 O (110 − 101) line is asymmetric and red-shifted (∆ v = +0.33km s−1 ) due to self absorption by the cool gas in the foreground of this optically thick line. Vertical scale is main beam brightness temperature and horizontal scale Doppler ve- locity in the comet rest frame. 5.1. 16O/18OratioincometLovejoy The analysis of Odin observations in Sect. 4.2 (Table 2) enabled us to derive an accurate H16 2 O production rate ratio. From the sinusoidal fits and derived mean Q0 production rates, we ob- tain: 2 O /H18 – Q0(H16 – Q0(H16 2 O )/Q0(H18 2 O )/Q0(H18 2 O ) = 493 ± 29 for the independent fits; 2 O ) = 499 ± 24 for the phased fits. The second approach, which corresponds to a H16 2 O ratio that does not vary with time, yields 16O/18O = 499 ± 24, which is exactly the terrestrial value (498.7). 2 O /H18 5 Biver et al.: Isotopic ratios of H, C, N, O, S in comets Lemmon and Lovejoy Table 2. Periodic variation of water production in comet Lovejoy Odin measurements 29 January – 3 February 2015, rh = 1.29 AU Data points Molecule Production rate variation fitteda Q0 ∆ Q [1027molec. s−1] T p [days] T0 [UT day] 23b 23b 23b 15 15b 15 38b H16 H16 H16 H18 H18 H18 2 O 2 O 2 O 2 O 2 O 2 O 2 O , H18 H16 2 O e 751.6 ± 29.5 724.4 ± 14.6 774.0 ± 12.4 1.524 ± 0.065 1.453 ± 0.062 1.421 ± 0.055 724 ± 17 140 ± 21 141 ± 22 0.0c 0.374 ± 0.085 0.314 ± 0.083 0.0c 0.938 ± 0.014 0.935d 31.797 ± 0.038-Jan. 31.742-Jan.d – – 0.922 ± 0.031 0.935d 31.400 ± 0.035-Jan. 31.742-Jan.d – – 147 ± 19 0.939 ± 0.013 31.912 ± 0.022-Jan. χ2 ν 1.03 1.09 2.78 1.21 1.23 2.44 1.00 a: Q = Q0 + ∆ Q × sin(2π(t − T0)/T p), where T p is the period in days of the fitted sine variation and T0 the reference time for the phase. b: including 5% uncertainty due to a 10′′ pointing uncertainty. c: fit of constant production. d: fixed value (see text). 2 O /H18 e: Assuming H16 2 O = 500. = 6.5 ± 1.6 × 10−4. This is four times the Earth value (VSMOW D/H=1.56 × 10−4), the highest value found in a comet, but com- patible with the Earth value at the 3 σ level. 5.2.2. D/H ratio in comet C/2014 Q2 (Lovejoy) 2 O and H18 The tracking of the comet activity via monitoring of the CH3OH production rate and the precise determination of the water out- gassing rate using OH, H16 2 O make us confident that we have a good H2O production rate reference in order to de- rive the D/H ratio. We report here the detection of HDO with the IRAM 30m telescope in this comet. The HDO (211 − 212) line was observed on 13, 15, 16, 23, and 24 January 2015. The average yields a 4 σ detection (Fig. 3) corresponding to QHDO = 15.3 ± 4.1 × 1025molec. s−1, hence a D/H ratio of 1.4 ±0.4 ×10−4. Although the detection is marginal, in the worse case taking a 5 σ upper limit would yield D/H < 1.86 × 10−4, suggesting that comet Lovejoy has a D/H ratio for water simi- lar to that of the Earth (VSMOW D/H=1.56 × 10−4), or possi- bly even lower. Previous measurements of the D/H ratio in Oort cloud comets yielded values larger than VSMOW (see Sect. 6.1). 5.3. 14N/15NratioinHCN One of the primary goals of the observing program of comet Lovejoy was to measure the 14N/15N ratio in HCN. On January 13, 15, and 16 we first observed the main isotopologue transition HCN (3-2) at 265.886 GHz, then carried out longer integrations using the setup covering the H13CN (3-2) line at 259.012 GHz and the HC15N (3-2) line at 258.157 GHz (Fig. 5). The HCN reference production rate was computed taking opacity effects into account. The average value for the 13–16 January period is QHCN = 4.87 ± 0.02 × 10 × 1026 molec. s−1, using pointing at the nucleus and at 5–8′′offsets – which both yield exactly the same value. The HCN production rate based on the hyperfine component HCN (J = 3 − 2, F = 2 − 2) with a total statistical weight of 3.7% is marginally higher (+14%, 2 σ), pos- sibly because of contamination by the wings of the main HCN hyperfine components. In order to minimize biases due to time variation of the out- gassing and other observing uncertainties (calibration, pointing uncertainty, etc.), we rescaled the HCN production rate to the Fig. 4. Comet C/2012 F6 (Lemmon): average on-nucleus spec- trum of the HDO (211 − 212) line at 241.561 GHz observed with the IRAM 30m telescope between 14 March and 8 April 2013. Vertical scale is main beam brightness temperature and horizon- tal scale Doppler velocity in the comet rest frame. 5.2. D/Hratio Thanks to the wide frequency coverage of IRAM observations, lines of deuterated species of known cometary molecules were observed. None were detected except the HDO (211 − 212) line at 241.562 GHz (Fig. 3). This is not surprising as water is over a hundred times more abundant than HCN, H2S, or H2CO. Nevertheless, the upper limits on the D/H ratio in H2S or H2CO in comet Lovejoy are the lowest obtained so far in a comet. 5.2.1. D/H ratio in water in comet C/2012 F6 (Lemmon) We observed the HDO (211 − 212) line at 241.561 GHz in comet C/2012 F6 (Lemmon) in March-April 2013. The detection of HDO (Fig. 4) is also marginal (4 σ) and the retrieved intensity was possibly affected by poorer baselines on 8 April when we could not use the telescope wobbler to cancel sky emission. The inferred average HDO production rate for the period 14 March to 8 April 2013 is QHDO = 13 ± 3 × 1026molec. s−1, yielding D/H 6 Biver et al.: Isotopic ratios of H, C, N, O, S in comets Lemmon and Lovejoy Fig. 5. Comet C/2014 Q2 (Lovejoy): Average on-nucleus spec- tra of the HCN (3-2) line at 265.886 GHz, H13CN (3-2) line at 259.012 GHz, and HC15N (3-2) line at 258.157 GHz ob- tained with the IRAM 30m telescope on 13–24 January 2015. For HC15N (3-2) the spectrum obtained with the VESPA au- tocorrelator (thick line) is superimposed on the FTS spectrum. Vertical scale is main beam brightness temperature and horizon- tal scale Doppler velocity in the comet rest frame. Fig. 6. Comet C/2012 F6 (Lemmon): average on-nucleus spec- trum of the HCN (3-2) line at 265.886 GHz obtained with the IRAM 30m telescope between 14 March and 8 April 2013. Spectra covering the same velocity range obtained at simi- lar times for H13CN (3-2) (259.012 GHz) and HC15N (3-2) (258.157 GHz) are also shown for comparison, but do not show a clear detection. Vertical scale is main beam brightness tem- perature and horizontal scale Doppler velocity in the comet rest frame. time when H13CN (3-2) and HC15N (3-2) were observed, using the CH3OH production rate as a reference since several methanol lines were observed in each setup. The correction applied to QHCN is +10% for the 13–16 January observations and +27% for 23–25 January. The compound HC15N (3-2) is clearly de- tected during both periods (signal-to-noise ratios of 9 and 6). The derived 14N/15N ratios are 145 ± 16 and 144 ± 23, for 13–16 January and 23–25 January, respectively. The compounds HC15N (3-2) and H13CN (3-2) were not clearly detected (Fig. 6) in comet C/2012 F6 (Lemmon), but the 2 σ value or lower limit we obtained for the 14N/15N ratio in HCN (≥ 106 or 152 ± 72) is fully compatible with values found in other comets. 5.4. 12C/13Cratio The H13CN (3-2) line at 259.012 GHz is in the same spectral band as HC15N (3-2). The production rates (assuming the same photo-dissociation lifetime as for HCN) and relative abundances are derived in the same way for both molecules. Owing to its hyperfine structure similar to that of H12CN (3-2) (Fig 5), the H13CN (3-2) line only contains 93% of the flux in the (-1.2 – 1.2 km s−1) window and this was taken into account. The derived 12C/13C ratios in comet Lovejoy and Lemmon are 109 ± 14 and 124 ± 64, respectively. Within the uncer- tainties, they are compatible with the terrestrial value (89.7) although slightly higher (1.4 σ for Lovejoy). This trend was also observed for comet 17P/Holmes (Bockel´ee-Morvan et al., 2008) and Hale-Bopp (Jewitt et al., 1997). Combining the three 12C/13C measurements yields a value 2.5 σ larger than the Earth value. This interesting result needs to be further investigated in other comets. The sensitivity was not sufficient to detect other 13C iso- topologues. The most sensitive limit was reached for the 13CH3OH (5+2 − 4+1E) and 12CH3OH (5+2 − 4+1E) lines at 261.113 and 266.838 GHz, respectively. The derived lower limit is 12C/13C > 61 in comet Lovejoy, compatible with the Earth value. 7 Biver et al.: Isotopic ratios of H, C, N, O, S in comets Lemmon and Lovejoy Fig. 7. Comet C/2014 Q2 (Lovejoy): Average on-nucleus spec- tra of the CS (5-4) line at 244.936 GHz and C34S (5-4) line at 241.016 GHz obtained with the IRAM 30m telescope on 13- 24 January 2015. Vertical scale is main beam brightness tem- perature and horizontal scale Doppler velocity in the comet rest frame. Fig. 8. Comet C/2012 F6 (Lemmon): average on-nucleus spec- tra of the CS (5-4) line at 244.936 GHz and C34S (5-4) line at 241.016 GHz obtained with the IRAM 30m telescope between 14 March and 8 April 2013. Vertical scale is main beam bright- ness temperature and horizontal scale Doppler velocity in the comet rest frame. 5.5. 32S/34Sratio The wide 8 GHz frequency coverage of the EMIR receiver at IRAM, allows us to observe both C32S and C34S or H32 2 S and H34 2 S lines in the same FTS spectrum since they are less than 4 GHz apart. As a consequence, the isotopic ratios do not de- pend on any calibration or pointing effects and, since the lines are mostly optically thin, the abundance ratio is very close to the line intensity ratio. The integration time on the H2S lines (either around 168 or 216 GHz) was not sufficient to yield a sig- nificant constraint on the H32 2 S ratio. On the other hand, the C34S (5-4) line at 241016.089 MHz was detected in both comets (Figs. 7 and 8). The accuracy of the 32S/34S ratio in CS is given by the signal-to-noise ratio of the C34S line, and we found 32S/34S = 20 ± 5 and 24.7 ± 3.5 in comets Lemmon and Lovejoy, respectively. We do not find any significant departure from the Earth value of 32S/34S=22.5. The C33S line lies between the C34S and C32S lines. The upper limit for the C33S (5-4) line at 242.913 GHz in comet Lovejoy yields 32S/33S> 50. This is the first reported estimate of 33S in a comet, but this is only a lower limit compatible with the Earth value (126.7). 2 S/H34 8 6. Discussion We measured a terrestrial value for the 16O/18O ratio in wa- ter in comet Lovejoy, as was the case in all comets in which this ratio had been measured before (Balsiger et al., 1995; Biver et al., 2007; Bockel´ee-Morvan et al., 2012; Altwegg et al., 2015; Bockel´ee-Morvan et al., 2015). The D/H and 14N/15N ratios in solar system bodies are of par- ticular importance as they vary from object to object, with the heavier elements enriched in the Earth and small bodies com- pared to Jupiter or the protosolar values. These ratios trace the various enrichment and fractionation mechanisms that were at play in the proto-planetary nebula. The other isotopic ratios in- vestigated here do not show significant variations in solar sys- tem objects within the accuracy of our measurements. Figure 9 shows the 14N/15N versus D/H ratios measured in solar system objects, including our new measurements presented here. 6.1. Earth-like water in comet Lovejoy We obtained a measurement or at least a stringent upper limit for the D/H ratio in water of two dynamically old Oort cloud comets. We measured the lowest D/H ratio in water in a comet so Biver et al.: Isotopic ratios of H, C, N, O, S in comets Lemmon and Lovejoy Table 3. Line intensities from IRAM observations and production rates Molecule Transition HCN HCN H13CN H13CN H13CN HC15N HC15N HC15N DCN H2S H34 2 S CS CS C34S C34S C34S HDO HDO HDO Date [yyyy/mm/dd.d − dd.d] Comet C/2012 F6 (Lemmon) March-April 2013 3 − 2 2013/03/14.5-18.6 3 − 2 2013/04/06.46 3 − 2 2013/03/14.5-18.6 3 − 2 2013/04/06.56 3 − 2 2013/03/14.5-37.6 3 − 2 2013/03/14.5-18.6 3 − 2 2013/04/06.46 3 − 2 2013/03/14.5-37.6 3 − 2 2013/03/15.6-38.6 110 − 101 2013/04/08.47 110 − 101 2013/04/08.47 5 − 4 2013/03/14.6-18.6 5 − 4 2013/04/06.56 5 − 4 2013/03/14.6-18.6 5 − 4 2013/04/06.56 5 − 4 2013/03/15.6-38.6 211 − 212 2013/03/14.6-18.6 211 − 212 2013/04/06.56 2013/03/14.6-37.6 211 − 212 Comet C/2014 Q2 (Lovejoy) 13–25 January 2015 2015/01/13.7-16.8 2015/01/13.7-16.8 2015/01/23.73 2015/01/13.7-23.7 2015/01/13.7-16.8 2015/01/23.73 2015/01/13.7-23.7 2015/01/13.7-24.7 2015/01/13.7-16.8 2015/01/13.7-23.7 2015/01/13.7-25.7 2015/01/13.7-25.7 2015/01/13.7-23.7 2015/01/13.7-23.7 2015/01/13.8-25.7 2015/01/13.8-25.7 2015/01/13.8-23.7 2015/01/13.8-16.9 2015/01/23.7-24.7 2015/01/13.8-23.7 2015/01/13.8-16.9 2015/01/23.73 2015/01/13.8-23.7 2015/01/13.8-23.7 2015/01/13.8-16.9 2015/01/23.7-24.7 2015/01/13.8-24.7 3 − 2 HCN H13CN 3 − 2 H13CN 3 − 2 H13CN 3 − 2 HC15N 3 − 2 HC15N 3 − 2 HC15N 3 − 2 3 − 2 DCN CH3OH 52 − 41E 13CH3OH 52 − 41E H2CO 313 − 212 312 − 211 H2CO 414 − 313 HDCO 404 − 303 HDCO 220 − 211 H2S H34 2 S 220 − 211 211 − 202 HDS 5 − 4 CS 5 − 4 CS 5 − 4 CS C34S 5 − 4 C34S 5 − 4 C34S 5 − 4 C33S 5 − 4 211 − 212 HDO HDO 211 − 212 211 − 212 HDO Frequency [MHz] Intensity [mK km s−1] Total production rate [1026 molec.s−1] 265886.432 259011.798 258156.996 217238.538 168762.762 167910.516 244935.557 241016.089 241561.550 265886.432 259011.798 258156.996 217238.538 266838.123 263113.343 211211.469 225697.772 246924.600 256585.531 216710.437 214376.924 257781.410 244935.557 241016.089 242913.610 241561.550 6055 ± 36 5966 ± 50 65 ± 44 43 ± 32 51 ± 26 108 ± 36 11 ± 22 40 ± 19 < 133 234 ± 36 < 66 1477 ± 38 1019 ± 25 50 ± 31 58 ± 16 53 ± 13 88 ± 29 44 ± 19 54 ± 14 10848 ± 39 114 ± 13 86 ± 18 100 ± 12 86 ± 9 85 ± 13 83 ± 7 < 39 1345 ± 24 < 23 978 ± 11 1181 ± 17 < 18 < 21 227 ± 12 16 ± 11 < 21 1049 ± 10 1089 ± 10 1076 ± 7 37 ± 9 53 ± 13 42 ± 6 < 21 31 ± 8 23 ± 11 25 ± 7 15.3 ± 2.1a 13.1 ± 0.8a < 0.28 < 0.20 0.12 ± 0.06 < 0.25 < 0.14 0.09 ± 0.04 < 0.52 77 ± 13a < 21 14.7 ± 0.6a 9.5 ± 0.2 0.51 ± 0.31 0.57 ± 0.16 0.54 ± 0.13 21 ± 7 12 ± 4 13 ± 3 4.9 ± 0.2a 0.049 ± 0.006 0.043 ± 0.009 0.046 ± 0.006 0.037 ± 0.004 0.043 ± 0.007 0.039 ± 0.003 < 0.031 117 ± 3a < 2.4 180 ± 10a 180 ± 10a < 0.38 < 0.36 17.3 ± 0.9a < 2.6 < 0.6 2.14 ± 0.02a 2.51 ± 0.02a 2.30 ± 0.02a 0.077 ± 0.019 0.126 ± 0.031 0.093 ± 0.013 < 0.046 1.8 ± 0.4 1.6 ± 0.7 1.5 ± 0.4 a: Average production rate for the period also taking into account measurements at 5 to 20′′ offsets. For formaldehyde the offset positions were used to constrain the contribution of the distributed source: 80% with a scale-length of 10 000 km. far: 1.4 ±0.4 ×10−4 (< 1.8 ×10−4; 5 σ) in Lovejoy, while the cor- responding ratio could be four times higher in comet Lemmon. These two D/H measurements are based on the same line in two comets of similar activity. Comet Lemmon was slightly more active but also two times farther away. We do not expect that the excitation mechanism of HDO in the coma of these comets can yield a factor of four difference in the derived production rates. So the observed diversity in the D/H ratio can be real and would suggest that these two comets, although they come from the same current reservoir, were formed in different places or at different times in the young solar system. In either case, this brings additional proof that comet dynamical origin does not imply a specific formation region in the disk. Previous studies (Hartogh et al., 2011; Lis et al., 2013; Bockel´ee-Morvan et al., 2012; Altwegg et al., 2015) have shown a growing diversity in the measured values of the D/H ratio in cometary water, and especially variations by more than a factor of 3 among Jupiter family comets, which are thought to originate from the Kuiper 9 Biver et al.: Isotopic ratios of H, C, N, O, S in comets Lemmon and Lovejoy Table 4. Isotopic ratios in comet C/2014 Q2 and C/2012 F6 Isotopic ratio Molecule value Comet C/2012 F6 (Lemmon) Mar.-Apr. 2013 6.5 ± 1.6 × 10−4 D/H Comet C/2014 Q2 (Lovejoy) Jan. 2015 D/H 1.4 ± 0.4 × 10−4 (2010), and Mandt et al. (2009) and Kerridge from Bockel´ee-Morvan et al. Fig. 9. D/H ratios and 14N/15N ratios in solar system objects, compared to the protosolar (gray bands) and Earth (blue bands) values (Robert et al., 2000; Marty et al., 2011). Jupiter and Saturn values are from Marty et al. (2011), Furi & Marty (2015), (2014); Titan values from Abbas et al. and Fletcher et al. (2009); (2010), Niemann et al. (1985); meteorites from Waite et al. and comet values (1998), Meier et al. (2004a), Hartogh et al. (2011), Bockel´ee-Morvan et al. (2012), Bockel´ee-Morvan et al. (2008), Hutsem´ekers et al. (2009), and this paper. The symbols corresponding to each source are given in the enclosed legend box. They are ordered according to the species in which the 14N/15N was measured from left to right: atomic nitrogen, N2 (as for Earth), ammonia, HCN, and CN. Empty symbols with names in italics are for lower limits on the 14N/15N . (1998), Crovisier et al. in comparison to the proto-planetary value (441, Marty et al., 2011) (Fig. 9). The enrichment is similar to that found in NH2 (Rousselot et al., 2014; Shinnaka et al., 2015), thought to be mostly produced by the photo-dissociation of NH3, the main volatile carrier of cometary nitrogen (Biver et al., 2012). Correspondingly low 14N/15N ratios are also detected in car- bonaceous IDPs (Messenger, 2000; Floss et al., 2004). These large enrichments with respect to solar or terrestrial values can be explained by interstellar chemistry theories involving ion-molecule 15N fractionation at 10 K (e.g., Wirstrom et al., 2012). Recently, significant nitrogen fractionation has been mea- sured in dark cloud cores with ratios as low as 150 for nitriles (Milam et al., 2015). Thus, the similarity of the 14N/15N ratios found in comets and IDPs strengthens a putative link to inter- stellar chemistry as the origin of isotopically anomalous organic particles in comets. Nevertheless, our observations confirms that the trend of a twofold enrichment in 15N compared to Earth observed in C15N is also present in HCN. A consequence concerning the debate of the origin of CN in cometary coma is that we cannot ex- clude HCN as the sole parent of CN on the basis of different 14N/15N ratios. 7. Conclusions Chemical diversity is observed in the population of Oort Cloud comets, with abundances varying by up to a factor of ten for several species. The origin of this diversity is unclear and might H2O HCN HCN HCN CS H2S H2O HCN H2S H2CO HCN CH3OH HCN H2O CS H2S CS 12C/13C 14N/15N 32S/34S 12C/13C 14N/15N 16O/18O 32S/34S 32S/33S < 0.045 124 ± 64 or ≥ 89b 152 ± 72 or ≥ 106b 20 ± 5 > 3.5 < 0.006 < 0.017 < 0.007 109 ± 14 > 61 145 ± 12 499 ± 24a 24.7 ± 3.5 > 7 > 50 Note: a: From time variable fits of the production rates. b: based only on the backends with the lowest noise. belt. Now we find that a similar range of values is present among Oort cloud comets. This is in line with the latest mod- els (Brasser & Morbidelli , 2013), which suggest that all comets were formed in the same extended massive original Kuiper Belt beyond the giant planets and were later scattered to the two cur- rent reservoirs, the scattered Kuiper disk and the Oort cloud. The diversity in deuterium enrichment would occur because the pre- solar cloud was initially enriched in HDO (higher D/H in water than in H2: protosolar D/H value in H2 is 2 × 10−5, Fig.9), but the inner warmer parts of the proto-planetary nebula lost more of their initial enrichment in deuterated water. 6.2. 14N/15Nincomets Measurements of 14N/15N in comets are relatively scarce. Although the 14N/15N ratio is 5–20 times lower than the H16 2 O /HDO ratio, the nitrogen bearing molecules in cometary comae are less than 1% in abundance relative to water. So the detection of 15N isotopologues requires very active comets. The most sensitive technique is optical high resolution spectroscopy, which has provided several detections of C15N (Manfroid et al., 2005, 2009; Hutsem´ekers et al., 2009) and recently detections of 15NH2 (Rousselot et al., 2014; Shinnaka et al., 2015), but this technique only gives access to daughter species. The com- pounds N2 and NH3 are not very abundant (Rubin et al., 2015; Biver et al., 2012) and do not have strong lines observable from the ground, nor do their 15N isotopologues. Only HCN is eas- ily detectable in the radio and infrared. We report here the third clear (S/N ¿ 10) detection of HC15N in a comet after its detection in comets Hale-Bopp and 17P/Holmes (Bockel´ee-Morvan et al., 2008). We measure the same enrichment in 15N in HCN in comet Lovejoy as in other comets, in spite of their different dynamical origins (Bockel´ee-Morvan et al., 2008). The same enrichment was also found in CN in many other comets (Hutsem´ekers et al., 2009). This is a factor of two enrichment in 15N in compari- son to the Earth atmospheric value (272) and a factor of three 10 Biver et al.: Isotopic ratios of H, C, N, O, S in comets Lemmon and Lovejoy reflect comet formation at different places or times in the early solar system. The observations presented here provide further evidence that diversity is present in the enrichment in deuterium in cometary water with respect to the protosolar value (HD/H2), both in Jupiter family and Oort cloud comets, in line with the latest scenarios of comet origins (Brasser & Morbidelli , 2013). On the one hand, we found Earth-like D/H, 16O/18O , 12C/13C , and 32S/34S ratios in comet C/2014 Q2 (Lovejoy), strengthening the role that some comets may have played in supplying mate- rial to the young Earth, especially complex organic molecules (Biver et al. , 2015). On the other hand we confirm the trend of finding a systematic twofold enrichment in 15N in cometary HCN relative to its abundance in Earth N2, whose origin is puz- zling. We were also able to obtain in comet C/2014 Q2 sensitive limits, some of the best obtained so far with remote observations, on isotopic ratios in other molecules such as D/H in H2S, H2CO, or C32S/C33S. Acknowledgements. The IRAM observations were conducted under the target of opportunity proposal D04-14 and regular proposal 128-14 and we gratefully acknowledge the support from the IRAM director for awarding us discretionary time and the IRAM staff for its support and for scheduling the observations on short notice. This research has been supported by the Programme national de plan´etologie de l'Institut des sciences de l'univers (INSU). SNM acknowledges the NASA Planetary Astronomy program for support. References Abbas, M. M., Kandadi, H., LeClair, A., et al., 2010, ApJ, 708, 342 Altwegg, K. & Bockel´ee-Morvan, D., 2003, Space Sci. Rev., 106, 139 Altwegg, K., Balsiger, H., Bar-Nun, A., et al., 2015, Science, 347, 1261952 Balsiger, H., Altwegg, K. & Geiss, J., 1995, J. Geophys. Res., 100, 5827 Biver, N., Bockel´ee-Morvan, D., Crovisier, J., et al. 2000, AJ, 120, 1554 Biver, N., Bockel´ee-Morvan, D., Crovisier, J., et al. 2006, A&A, 449, 1255 Biver, N., Bockel´ee-Morvan, D., Crovisier, J., et al., 2007, Planet. Space Sci., 55, 1058 Biver, N., Bockel´ee-Morvan, D., Colom, P., et al., 2011, A&A, 528, A142 Biver, N., Crovisier, J., Bockel´ee-Morvan, D., et al., 2012, A&A539, A68 Biver, N., Bockel´ee-Morvan, D., Debout, V., et al., 2014, A&A, 566, L5 Biver, N., Bockel´ee-Morvan, D., Moreno, R., et al., 2015, Science Advances, 1:e1500863 Bockel´ee-Morvan, D., Gautier, D., Lis, D.C., et al. 1998, Icarus, 133, 147–162 Bockel´ee-Morvan, D., Lis, D.C., Wink, J.E., et al. 2000, A&A, 353, 1101 Bockel´ee-Morvan, D., Biver, N., Jehin, E., et al. 2008, ApJ, 679, L49 Bockel´ee-Morvan, D., Biver, N., Swinyard, B., et al., 2012, A&A, 544, L15 Bockel´ee-Morvan, D., Calmonte, U., Charnley, S., Duprat, J., Engrand, C., et al., 2015, Space Sci. Rev., 197, 47 Brasser, R., & Morbidelli, A., 2013, Icarus, 225, 40–49 Carter, M, Lazareff, R., Maier, D., et al. 2012, A&A, 538, A89 Combi, M.R., Bertaux, J.-L., Qu´emerais, E., et al. 2014, AJ, 147, A126 Crovisier, J., 1994, J. Geophys. Res., 99-E2, 3777-3781 Crovisier, J., Bockel´ee-Morvan, D., Biver, N., et al. 2004a, A&A, 418, L35 Crovisier, J., Bockel´ee-Morvan, D., Colom, P., et al. 2004b, A&A, 418, 1141 Fletcher, L.N., Greathouse, T.K., Orton, G.S., et al., 2014, Icarus, 238, 170 Floss, C., Stadermann, F. J., Bradley, J., et al. 2004, Science, 303, 1355 Frisk, U., Hagstrom, M., Ala-Laurinaho, J. et al., 2003, A&A, 402, L27 Furi, E. & Marty B., 2015, Nature Geoscience, 8, 515–522 G´erard, E., Crovisier, J., Colom, P., et al. 1998, Planet. Space Sci., 46, 569-577 Hartogh, P., Lis, D.C., Bockel´ee-Morvan, D., et al. 2011, Nature, 478, 218 Hutsem´ekers, D., Manfroid, J., Jehin, E., & Arpigny, C. 2009, Icarus, 204, 346 Jewitt, D., Matthews, H. E., Owen, T. & Meier, R., 1997, Science, 278, 90 Kerridge, J. F., 1985, Geochim. Cosmochim. Acta, 49, 1707–1714 Lis, D. C., Biver, N., Bockel´ee-Morvan, D., et al. 2013, ApJ, 774, L3 Mandt, K.E., Waite, J. H., Lewis, W., et al., 2009, Planet. Space Sci., 57, 1917– 1930 Mandt, K.E., Mousis, O., Lunine, J., & Gautier, D., 2014, ApJ, 788, L24 Manfroid, J., Jehin, E., Hutsem´ekers, D., et al. 2005, A&A, 432, L5 Manfroid, J., Jehin, E., Hutsem´ekers, D., et al. 2009, A&A, 503, 613 Marty, B., Chaussidon, M., Wiens, R. C., Jurewicz, A. J. G., & Burnett, D. S., 2011, Science, 332, 1533 Messenger, S., 2000, Nature, 404, 968 Meech, K.J., and 196 co-authors, 2011, ApJ, 734, L1 Meier, R., Owen, T. C., Matthews, H.E., et al., 1998, Science, 279, 842 Milam, S. N., Adande, G., Cordiner, M. A., Wirstrom, E. & Charnley, S. B., 2015, LPI Contribution, 1832, 1934 Muller, H. S. P., Schloder, F., Stutzki J. & Winnewisser, G. 2005, J. Mol. Struct., 742, 215–227 (http://www.astro.uni-koeln.de/cdms/) Niemann, H. B., Atreya, S. K., Demick, J. E., et al. 2010, J. Geophys. Res., 115, E12, E12006 O'Brien, D.P., Walsh, K.J., Morbidelli, A., Raymond, S.N. & Mandell, A.M., 2014, Icarus, 239, 74–84 Paganini, L., DiSanti, M. A., Mumma, M. J., et al. 2014, AJ, 147, A15 Pickett, al. Poynter, Cohen, H.M., R.L., E.A., et 1998, J. Quant. Spec. Radiat. Transf., 60, 883–890 (http://spec.jpl.nasa.gov/) Robert, F., Gautier, D. & Dubrulle, B., 2000, Space Sci. Rev., 92, 201 Rousselot, P., Pirali, O., Jehin, E., et al. 2014, ApJ, 780, L17 Rubin, M., Altwegg, K., Balsiger, H., et al., 2015, Science, 348, 232-235 Shinnaka, Y., Kawakita, H., Kobayashi, H., et al. 2015, ApJ, 782, L16 Waite, J. H., Jr., Lewis, W. S., Magee, B. A., et al. 2009, Nature, 460, 487–490 Wirstrm, E. S., Charnley, S. B., Cordiner, M. A. & Milam, S. N., 2012, ApJ, 757, L11 Zakharov, V., Bockel´ee-Morvan, D., Biver, N., Crovisier, J., Lecacheux, A. 2007, A&A, 473, 303–310 11
1605.03022
1
1605
2016-05-10T14:08:58
Development of HPLC-Orbitrap method for identification of N-bearing molecules in complex organic material relevant to planetary environments
[ "astro-ph.EP" ]
Although the Cassini Spacecraft and the Huygens lander provided numerous information about Titan atmospheric chemistry and the formation of its aerosols, the exact composition of these aerosols still remains unknown. A fruitful proxy to investigate these aerosols is the use of laboratory experiments that allow producing and studying analogs of Titan aerosol, the so35 called tholins. Even when produced in the laboratory, unveiling the exact composition of the aerosol remains problematic due to the high complexity of the material. Numerous advances have been recently made using high-resolution mass spectrometry (HRMS) (Pernot et al. 2010, Somogyi et al. 2012, Gautier et al. 2014) that allowed the separation of isobaric compounds and a robust identification of chemical species composing tholins regarding their molecular formulae. Nevertheless isomeric species cannot be resolved by a simple mass measurement. We propose here an analysis of tholins by high performance liquid chromatography (HPLC) coupled to HRMS to unveil this isomeric ambiguity for some of the major tholins compounds. By comparing chromatograms obtained when analyzing tholins and chemical standards, we strictly identified seven molecules in our tholins samples: melamine, cyanoguanidine, 6-methyl-1,3,5-triazine-2,4-diamine, 2,4,6-triaminopyrimidine, 3-amino- 1,2,4-triazole, 3,5-Dimethyl-1,2,4-triazole and 2,4-diamino-1,3,5-triazine. Several molecules, including hexamethylenetriamine (HMT) were not present at detectable levels in our sample. The use for the first time of a coupled HPLC-HRMS technique applied to tholins study demonstrated the interest of such a technique compared to single high-resolution mass spectrometry for the study of tholins composition.
astro-ph.EP
astro-ph
Development of HPLC-Orbitrap method for identification of N-bearing molecules in complex organic material relevant to planetary environments Thomas Gautier*1,2, Isabelle Schmitz-Afonso3,4, David Touboul3, Cyril Szopa2,5, Arnaud Buch6 , Nathalie Carrasco2,5 1 NASA Goddard Space Flight Center, code 699, 8800 Greenbelt Rd, Greenbelt, MD 20771, USA 2 LATMOS/IPSL, UVSQ Université Paris-Saclay, UPMC Univ. Paris 06, CNRS, Guyancourt, France 3Institut de Chimie des Substances Naturelles ICSN-CNRS UPR 2301, Université Paris-Sud, 1 avenue de la terrasse, 91198, Gif-sur-Yvette, France 4Normandie Université, COBRA, UMR 6014 et FR3038, Université de Rouen ; INSA de Rouen ; CNRS, IRCOF, 1 rue Tesnière, 76821 Mont-Saint-Aignan Cedex, France 5Institut Universitaire de France, 103 Bvd St-Michel, 75005 Paris, France 6LGPM, Ecole Centrale de Paris, Grande voie des Vignes, 92295 Chatenay-Malabry Cedex, France Corresponding author, current address: *Thomas Gautier NASA – Goddard Space Flight Center 8800 Greenbelt Road Greenbelt, MD 20771 Phone: Mail: [email protected] Highlights ➢ ➢ ➢ ➢ We analyze Titan's tholins using HPLC-Orbitrap We strictly identify the isomers of seven of the major molecules constituting tholins All confirmed molecules bear nitrogen and most of them are aromatics. This supports the hypothesis of a tholins formation passing through PANH 1 Abstract Although the Cassini Spacecraft and the Huygens lander provided numerous information about Titan atmospheric chemistry and the formation of its aerosols, the exact composition of these aerosols still remains unknown. A fruitful proxy to investigate these aerosols is the use of laboratory experiments that allow producing and studying analogs of Titan aerosol, the so- called tholins. Even when produced in the laboratory, unveiling the exact composition of the aerosol remains problematic due to the high complexity of the material. Numerous advances have been recently made using high-resolution mass spectrometry (HRMS) (Pernot et al. 2010, Somogyi et al. 2012, Gautier et al. 2014) that allowed the separation of isobaric compounds and a robust identification of chemical species composing tholins regarding their molecular formulae. Nevertheless isomeric species cannot be resolved by a simple mass measurement. We propose here an analysis of tholins by high performance liquid chromatography (HPLC) coupled to HRMS to unveil this isomeric ambiguity for some of the major tholins compounds. By comparing chromatograms obtained when analyzing tholins and chemical standards, we strictly identified seven molecules in our tholins samples: melamine, cyanoguanidine, 6-methyl-1,3,5-triazine-2,4-diamine, 2,4,6-triaminopyrimidine, 3-amino- 1,2,4-triazole, 3,5-Dimethyl-1,2,4-triazole and 2,4-diamino-1,3,5-triazine. Several molecules, including hexamethylenetriamine (HMT) were not present at detectable levels in our sample. The use for the first time of a coupled HPLC-HRMS technique applied to tholins study demonstrated the interest of such a technique compared to single high-resolution mass spectrometry for the study of tholins composition. Keywords: Titan's atmosphere; Atmospheres, chemistry; Organic chemistry; Prebiotic chemistry 2 1. Introduction The atmosphere of Titan is mainly constituted of N2 and CH4. Solar irradiation and electrically charged particles accelerated in Saturn's magnetosphere (Sittler Jr. et al. 2009) induce organic chemical reactions within the atmosphere. These reactions lead to the production in Titan's atmosphere of an opaque layer of organic solid aerosol. To study these Titan's aerosols, it is possible to produce laboratory analogs, the so-called "tholins". A discussion on the different experimental setups designed for such a purpose can be found in Cable et al. (2012). The properties of the produced tholins allow a better analysis and understanding of observational data obtained in the atmosphere of Titan. Recent advances in HRMS, with OrbitrapTM or Fourier-Transform Ion Cyclotron Resonance mass spectrometers, gave a first proxy on the aerosol constitution and revealed that tholins are formed of thousands of different chemical compounds (Pernot et al. 2010, Somogyi et al. 2012, Gautier et al. 2014). However, each detected mass peak only gives access to the raw molecular formula of the compound whereas, especially for high masses, one formula can be attributed to several different isomeric molecules bearing highly different structures and reactivity. It is then of prime importance to be able to strictly identify the molecules within the tholins, by separating all the isomers. The separation methods mostly used for tholins analysis are pyrolysis GC/MS (Khare et al. 1984, McGuigan et al. 2006, Coll et al. 2013) and GC/MS (Pilling et al. 2009, He and Smith 2014). Thin-layer chromatography prior to mass spectrometry (Jagota et al. 2014) and microelectrophoresis (Cable et al. 2014) were recently used whereas very few liquid chromatography separations of tholins have been published (Ruiz-Bermejo et al. 2008, Cleaves et al. 2014). As mentioned by Cable et al. 2012, liquid chromatography could provide a more comprehensive analysis of the tholins molecular structure with a separation for both polar and charged species. 3 Therefore, to go further in tholinomics and unveil the degeneracy of the molecular formulae, we separated the tholins soluble fraction by high-performance liquid chromatography (HPLC) coupled to a hybrid linear trap/OrbitrapTM mass spectrometer and compared retention times with those of standard compounds. The method presented in this paper has been developed for Titan's tholins but it could be applied to any complex organic material in the solar system such as meteorite organic soluble matter, kerogens, interstellar ice analogs etc. 2. Experimental 2.1. Tholins preparation and HPLC-HRMS developments Tholins were produced with a cold plasma in a 95-5% N2-CH4 mixture using the PAMPRE reactor (Szopa et al. 2006). They were produced and prepared following the procedure described in Gautier et al. (2014). These tholins were previously analyzed by direct HR-MS in Gautier et al. 2014. For mass spectrometric analysis, tholins were first dissolved in methanol (HPLC grade, Baker) at a concentration of 4 mg.mL-1, then mixed thoroughly. The resulting solution was filtered through a PTFE 0.2 µm membrane to remove remaining tholins that did not dissolve in methanol (Carrasco et al. 2009). HPLC was performed with a HPLC Ultimate 3000 system (Dionex). Three different columns were tested to achieve proper and complementary separation. First we used an octadecyl C18 column 2.1 mm internal diameter (i.d.), 100 mm, particle size 3.5 µm (Zorbax SB-Aq, Agilent technologies), equipped with a guard column and thermostated at 25°C. Elution was performed with a mobile phase of pH=6.85 consisting of 20 mM of ammonium acetate in water (A) and acetonitrile (HPLC grade, Baker) (B) with the following gradient: 0 to 2 min, isocratic with 0% B – 2 to 11 min, 0 to 21% B - 11 to 13 min, 21 to 100% B then maintaining 3 min at 100% B to rinse the column and finally return to 4 initial conditions and stabilization of the column. Second, a cyano column 2.1 mm i.d., 150 mm, particle size 5 µm (Zorbax Eclipse XDB-CN, Agilent technologies), thermostated at 25°C, was tested. Elution was performed with a mobile phase consisting of 20 mM of ammonium acetate in water (A) and acetonitrile (B) with the following gradient: 0 to 3 min, isocratic with 4% B – 3 to 15 min, 4 to 100% B then maintaining 3 min at 100% B to rinse the column and finally return to initial conditions and stabilization of the column. The third column was an aminopropyl column 2 mm i.d., 150 mm, particle size 3 µm (Luna NH2, Phenomenex), thermostated at 20°C. This column was operated in hydrophilic interaction chromatography mode with a mobile phase consisting of 20 mM of ammonium acetate in water (A) and acetonitrile (B): 0 to 10 min, 95 to 73% B – 10 to 17 min, 73 to 50% B then maintaining 8 min at 50% B to rinse the column and finally return to initial conditions and stabilization of the column. The tholins extract was diluted by a factor of two to get a sample in water/methanol (50/50), for the reversed phase separations (octadecyl and cyano columns) and in acetonitrile/methanol (50/50) for the hydrophilic interaction chromatography (aminopropyl column). For the three columns, flow rate was set to 0.25 mL.min-1 and injection volume was 7 µL. The HPLC was directly coupled to a hybrid linear trap/OrbitrapTM (LTQ/OrbitrapTM, ThermoScientific) mass spectrometer equipped with an ElectroSpray Ionization (ESI) source. Analyses were performed in the positive ion mode. Acquisition parameters of the ESI- LTQ/OrbitrapTM were: needle voltage 4.5 kV; capillary temperature 275°C; capillary voltage 25 V; tube lens voltage 65 V; sheath gas flow rate 40 arbitrary unit (a.u.). The mass spectrometer was externally calibrated using caffeine, MRFA (met-arg-phe-ala) peptide and Ultramark 1621 allowing a mass precision below 2 ppm. Data were acquired with a mass resolution set to 100 000 at m/z 450. The experimental resolution m/Δm was determined to be ~200 000 at m/z 150 for both MS and MS/MS measurements. 5 2.2 Molecules of interest; choice of standards In the present work we aim to determine the exact composition of tholins produced with PAMPRE (introducing 5% of methane in dinitrogen for the gas mixture) by determining the exact composition (formula and structure) of the most abundant peaks detected in the high- resolution mass spectrum presented in Figure 1 (Gautier et al. 2014). Samples analyzed in the present work were produced following the same protocol as the one described in Gautier et al. 2014. Figure 1: Mass spectrum of an extract of tholins (infusion of the sample) produced with 5% of methane in nitrogen. Numbers in the figure indicate the most intense ion for each cluster. Adapted from Gautier et al. 2014 The interest of the method proposed here is to unveil the isomeric ambiguity remaining after high-resolution Orbitrap analysis and exact mass determination. However, the high complexity of the mass spectrum of tholins obtained with the Orbitrap (>15,000 peaks detected) makes an exhaustive analysis of all peaks impossible. We focused here on a limited set of compounds to be tested with our method. 6 The compounds of interest for HPLC-MS analysis are chosen based on two criteria. As a first approach, we limit this study to the most intense ions detected in this spectrum that might represent a significant portion of the material composing sample (see table 1). Second, we choose the standard compounds depending on their availability and affordability for comparison of retention time. The complete set of compounds tested can be find in Table 1 for the compounds detected in tholins and in Table SI 1 for those non confirmed to be present in tholins. Standards 1, 2, 5, 8, 9 and 10 were purchased from VWR. Standards 3, 4, 6, 7, 11 and 12 were purchased from Sigma Aldrich. Standard 13 was purchased from Fluka. All stock solutions were prepared in water except for standards 7, 10, 12 and 13 prepared in MeOH/water, 50/50, v/v to be able to solubilize them. After dilution in methanol/water or methanol/acetonitrile, each individual standard and a mixture of all standards were analyzed in the three chromatographic conditions developed with the tholins extract. Validation of the presence of the standard in the tholins extract was done by retention time comparison. Concentrations tested ranged from 3.10-4 to 1.10-2 mg.mL-1, varying for each compound. 3. Results The three columns, octadedyl, cyano and aminopropyl, have been implemented for complementary chromatographic separation in regard with known tholins chemical properties. The octadecyl column, with an apolar stationary phase, has been chosen for reversed phase separation starting with 100% aqueous phase to get maximum retention of polar analytes, with retention time increasing with decreasing polarity of the analyzed compounds. The cyano column with a more polar stationary phase, dimethyl-cyanopropylsilane, is especially suited for separation of highly polar, acidic and basic compounds with better retention of polar analytes through interactions with the stationary phase in reversed phase mode. The 7 aminopropyl column has been selected to work with hydrophilic interaction chromatography (HILIC), thus inducing different retention capabilities (Buszewski and Noga 2012). Two types of interaction may occur in HILIC mode in our study: hydrophilic interactions with retention increasing with polarity, and electrostatic repulsion between positively charged analytes and the positively charged stationary phase. In the particular case of our basic compounds, no hydrogen bonding may occur with the stationary phase. Furthermore, HILIC is particularly adapted for polar nitrogen compounds, which are one of the major parts of the tholins composition (Zheng et al 2012). 3.1. Interest of the method Figure 2 provides the Base Peak Chromatogram (BPC) of tholins sample using the aminopropyl column. This congested BPC of tholins denotes the need of both separation with HPLC and high resolution in mass with OrbitrapTM to isolate precisely compounds from the bulk material in order to identify them. #106 10 y t i s n e t n I 8 6 4 2 0 0 2 4 6 Time (min) 8 10 12 14 Figure 2: Base Peak Chromatogram (BPC) Intensity chromatogram of a tholins sample analyzed with the aminopropyl column 8 The complete set of chromatograms is given in the supplementary information together with an example of blank experiment (figure SI7). We provide here an example set of chromatograms at m/z 85 and 98 in the tholins extract and in the mixture of standards, with separation by aminopropyl, cyano and octadecyl columns. Figure 3 shows the different extracted ion chromatograms (EIC) for m/z 85. Figure 3a represents the elution using an aminopropyl column, 3b using a cyano column and 2c using a C18 column. Figure 4 represents the same set of data for m/z 98. These two figures clearly show the complementarity of the different stationary phases for the separation of the different analytes presents in the tholins. Figure 3 demonstrates that the aminopropyl column leads to the best separation of isomers at m/z 85 present in tholins, as well as the two standards. On figure 4, standards 5 and 6 are better separated using the C18 column. Standard 5 retention time does not match with any tholins compound with the C18 column whereas it seems to match with a tholins compound with the amino and cyano columns. Standard 6 is confirmed with the three columns. The aminopropyl column evidences many isomeric compounds for the same exact mass. Regarding the eight m/z ratio studied here (Other EIC are presented in the Supplementary Information), the most efficient column is the aminopropyl column with around 39 different isomers separated, then the octadecyl column with 25 compounds and finally the cyano column with 20 compounds. Only major compounds with a signal-to-noise ratio above 100 (Xcalibur Thermo software, automatic integration) were taken into account for each EIC. The aminopropyl stationary phase in hydrophilic chromatography seems the most powerful for the separation of tholins isomers but octadecyl phase in reversed phase mode stays useful for complementary separation for some isomers. 9 #3- #4- #3-and-#4-co=eluted- #3- #4- Tholins- Mix-std- Tholins- Mix-std- Tholins- Mix-std- (a)  Aminopropyl- - (b)-Cyano- - (c)-Octadecyl- Figure 3: Extracted ion chromatograms of m/z 85.0500 to 85.0518, tholins sample and standards 3 and 4, for columns (a) aminopropyl (b) cyano (c) octadecyl. 10 #5- #6- #5- #6- #3- #4- #5- #6- Tholins- Mix-std- Tholins- Mix-std- Tholins- Mix-std- (a)  Aminopropyl- - (b)-Cyano- - (c)-Octadecyl- Figure 4: Extracted ion chromatograms of m/z 98.0706 to 98.0720, tholins sample and standards 5 and 6, for columns (a) aminopropyl (b) cyano (c) octadecyl. 11 3.2. Standard confirmation The retention time obtained both for the tholins sample and the standard tested are summarized in table 1. A complete set of chromatograms is given in supplementary material. Table 1: Column one to six respectively give the m/z ratio, the standard number, the name, the structure, the CAS Number and the Formulae of standard compounds tested and confirmed to be present in tholins. Retention time of sample and confirmed standard on the three columns are given columns seven to twelve. Values correspond to the average on three injections and displayed uncertainties were calculated at 1 σ. A list of retention time for compounds not confirmed to be present in tholins can be find in Table SI 1 of the supplementary material. [M+H]+ m/z Std Name Structure CAS Number Formulae Aminopropyl Rt std (min) Rt 85.0509 98.0713 112.0618 126.0774 4 3 5 6 7 8 9 Cyanoguanidine 3-amino-1,2,4- triazole 5-methyl-3- amino-1,2- pyrazole 3,5-Dimethyl- 1,2,4-triazole 2,4-diamino- 1,3,5-triazine 6-methyl-1,3,5- triazine-2,4- diamine 2,4,6- triaminopyrimidi ne 127.0727 10 Melamine 461-58-5 C2N4H5 61-82-5 C2N4H5 31230-17- 8 C4N3H8 7343-34-2 C4N3H8 504-08-5 C3N5H6 542-02-9 C4N5H8 1004-38-2 C4N5H8 108-78-1 C3N6H7 3.97 ± 0.08 5.31 ±0.00 2.85 ±0.01 3.26 ±0.00 4.16 ±0.00 4.21 ±0.02 6.28 ±0.00 6.97 ±0.02 tholins (min) 3.95 ± 0.02 5.33 ± 0.00 2.81 ± 0.02 3.26 ± 0.03 4.16 ± 0.00 Rt std (min) 1.93 ±0.02 1.94 ±0.02 2.27 ±0.00 2.79 ±0.02 2.31 ±0.01 2.32 ±0.03 2.82 ±0.00 2.29 ±0.00 Octadecyl Rt Rt std (min) Cyano Rt tholins (min) 1.93 ±0.00 1.93 ±0.00 tholins (min) 1.85 ± 0.00 2.13 ± 0.00 4.36 ± 0.02 7.15 ±0.03 6.67 ± 0.01 1.85 ±0.02 2.14 ±0.01 3.41 ±0.00 7.17 ±0.02 6.65 ±0.00 4.22 ± 0.02 2.77 ±0.00 3.22 ±0.03 6.31 ± 0.02 2.77 ±0.00 8.04 ±0.02 6.29 ±0.02 3.24 ±0.00 8.04 ± 0.02 6.31 ± 0.02 6.99 ± 0.02 2.19 ±0.00 2.21 ±0.00 6.08 ±0.00 6.10 ± 0.00 1 2 Using the HPLC coupling with high resolution MS we have been able to identify strictly the following molecules in our tholins sample: 3-amino-1,2,4-triazole, cyanoguanidine, 3,5- 12 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 Dimethyl-1,2,4-triazole, 2,4-diamino-1,3,5-triazine, 6-methyl-1,3,5-triazine-2,4-diamine, 2,4,6-triaminopyrimidine and melamine (standards 3, 4, 6 to 10 respectively). To our knowledge, and with exception of Cyanoguanidine and Melamine, none of these molecules have been previously reported in the literature to be present in tholins. These results highlighted the importance of separation methods coupled to mass spectrometry to characterize the compounds present in tholins as they present many isomeric species. In direct introduction methods, even with a high-resolution instrument, these isomeric compounds cannot be differentiated and MS/MS spectra would be representative of several compounds thus inducing interpretation errors. Our results indicate that 1-methyl-1,2,4-triazole, 3- aminopyrazole, (S)-5-(2-pyrrolidinyl)-1H-tetrazole, 4-amino-2-dimethylamino-1,3,5-triazine and HMT (standards 1, 2, 11, 12 and 13, respectively) are not detected in our tholins sample with the analytical method set up. For standard 5, the result is more ambiguous since there are many isomers detected in the tholins sample; this standard would be confirmed only on the aminopropyl and cyano column. As we consider that the three columns should validate each compound, we would conclude that 3-amino-5-methylpyrazole (standard 5) is not detected in PAMPRE tholins. To provide a rough estimate of compounds concentration in tholins soluble fraction we used the standard addition method. We measured the area of a given compound in the tholins chromatogram (At) and the area of the same peak after adding a known amount (Cs) of standard compound in the tholins sample (Ats). Concentration of this compound in tholins solution can therefore be estimated as Ct=(At*Cs)/(Ats-At). Concentrations for compounds detected in tholins are given in table 2. We would like to emphasize though that these values represent only a rough estimate of compound concentration in tholins solution. Given the fact that tholins are not 100% soluble (Carrasco et al. 2009) and that solubility of tholins might differ from one compound to another, it is impossible to interpret these values as actual 13 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 concentration within the bulk tholins. However, values provided in table 2 can be used as lower limits for compounds concentration in tholins. Table 2: Calculation of identified compounds concentration in tholins soluble fraction (Ct). Cs represents the known concentration of standard added, Ats the area of the chromatographic peak for tholins + standard (averaged on 3 injections), and At the area of the peak for tholins solely (averaged on 3 injections). [M+H]+ m/z Std number Cs (mg.mL-1) Ats At Ct (mg.mL-1) 85.0509 98.0713 112.0618 126.0774 4 3 6 7 8 9 127.0727 10 5.00E-03 2.50E-03 2.50E-04 5.00E-03 2.50E-04 5.00E-04 1.25E-03 5.32E+07 7.36E+07 3.07E+07 1.86E+08 1.27E+08 2.89E+07 3.76E+07 2.72E+07 2.07E+07 5.00E+06 6.07E+07 1.03E+08 9.17E+06 1.88E+07 5.2E-03 9.8E-04 4.9E-05 2.4E-03 1.1E-03 2.3E-04 1.3E-03 A compound of interest we were particularly looking for in our sample was HMT (standard 13). HMT is a molecule detected in the final products of many astrophysical laboratory experiments, especially the one simulating interstellar ice irradiations (Bernstein et al. 1995, Muñoz Caro et al. 2003, Vinogradoff et al. 2012). The detection of HMT in laboratory tholins a priori similar to our samples has been reported recently by He et al. (2012) using NMR and high resolution mass spectrometry and was proposed to be a major component of their tholins. In our case a major product at m/z 141.1131 was detected corresponding to the exact mass (m/zth. 141.1135; ∆m/z = -2.4 ppm) and thus formulae of protonated HMT. However our HPLC results show that no signal at m/z 141 was detected in the tholins sample at the retention time 14 44 45 46 47 48 49 of the HMT standard (Figure SI6). Furthermore, MS/MS spectra of the isomers in tholins sample (Figure 5a) and HMT (Figure 5b) show different fragmentation patterns. Fragmentation pattern also does not match any HMT derivative detected in ice irradiation (Danger et al. 2013). For the tholins sample, the losses of NH3 and HCN, already described in MS/MS experiments of tholins (Somogyi et al. 2005), would suggest either the presence of NH2 function and/or a cyano function and/or a cycle including a nitrogen atom. 50 51 52 Figure 5: LC-MS/MS spectra of m/z 141.1131 (C6H13N4) for (a) tholins isomer (b) HMT, recorded with the same collision energy, CE = 24 a.u., activation qz 0.25. 15 53 54 55 56 57 58 59 60 61 62 63 64 65 66 67 68 69 70 71 72 73 74 75 76 This clearly indicates that, at least in our sample, the major compound detected at m/z 141.1131 cannot be attributed to HMT. The isomeric compound detected in our tholins sample is less polar than HMT regarding its higher retention time for octadecyl column and shorter retention time for aminopropyl column. 4. Discussion This difference between our results and the one from He et al. (2012) is difficult to understand as the pathway they proposed for HMT formation passes through gas phase reactions of methanimine. Methanimine has been firmly detected as a volatile compound present in the PAMPRE reactor during standard experiments (Carrasco et al. 2012), that would lead to the formation of HMT according to He et al. proposed mechanism. Considering the complexity of the molecular identification in tholins samples (which required comparisons to standards in our case), a first explanation for the discrepancies between the experiments might then be that the compound identified by He et al. (2012) using NMR is not HMT but one of its isomer, as in PAMPRE tholins. Another possibility would be that HMT is actually produced in the case of He et al (2012); but with a mechanism specific to this experiment. Indeed He et al. 2012 used an AC plasma, where the solid samples remains in the reactor for the whole duration of the synthesis whereas the tholins in PAMPRE are dragged out of the plasma after a few minutes. The solid samples are therefore longer exposed to the plasma bombardment in the case of He et al (2012). Moreover in the case of He et al. 2012, tholins are not collected directly after production at cold temperature but are warmed up to room temperature for 24 h before collection. Recent investigation on PAMPRE showed that such a process tends to form a complex organic residue in addition to the tholins by reaction taking place within the solid state (Gautier et al. 2014). These secondary reactions may explain the formation of HMT, whose formation 16 77 78 79 80 81 82 83 84 85 86 87 88 89 90 91 92 93 94 95 96 97 98 99 pathways are allowed by thermal reactivity in cometary ice analogs (Vinogradoff et al. 2011, Vinogradoff et al. 2012, Theulé et al. 2013). Among the compounds detected in our sample are cyanoguanidine (standard 4) and melamine (standard 10). HPLC-LTQ/OrbitrapTM provides here a firm detection of these two compounds recently suggested to be present in tholins by NMR studies, in sample produced both with PAMPRE (Derenne et al. 2012) and with another experimental setup (He and Smith 2013, He and Smith 2014). As stated by He et al. 2014, cyanoguanidine and melamine are a dimer and a trimer of cyanamide, respectively. These two compounds may be of interest regarding aerosols formation in Titan's atmosphere. In particular, melamine is well known for its polymeric properties and tends to form large-scale structures of highly cross-linked material (Wohnsiedler 1953, Philbrook et al. 2005, Merline et al. 2013) that could serve as a base structure for aerosol growth. Regarding the molecules present in our tholins, all have high nitrogen content and six out of seven are aromatic molecules. As a reminder these molecules correspond to the most predominant peaks in the mass spectrum of tholins recorded with an OrbitrapTM analyser. Even though N-bearing compounds do not form the entire material (some pure hydrocarbons are detected by OrbitrapTM analyser, see Pernot et al. (2010), Gautier et al. (2014)), it is clear that N-bearing aromatics constitute a significant part of tholins material. Triazine-like cycle is a predominant feature in the compounds detected in this study. This is in agreement with results from Quirico et al. (2008) that also proposed triazine as an important contribution to tholins signature in infrared absorption spectroscopy. The predominance of such compounds in tholins is a possible indication of N-Bearing Polyaromatic Hydrocarbons (PANH) 100 formation pathway, rather than a pure Polyaromatic Hydrocarbons (PAH) one. 17 101 5. Conclusion 102 We used a complementary HPLC-OrbitrapTM HRMS analysis on Titan's tholins and standard 103 compounds. Our study confirms the importance of a proper chromatographic separation for 104 the analysis of complex (several thousands of peaks in the MS) organic mixtures, such as 105 tholins, where a single high resolution MS peak can be due to the contribution of a dozen of 106 different isomers. We saw that in such a case two chromatographic columns, at the very least, 107 are necessary for a proper decomposition of the material due to the large variability in polarity 108 of the compounds composing tholins. In case of using only two columns we would 109 recommend the C18 in reversed phase mode and the aminopropyl in hydrophilic interaction 110 mode due to their complementarity. We expect the approach developed in this paper to be 111 easily applicable to other organic complex mixtures of interest for planetary sciences, such as 112 meteorites soluble organic matter or laboratory analogues of cosmic ices. 113 On our sample, we were able to strictly identify seven isomers of the major compounds 114 present in Titan's tholins, and to discard six others including HMT, based on three 115 complementary chromatographic separation coupled to HRMS analysis. All the detected 116 molecules bear nitrogen within their structure and most of them are aromatics. This supports 117 the idea of a tholins formation pathways passing through PANH. 118 119 120 121 122 123 Funding sources The research presented in this paper was partially funded through the French Programme National de Planétologie (PNP). NC acknowledges the European Research Council for their financial support (ERC Starting Grant PRIMCHEM, grant agreement n°636829). TG acknowledges the NASA Postdoctoral Program at the Goddard Space Flight Center, administered by Oak Ridge Associated Universities. 18 124 References 125 Bernstein M.P., S.A. Sandford, L.J. Allamandola, S. Chang, M.A. Scharberg (1995) "Organic 126 compounds produced by photolysis of realistic interstellar and cometary ice analogs 127 containing methanol." ApJ 454:327-344 128 Buszewski B., Noga S. (2012). "Hydrophilic interaction liquid chromatography (HILIC) – a 129 powerful separation technique." Analytical and Bioanalytical Chemistry 402: 231-247. 130 Cable, M. L., S. M. Hörst, C. He, A. M. Stockton, M. F. Mora, M. A. Tolbert, M. A. Smith 131 and P. A. Willis (2014). "Identification of primary amines in Titan tholins using microchip 132 nonaqueous capillary electrophoresis." Earth and Planetary Science Letters 403(0): 99-107. 133 Cable, M. L., S. M. Hörst, R. Hodyss, P. M. Beauchamp, M. A. Smith and P. A. Willis 134 (2012). "Titan Tholins: Simulating Titan Organic Chemistry in the Cassini-Huygens Era." 135 Chemical Reviews 112(3): 1882-1909. 136 Carrasco, N. , I. Schmitz-Afonso, J.-Y. Bonnet, E. Quirico, R. Thissen, O. Dutuit, A. Bagag, 137 O. Laprevote, A. Buch, A. Giuliani, G. Adande, F. Ouni, E. Hadamcik, C. Szopa and G. 138 Cernogora (2009). "Chemical characterization of Titan's tholins: Solubility, Morphology and 139 Molecular Structure Revisited." J. Phys. Chem. A 113 (42): 11195-11203. 140 Carrasco, N., T. Gautier, E.-t. Es-sebbar, P. Pernot and G. Cernogora (2012). "Volatile 141 products controlling Titan's tholins production." Icarus 219(1): 230-240. 142 Cleaves, H. J., C. Neish, M. P. Callahan, E. Parker, F. M. Fernández and J. P. Dworkin 143 (2014). "Amino acids generated from hydrated Titan tholins: Comparison with Miller–Urey 144 electric discharge products." Icarus 237(0): 182-189. 19 145 Coll, P., R. Navarro-González, C. Szopa, O. Poch, S.I. Ramírez, D. Coscia, F. Raulin, M. 146 Cabane, A. Buch, G. Israël (2013). "Can laboratory tholins mimic the chemistry producing 147 Titan's aerosols? A review in light of ACP experimental results" Planetary and Space Science 148 117: 91-103. 149 Danger, G., F.-R. Orthous-Daunay, P. de Marcellus, P. Modica, V. Vuitton, F. Duvernay, L. 150 Flandinet, L. Le Sergeant d'Hendencourt, R. Thissen, T. Chiavassa (2013). "Characterization 151 of laboratory analogs of interstellar/cometary organic residues using very high resolution 152 mass spectrometry" Geochimica et Cosmochimica Acta 118:184-201. 153 Derenne, S., C. Coelho, C. Anquetil, C. Szopa, A. S. Rahman, P. F. McMillan, F. Corà, C. J. 154 Pickard, E. Quirico and C. Bonhomme (2012). "New insights into the structure and chemistry 155 of Titan's tholins via 13C and 15N solid state nuclear magnetic resonance spectroscopy." 156 Icarus 221(2): 844-853. 157 Gautier, T., N. Carrasco, I. Schmitz-Afonso, D. Touboul, C. Szopa, A. Buch and P. Pernot 158 (2014). "Nitrogen incorporation in Titan's tholins inferred by high resolution orbitrap mass 159 spectrometry and gas chromatography–mass spectrometry." Earth and Planetary Science 160 Letters 404: 33-42. 161 He, C., G. Lin and M. A. Smith (2012). "NMR identification of hexamethylenetetramine and 162 its precursor in Titan tholins: Implications for Titan prebiotic chemistry." Icarus 220(2): 627- 163 634. 164 He, C. and M. A. Smith (2013). "Identification of nitrogenous organic species in Titan 165 aerosols analogs: Nitrogen fixation routes in early atmospheres." Icarus 226(1): 33-40. 20 166 He, C. and M. A. Smith (2014). "Identification of nitrogenous organic species in Titan 167 aerosols analogs: Implication for prebiotic chemistry on Titan and early Earth." Icarus 238(0): 168 86-92. 169 Jagota, S., J. Kawai, D. Deamer, C. McKay, B. Khare and D. Beeler (2014). "Surface-active 170 substances in a laboratory simulated Titan׳s organic haze: Prebiotic microstructures." 171 Planetary and Space Science 103(0): 167-173. 172 Khare, B. N., C. Sagan, E. T. Arakawa, F. Suits, T. A. Callcott and M. W. Williams (1984). 173 "Optical constants of organic tholins produced in a simulated Titanian atmosphere: From soft 174 x-ray to microwave frequencies." Icarus 60(1): 127-137. 175 McGuigan, M., J. H. Waite, H. Imanaka and R. D. Sacks (2006). "Analysis of Titan tholin 176 pyrolysis products by comprehensive two-dimensional gas chromatography–time-of-flight 177 mass spectrometry." Journal of Chromatography A 1132(1–2): 280-288. 178 Merline, D. J., S. Vukusic and A. A. Abdala (2013). "Melamine formaldehyde: curing studies 179 and reaction mechanism." Polym J 45(4): 413-419. 180 Muñoz Caro, G.M. and A. Schutte (2003). "UV-photoprocessing of interstellar ice analogs: 181 New infrared spectroscopic results." Astronomy and Astrophysics 412(1):121 182 Pernot, P., N. Carrasco, R. Thissen and I. Schmitz-Afonso (2010). "Tholinomics-Chemical 183 Analysis of Nitrogen-Rich Polymers." Analytical Chemistry 82(4): 1371-1380. 184 Philbrook, A., C. J. Blake, N. Dunlop, C. J. Easton, M. A. Keniry and J. S. Simpson (2005). 185 "Demonstration of co-polymerization in melamine–urea–formaldehyde reactions using 15N 186 NMR correlation spectroscopy." Polymer 46(7): 2153-2156. 21 187 Pilling, S., D. P. P. Andrade, Á. C. Neto, R. Rittner and A. Naves de Brito (2009). "DNA 188 Nucleobase Synthesis at Titan Atmosphere Analog by Soft X-rays†." The Journal of Physical 189 Chemistry A 113(42): 11161-11166. 190 Quirico, E., G. Montagnac, V. Lees, P. F. McMillan, C. Szopa, G. Cernogora, J.-N. Rouzaud, 191 P. Simon, J.-M. Bernard, P. Coll, N. Fray, R. D. Minard, F. Raulin, B. Reynard and B. 192 Schmitt (2008). "New experimental constraints on the composition and structure of tholins." 193 Icarus 198(1): 218-231. 194 Ruiz-Bermejo, M., C. Menor-Salván, E. Mateo-Martí, S. Osuna-Esteban, J. Á. Martín-Gago 195 and S. Veintemillas-Verdaguer (2008). "CH4/N2/H2 spark hydrophilic tholins: A systematic 196 approach to the characterization of tholins." Icarus 198(1): 232-241. 197 Sittler Jr, E. C., A. Ali, J. F. Cooper, R. E. Hartle, R. E. Johnson, A. J. Coates and D. T. 198 Young (2009). "Heavy ion formation in Titan's ionosphere: Magnetospheric introduction of 199 free oxygen and a source of Titan's aerosols?" Planetary and Space Science 57(13): 1547- 200 1557. 201 Somogyi, A., C.-H. Oh, M. A. Smith and J. I. Lunine (2005). "Organic Environments on 202 Saturn's Moon, Titan: Simulating Chemical Reactions and Analyzing Products by FT-ICR 203 and Ion-Trap Mass Spectrometry." Journal of the American Society for Mass Spectrometry 204 16(6): 850-859. 205 Somogyi, Á., M. A. Smith, V. Vuitton, R. Thissen and I. Komáromi (2012). "Chemical 206 ionization in the atmosphere? A model study on negatively charged "exotic" ions generated 207 from Titan's tholins by ultrahigh resolution MS and MS/MS." International Journal of Mass 208 Spectrometry 316–318(0): 157-163. 22 209 Szopa, C., G. Cernogora, L. Boufendi, J. J. Correia and P. Coll (2006). "PAMPRE: A dusty 210 plasma experiment for Titan's tholins production and study." Planetary and Space Science 211 54(4): 394-404. 212 Theulé, P., F. Duvernay, G. Danger, F. Borget, J. B. Bossa, V. Vinogradoff, F. Mispelaer and 213 T. Chiavassa (2013). "Thermal reactions in interstellar ice: A step towards molecular 214 complexity in the interstellar medium." Advances in Space Research 52(8): 1567-1579. 215 Vinogradoff, V., F. Duvernay, G. Danger, P. Theulé and T. Chiavassa (2011). "New insight 216 into the formation of hexamethylenetetramine (HMT) in interstellar and cometary ice 217 analogs." A&A 530: A128. 218 Vinogradoff, V., A. Rimola, F. Duvernay, G. Danger, P. Theule and T. Chiavassa (2012). 219 "The mechanism of hexamethylenetetramine (HMT) formation in the solid state at low 220 temperature." Physical Chemistry Chemical Physics 14(35): 12309-12320. 221 Wohnsiedler, H. P. (1953). "Polymerization in Melamine-Formaldehyde Molded Resins." 222 Industrial & Engineering Chemistry 45(10): 2307-2311. 223 Zheng, X.L., Yu B.S., Li K.X., Dai Y.N. (2012). "Determination of melamine in dairy 224 products by HILIC-UV with NH2 column." Food Control (23): 245:250. 23
1006.4020
1
1006
2010-06-21T10:04:56
Formation of Cosmic Crystals in Highly-Supersaturated Silicate Vapor Produced by Planetesimal Bow Shocks
[ "astro-ph.EP" ]
Several lines of evidence suggest that fine silicate crystals observed in primitive meteorite and interplanetary dust particles (IDPs) nucleated in a supersaturated silicate vapor followed by crystalline growth. We investigated evaporation of $\mu$m-sized silicate particles heated by a bow shock produced by a planetesimal orbiting in the gas in the early solar nebula and condensation of crystalline silicate from the vapor thus produced. Our numerical simulation of shock-wave heating showed that these {\mu}m-sized particles evaporated almost completely when the bow shock is strong enough to cause melting of chondrule precursor dust particles. We found that the silicate vapor cools very rapidly with expansion into the ambient unshocked nebular region; the cooling rate is estimated, for instance, to be as high as 2000 K s$^{-1}$ for a vapor heated by a bow shock associated with a planetesimal of radius 1 km. The rapid cooling of the vapor leads to nonequilibrium gas-phase condensation of dust at temperatures much lower than those expected from the equilibrium condensation. It was found that the condensation temperatures are lower by a few hundred K or more than the equilibrium temperatures. This explains the results of the recent experimental studies of condensation from a silicate vapor that condensation in such large supercooling reproduces morphologies similar to those of silicate crystals found in meteorites. Our results suggest strongly that the planetesimal bow shock is one of the plausible sites for formation of not only chondrules but also other cosmic crystals in the early solar system.
astro-ph.EP
astro-ph
Formation of Cosmic Crystals in Highly-Supersaturated Silicate Vapor Produced by Planetesimal Bow Shocks H. Miura1, K. K. Tanaka2, T. Yamamoto2, T. Nakamoto3, J. Yamada1, K. Tsukamoto1, and J. Nozawa1,4 [email protected] ABSTRACT Several lines of evidence suggest that fine silicate crystals observed in primi- tive meteorite and interplanetary dust particles (IDPs) nucleated in a supersatu- rated silicate vapor followed by crystalline growth. We investigated evaporation of µm-sized silicate particles heated by a bow shock produced by a planetesi- mal orbiting in the gas in the early solar nebula and condensation of crystalline silicate from the vapor thus produced. Our numerical simulation of shock-wave heating showed that these µm-sized particles evaporated almost completely when the bow shock is strong enough to cause melting of chondrule precursor dust par- ticles. We found that the silicate vapor cools very rapidly with expansion into the ambient unshocked nebular region; the cooling rate is estimated, for instance, to be as high as 2000 K s−1 for a vapor heated by a bow shock associated with a planetesimal of radius 1 km. The rapid cooling of the vapor leads to nonequi- librium gas-phase condensation of dust at temperatures much lower than those expected from the equilibrium condensation. It was found that the condensa- tion temperatures are lower by a few hundred K or more than the equilibrium temperatures. This explains the results of the recent experimental studies of condensation from a silicate vapor that condensation in such large supercooling reproduces morphologies similar to those of silicate crystals found in meteorites. Our results suggest strongly that the planetesimal bow shock is one of the plau- sible sites for formation of not only chondrules but also other cosmic crystals in the early solar system. 1Department of Earth and Planetary Materials Science, Graduate School of Science, Tohoku University, Aoba 6-3, Aramaki, Aoba-ku, Sendai 980-8578, Japan 2Institute of Low Temperature Science, Hokkaido University, Sapporo 060-0819, Japan 3Earth and Planetary Sciences, Tokyo Institute of Technology, Meguro, Tokyo 152-8551, Japan 4Institute for Materials Research, Tohoku University, 2-1-1 Katahira, Aoba-ku, Sendai 980-8577, Japan -- 2 -- Subject headings: interplanetary medium -- meteorites, meteors, meteoroids -- planet?disk interactions -- planets and satellites: formation -- shock waves 1. Introduction Vapor-solid (VS) growth is a major process for dust formation in the inner region of the early solar nebula, where the gas pressure was too low for a liquid phase to exist stably. Actually, there are several lines of evidence in interplanetary dust particles (IDPs) and primitive meteorites that the dust condensed directly from the vapor phase as suggested for example by enstatite whiskers elongated along the a-axis found in IDPs (Bradley et al. 1983) and µm-sized polyhedral olivine crystals with various morphologies found in matrix of primitive meteorite (Nozawa et al. 2009). Hereafter, we refer to these fine crystals as cosmic crystals. It is an important issue to reveal the formation environment of these cosmic crystals for understanding the early history of the solar system. To reproduce the cosmic crystals, evaporation and condensation experiments have been performed by many authors so far. Mysen & Kushiro (1988) carried out experiments in the system composed of Mg2SiO4-SiO2-H2 in the pressure ranging from 10−4 to 104 dyn cm−2 and in the temperature ranging from 1620 K to 1920 K to determine phase relations of the system. In these experiments, condensation of MgSiO3 and SiO2 phases took place from the Si-rich va- por, which was produced by incongruent vaporization of enstatite. Tsuchiyama et al. (1988) analyzed these condensates with the use of an analytical transmission electron microscope (ATEM) and a scanning electron microscope (SEM) to compare the condensates produced in the experiment with enstatite crystals found in IDPs (Bradley et al. 1983). They concluded that the characteristic textures of clinoenstatite found in IDPs were not reproduced in the evaporation and condensation experiments by Mysen & Kushiro (1988). Recently, Kobatake et al. (2008) carried out evaporation and condensation experiments to investigate a relationship between the growth conditions and morphologies of cosmic crys- tals condensed from highly supersaturated vapors. They used a sphere with forsteritic com- position (Mg2SiO4) as an evaporation source and succeeded in reproducing various morpholo- gies observed in µm-sized cosmic olivine crystals. They found that types of the morphology depend on the condensation temperature Tc. Namely, they found bulky-types at Tc > 1270 K, platy-types at 970 < Tc < 1270 K, and columnar-needle-types at 770 < Tc < 1090 K under the total pressure of 103 to 104 dyn cm−2. Furthermore, Yamada (2009) carried out the same experiments except that the evaporation source has enstatitic composition (MgSiO3) and suc- ceeded in reproducing enstatite whiskers elongated along the a-axis at 700 < Tc < 1150 K. They also reproduced enstatite crystals of platy-types at 1150 < Tc < 1300 K. Crys- -- 3 -- tals produced in their experiments have morphologies very similar to the cosmic crystals (Bradley et al. 1983; Nozawa et al. 2009). The condensation temperatures of these cosmic crystal analogs produced in the experiments aforementioned are substantially lower than the temperature of ∼ 1400 K, at which forsterite and enstatite condense in equilibrium from the solar nebula gas with the total pressure of 103 dyn cm−2 (Grossman 1972, e.g.,). The reproduction of the morphologies of cosmic crystals in the experiments suggests that the cosmic crystals were formed in highly supercooled vapors through nucleation and successive crystal growth; the condensation temperatures of the cosmic crystals are lower than those expected from the equilibrium theory by a few hundred degrees or more. To produce such supercooled silicate vapor, one would require significant evaporation of silicate dust followed by rapid cooling of its vapor. As a possible site of formation of the cosmic crystals, we notice a localized bow shock produced by a planetesimal revolving in a highly eccentric orbit in a gas of the solar nebula; the shocked region was originally proposed as a plausible site for chondrule formation (Hood 1998; Ciesla et al. 2004; Hood et al. 2005, 2009). When chondrule precursor dust aggregates enter the bow shock, they do not evaporate significantly because of their large sizes (∼ mm) but melt then cool and solidify to form chondrules. However, small dust particles of µm in size will evaporate completely behind the shock front if the shock is strong enough, and produce silicate vapor (Miura & Nakamoto 2005). The vapor cools rapidly with expansion in the vicinity of the planetesimal in orbit, resulting in a highly supercooled state. In this paper, we examine formation of cosmic crystals in the regions behind planetesimal bow shocks. The point of our discussion is whether the silicate vapor realizes the highly supercooling in which cosmic crystals of various morphologies are produced, or not. We give an overview of our model in §2. Section 3 describes the process of dust evaporation by shock-wave heating to evaluate the evaporation fraction of silicate dust particles in a wide range of the physical parameters. In §4, we investigate expansion of the silicate vapor behind the bow shock and estimate the cooling rate. Section 5 examines the properties of the condensates such as their particle sizes and morphologies expected from our model and compare the results with those obtained by the evaporation and condensation experiments. We discuss a comprehensive scenario on the formations of chondrules, cosmic crystals, and other materials in chondrites in §6. 2. Outline of the Model As stated in §1, a possible site for formation of fine cosmic crystals is a localized bow shock region associated with a planetesimal orbiting in an eccentric orbit. There arises -- 4 -- relative velocity between a planetesimal and the nebular gas both orbiting around the Sun, where the eccentricity of the planetesimal orbit is large. Weidenschilling et al. (1998) showed that the Jupiter mean motion resonances can excite planetesimal eccentricities up to e ≃ 0.3 or more. Nagasawa et al. (2005) analyzed the orbital evolution of terrestrial planetary embryos including the effect of the sweeping Jupiter secular resonance combined with tidal drag during dissipation of the protoplanetary gas disk. They found that the eccentricities of planetary embryos with mass of 0.01ME are excited up to e ≃ 0.6 or more at maximum and oscillate around the mean value of e ≃ 0.3 − 0.4 if Jupiter has the eccentricity of eJ = 0.05 (the current value is eJ = 0.0485), where ME is the Earth mass. The relative velocity between the eccentric planetesimal and the circularly orbiting nebular gas is estimated to be vp ≃ √e2 + i2 vK, where e and i are, respectively, eccentricity and inclination of a planetesimal orbit and vK is its Keplerian velocity. For i ≪ e, we obtain 3 AU(cid:17)−1/2 vp = 10.3(cid:16) e 0.6(cid:17)(cid:16) a km s−1, (1) where a is semi-major axis of a planetesimal orbit. The supersonic velocities relative to the nebula gas produce bow shocks in front of the planetesimals (Hood 1998; Ciesla et al. 2004). The relative velocity of vp = 10.3 km s−1 for e = 0.6 at a = 3 AU is large enough to cause melting of mm-sized silicate dust aggregates (Iida et al. 2001) and evaporation of µm-sized dust particles (Miura & Nakamoto 2005) in a region of the asteroid belt, where the gas density is n0 ∼ 3× 1013 cm−3 at the midplane of the gas disk in the minimum mass solar nebula model (Hayashi et al. 1985). An outline of the model for cosmic crystal formation is illustrated in Fig. 1. When the shock front formed by supersonic orbital motion of a planetesimal passes through a region in the nebula, the nebular gas is abruptly accelerated whereas µm-sized dust particles tend to keep their initial position because of their relatively large inertia. As a result, the dust particles find that they are exposed to a high-velocity gas flow suddenly and are heated to their evaporation temperature if the relative velocity and the gas density are large enough. Evaporation of the µm-sized dust particles in the post-shock region will be discussed in §3. The silicate vapor thus produced expands outward because its pressure is higher than that of the ambient unshocked region. The cooling associated with the expansion will produce supercooled silicate vapor. We discuss the cooling process of the silicate vapor and its cooling rate in §4. The present model supposes that the cosmic crystals observed in meteorites are condensation products in the expanding silicate vapor supercooled behind the planetesimal bow shock. Their sizes, morphologies, and condensation temperatures strongly depend on the density of silicate vapor and the cooling rate. We shall show that various kinds of cosmic crystals observed in meteorites are formed in the cooling of the silicate vapor produced by planetesimal bow shocks (§5). We shall point out that the present model -- 5 -- leads to simultaneous formation of chondrules and fine cosmic crystals (§6), and that their formation is an inevitable consequence of formation of planetary systems. 3. Evaporation of µm-Sized Dust Particles 3.1. Evaporation fraction We carry out numerical simulations of shock-wave heating by using a one-dimensional plane-parallel model developed by Miura & Nakamoto (2006). Actually, the structure of nebular gas around a planetesimal is not of one-dimensional plane-parallel structure. How- ever, the two-dimensional hydrodynamic simulation by Ciesla et al. (2004) showed that the one-dimensional plane-parallel approximation was valid in the vicinity of a planetesimal we are concerned with, say, a few times of planetesimal radius Rp. We set a computational do- main along the x-axis to be −Rp ≤ x ≤ Rp, where the x-axis is parallel to the gas flow and x = 0 at the shock front. In this region, the shocked gas structure can be regarded as one- dimensional plane-parallel (Hood et al. 2009). The simulations were carried out with varying the following input parameters: the planetesimal radius Rp, the pre-shock gas number den- sity n0, the dust-to-gas mass ratio ξ, and the shock velocity vs. In the simulations we set the ranges of parameters to be 1 ≤ Rp ≤ 1000 km, 1013 ≤ n0 ≤ 1015 cm−3, 0.01 ≤ ξ ≤ 0.1, and 5 ≤ vs ≤ 60 km, respectively. We take particle radius to be ad = 1 µm as a typical size of fine dust particles; a scaling to other sizes is easily done with the use of Eq. (3). The case of ξ = 0.1 is investigated to see the dependence of the evaporation fraction on ξ, although it will require significant settling or concentration of dust particles. Figure 2 shows the result for Rp = 100 km, n0 = 1015 cm−3, vs = 8 km s−1, and ξ = 0.01. Panel (a) shows temperature profiles of the gas (Tg, solid line), the dust particles (Td, dashed), and the radiation field (Trad, dotted) in the vicinity of the shock front (−1.0 ≤ x ≤ 1.0 km). Dust temperature increases rapidly just behind the shock front by gas frictional heating due to the velocity difference between gas and dust. In this stage, which we call the first stage, the dust temperature is determined by a balance among frictional heating, radiative cooling, and interaction with the ambient radiation field. The first stage ceases in a short period of time (less than 0.1 s in this case) because dust particles come to stop relative to the ambient gas. Panel (b) shows density profiles of dust (ρd, dashed) and silicate vapor (ρv, solid), which is produced by evaporation of the dust. It is found that the dust density ρd increases behind the shock front (0 . x . 0.3 km) because of deceleration by the gas friction. On the other hand, the vapor density ρv remains much smaller than ρd, indicating that evaporation of the dust particles during the first stage is negligible because of the very short duration of the first stage. -- 6 -- Panel (c) shows temperature profiles over a wide region around the planetesimal. The relative velocity between the gas and the dust particles is almost zero in almost all the region shown here (x & 0.5 km), so the frictional heating does not work. However, the dust temperature is kept above 1500 K because of efficient collisional heating by the ambient hot gas; we call this stage of collisional heating the second stage. During the second stage, dust particles continue to evaporate gradually as is seen from the density profiles shown in panel (d). One should note that evaporation of dust occurs mainly in this stage. At the edge of the calculation zone (x = 100 km), the vapor density is ρv = 1.76 × 10−10 g cm−3, while the density of survived solid dust particles is ρd = 1.73× 10−11 g cm−3. Therefore, the evaporation fraction η defined by η = ρv ρd + ρv (2) equals 0.91 in the case shown in Fig. 2, implying that 90 % of the dust mass evaporate during the second stage. Figure 3 summarizes the evaporation fraction η for various sets of values of the input parameters. Here, η is shown as a function of dust temperature at the second stage. We take the dust temperature Td2 at the time when vrel/vT = 0.1, where vT is root mean square of thermal velocities of the gas molecules and vrel is the velocity of dust particles relative to the gas; the result changes little even if we take vrel/vT = 0.05. Open circles in panel (a) show numerical results for all sets of the input parameters. Figure 3 indicates that η increases rapidly with increasing Td2. The temperature dependence of η is given by η = 1 −(cid:18)1 − ∆a ad (cid:19)3 , (3) where ∆a is size decrease of a dust particle after finishing substantial evaporation and is given by ∆a = jevap(Td2) ρc Td2 H Td2 −(dT /dt)Td2 for a spherical dust particle (see Appendix A for the derivation). Here, ρc is material density of the dust particle, jevap is the evaporation rate, i.e. mass of vapor evaporating from unit surface area of the particle per unit time, and H is latent heat of evaporation divided by the gas constant. Note that ∆a is independent of the original size ad and η is small for a large dust particle. The factor ∆tevap = Td2 H ∆t indicates an effective duration of evaporation during the cooling timescale defined by ∆t = Td2 −(dT /dt)Td2 (4) (5) (6) -- 7 -- at T = Td2. Note that the numerical results are reproduced well by Eqs. (3) and (4). This implies that the evaporation fraction η mainly determined by dust temperature Td2 in the second stage in spite that there are many other factors (Rp, n0, vs, and ξ) that may affect evaporation of the dust particles behind planetesimal bow shock. The duration of substantial evaporation ∆tevap is proportional to the cooling timescale ∆t of the hot gas. The expression of ∆t is very complex in general because it depends on var- ious physical processes such as vibrational/rotational transitions of CO and H2O molecules, thermal dissociation of H2 molecules, Lyman-α emission, and so forth. For a gas of the solar abundance, however, the major cooling process is Lyman-α emission for T & 104 K and thermal dissociation of H2 molecules for T & 3000 K (Miura & Nakamoto 2005, see Fig. 7). The timescale of cooling due to Lyman-α emission is shorter than ∼ 100 s. Below 3000 K, the gas cools within a timescale of ∼ 100 s due to vibrational/rotational transitions of CO and H2O molecules (Miura & Nakamoto 2005, see Fig. 7). The cooling timescale of the hot gas does not depend on the number density of the gas significantly. In the present case, the cooling timescale of ∆t ∼ 100 s reproduces the numerical results well as is seen from Fig. 3. Finally, let us examine the dependences of the evaporation fraction η on the parameters other than the temperature. Panels (b), (c), and (d) examine the dependence of the evap- oration fraction η on Rp, n0, and ξ, respectively. In panel (b), η for Rp = 1, 10, 100, and 1000 km are plotted by different symbols to see the dependence of η on Rp. There seems no clear systematic dependence of η on Rp even if we vary Rp by three orders of magnitude. Panel (c) examines the dependence on n0, the number density of pre-shock gas. There seems to be a slight trend that η decreases with increasing n0 but the dependence is unremarkable compared with the scatter of the data for each value of n0. Panel (d) examines the depen- dence on the gas-to-dust mass ratio ranging from ξ = 0.01 to 0.1 but we found no systematic trend of η on ξ, neither, within the plausible range of ξ. 3.2. Analytic estimation of the dust temperature We have shown that the evaporation fraction η is determined mainly by the dust tem- perature Td2 in the second stage. However, one needs to elaborate numerical simulations to calculate Td2. Instead, we derived an approximate analytic expression (B2) of Td2 in Appendix B by considering the energy balance of a dust particle in the second stage. The analytic formula of Td2 will also be useful for calculating the dust temperature and its evap- oration in a planetesimal bow shock in general. Figure 4 compares Td2 given by Eq. (B2) with that obtained from the numerical results. -- 8 -- It is found that both agrees with the difference less than ±50 K for Td2 . 1500 K. For Td2 & 1500 K, the numerical values of Td2 are systematically lower than those given by Eq. (B2). The reason of the deviation is that the analytic estimation ignores decrease in the optical depth due to dust evaporation in the shocked region. Actually, the decrease in the optical depth weakens the intensity of the ambient radiation field, which heat the dust. In consequence, the dust temperature decreases and its evaporation is suppressed. This negative feedback taken into account in the numerical simulation results in the numerical value of Td2 lower than that of the analytic estimation. The deviation at Td2 & 1500 K, however, does not influence the estimation of the evaporation fraction η much because η ≃ 1 in any case at these temperatures as seen from Fig. 3. Figure 5 shows the evaporation fraction η as a function of dust temperature Td2 as does Fig. 3 but Td2 in the horizontal axis is replaced by the one calculated by using Eq. (B2). Although the scatter of the data plotted is larger than in Fig. 3, we see that the analytic formulae still reproduce the evaporation fraction η. 4. Expansion and Cooling of the Shocked Gas 4.1. Equation of expansion When the hot gas in the shocked region cools down to the temperatures lower than ∼ 1500 K, dust particles re-condense from the vapor produced by evaporation of the original dust. In this subsection, we consider hydrodynamics and cooling of the expanding gas cloud to characterize the environment for formation of the cosmic crystals. Let us assume cylindrical expansion with initial radius R0 (see Fig. 1). Initial radius of the shocked region R0 is on the same order of magnitude as planetesimal radius Rp (Ciesla et al. 2004). Neglecting the expansion along the x-axis, the expansion velocity vr is described by dvr dt 1 ρ ∂p ∂r , = − (7) where ρ is the gas density and p is the gas pressure. We use a one-zone approximation and approximate vr and ∂p/∂r as vr ∼ ∂p ∂r ∼ p R , (8) where R is radius of the gas cloud at time t. We adopt a polytropic equation of state for the gas given by dR dt , − ρ0(cid:19)γ p = p0(cid:18) ρ , (9) -- 9 -- where ρ0 is initial gas density and γ > 1 is a parameter relating to the polytrope index. The conservation of mass during the expansion is expressed as Using Eqs. (7) to (10), we obtain the equation of expansion of the gas given by R2ρ = R2 0ρ0. d2 R dt2 = 1 γ R−2γ+1, with R = R/R0 and t = (R0/cs0)t, where cs0 =r γp0 ρ0 (10) (11) (12) is sound speed in the gas at t = 0. The dimensionless equations for expansion make it clear that the timescale of expansion of the shocked gas behind a planetesimal can be scaled by the sound-crossing time R0/cs0. Figure 6 shows the solutions of Eq. (11) for the initial conditions of R = 1 and vr = 0 (see Appendix C). It is clearly seen that the expansion is separated into two phases; the acceleration phase, in which vr increases with time but R remains almost at the initial radius R0, and the expansion phase, in which the shocked region begins to expand and vr almost equals a constant terminal velocity. The dashed curves in panel (b) show approximations of vr in the two limiting cases of t ≪ 1 and t → ∞ (see Appendix C) given by vr =  where ts0 ≡ R0/cs0. cs0 γ t ts0 cs0(cid:20) 1 γ(γ − 1)(cid:21)1/2 (cs0t ≪ R0), (cs0t ≫ R0), (13) 4.2. Cooling rate of the shocked gas Using the relation T = T0(ρ/ρ0)γ−1 = T0 R−2(γ−1) (14) given by Eqs. (9) and (10), and T ∝ p/ρ, we obtain the time variation of the gas temperature T as dT dt − dT dR = − vr = 2(cid:18)γ − 1 γ (cid:19)1/2 T0cs0 R0 R−2γ+1[1 − R−2(γ−1)]1/2 (15) -- 10 -- with the use of Eq. (C1) in Appendix C. One sees from Eq. (15) that the cooling rate −dT /dt as a function of R increases with increasing R at first, reaches a peak, and decreases in proportion to R−2γ+1. Figure 7 shows the cooling rate −dT /dt as a function of T which decreases monotonically with time. To evaluate the cooling rate, we need to specify a value of the initial temperature T0. The gas temperature just behind the shock front could be higher than 2000 K or more depending on the Mach number vr/cs0. However, even if the temperatures of the gas exceeds 2000 K, it cools rapidly by dissociation of hydrogen molecules and is kept around 2000 K owing to the energy balance between re-formation of hydrogen molecules by three-body reaction and their dissociation (Iida et al. 2001). We set T0 = 2000 K to estimate the cooling rate around the condensation temperatures. To consider condensation through nucleation, on the other hand, we should refer to the cooling rate −dT /dt when the vapor becomes supersaturated. Taking the equilibrium condensation temperatures of Te = 1444 K for forsterite and Te = 1349 K for enstatite for the total pressure of 103 dyn cm−2 (Grossman 1972) as a measure of estimating the condensation temperature, we have Te/T0 = 0.65− 0.75 and dT /dtTe = (0.25 − 0.35)T0/(R0/cs0) for γ = 7/5 and 5/3 (see Fig. 7). We set R0 to be planetesimal radius Rp in what follows. In consequence, the cooling rate is estimated to be: −(cid:18)dT dt(cid:19)Te ≃ (0.25 − 0.35) ≃ 2000(cid:18) Rp T0 Rp/cs0 1 km(cid:19)−1(cid:18) T0 2000 K(cid:19)(cid:16) cs0 3.7 km s−1(cid:17) K s−1. (16) It should be pointed out that cooling of the shocked gas given by Eq. (16) can be used so far as the pressure of the shocked gas p is much larger than the ambient gas pressure pamb. The shocked gas pressure before the expansion is p ∼ 100 pamb for the shock velocity of an H2 gas of 10 km s−1. The gas temperature at that time is ∼ 2000 K as a result of the balance between H2 dissociation and its re-formation (Iida et al. 2001; Miura & Nakamoto 2005). The pressure and temperature decrease by subsequent cylindrical expansion. When the temperature drops to the typical condensation temperature of ∼ 1000 K, the radius of the cylinder is 2.4 times the initial one for adiabatic expansion, and the shocked gas pressure also decreases to ∼ 1/10 of that before expansion. However, the gas pressure is still higher than pamb by an order of magnitude. Therefore, Eq. (16) is applicable throughout the expansion phase of interest including the time of condensation. We focus here the adiabatic expansion because the radiative losses are negligibly small for small shocks as is shown below. Main coolants of the nebula gas at 2000 K are vibra- tional emissions of CO and H2O molecules. The cooling timescale due to these vibrational emissions was estimated to be ∼ 100 sec, which does not significantly depend on the gas -- 11 -- density (Miura & Nakamoto 2005). On the other hand, the cooling timescale due to the adiabatic expansion behind a planetesimal is shorter than ∼ 100 sec for planetesimal radius of < 100 km (see Eq. (16)). Therefore, the shocked gas cools by the expansion before the vibrational emissions remove the thermal energy significantly. The radiative losses might work for large shocks (& 100 km) because the expansion takes longer time. However, a large optical depth for these emissions resulting from large shocks prevents the radiative losses from being efficient. 4.3. Possibility of chondrule formation In the formation of chondrules, their cooling rate during solidification is one of the key physical quantities. According to the planetesimal bow shock model, the cooling rate was estimated to be ∼ 103 K hr−1 for planetesimal radius Rp = 1000 km, > 104 K hr−1 for Rp = 100 km, and > 105 K hr−1 for Rp = 10 km (Hood et al. 2005). On the other hand, the cooling rate of the shocked gas calculated from Eq. (16) is 7×103, 7×104, and 7×105 K hr−1 for Rp = 1000, 100, and 10 km, respectively. Although Eq. (16) is not a cooling rate of a chondrule itself but of the shocked gas strictly speaking, we note that both estimations of the cooling rates are comparable; this is because the cooling of chondrules is regulated by that of the shocked gas (Iida et al. 2001). Therefore, Eq. (16) measures the cooling rate of chondrules. A widely accepted range of the cooling rate of chondrules at solidification is ∼ 10 − 1000 K hr−1 (Hewins et al. 2005, and references therein), which is much slower than that pre- dicted by Eq. (16). However, we consider that this disagreement does not necessarily exclude planetesimal bow shocks as a chondrule formation site. In fact, some authors assert rapid cooling rates, which are in the range estimated from Eq. (16). Yurimoto & Wasson (2002) proposed that rapid cooling (∼ 105 − 106 K hr−1) was necessary to account for the observed Fe-Mg and O-isotopic exchange in a CO-chondrite type-II chondrule. Wasson & Rubin (2003) proposed that very thin overgrowths on some relict grains in chondrules must have been formed by the rapid cooling. The crystallization experiments of a melt droplet by a levi- tation method succeeded in reproducing chondrule-solidification textures in the experimental conditions of the rapid cooling (Tsukamoto et al. 1999; Nagashima et al. 2006). Although the rapid cooling scenario does not seem to have been accepted widely to the meteoritic community (Hewins et al. 2005), there has been no definite evidence that rejects the rapid cooling scenario completely. Therefore, we consider that the planetesimal bow shock is still one of the possible models to be studied for chondrule formation. -- 12 -- 5. Formation of Cosmic Crystals 5.1. Cooling parameter Λ for homogeneous nucleation Cosmic crystals condense in the course of cooling of the vapor produced by a planetes- imal bow shock. When almost all dust particles evaporate by the bow shock, there is no solid surface available on which the supersaturated vapor condenses. In this case, cosmic crystals are formed through homogeneous (spontaneous) nucleation. In homogeneous nucle- ation, condensation does not begin when the cooling vapor becomes saturated but begins effectively after the vapor becomes supersaturated to a certain degree. Yamamoto & Hasegawa (1977) and Draine & Salpeter (1977) formulated a grain forma- tion process though homogeneous nucleation in a vapor and derived analytical expressions of a typical size of grains and their actual condensation temperature Tc in a supercooling state as functions of two dimensionless parameters. One is a cooling parameter defined by Λ = νcolltT H/Te − 1 , (17) where νcoll is collision frequency of vapor molecules in thermal motion, tT = Te/ (−dT /dt)Te is cooling timescale of a vapor at T = Te with Te being equilibrium condensation temperature, and H is latent heat of condensation divided by the gas constant and equivalent to that of evaporation (see Appendix A). Note that Te (> Tc) is a temperature at which a vapor and a bulk condensate co-exit in chemical equilibrium and approaches Tc as tT gets so long that the equilibrium between the vapor and the condensate is realized. Grain size ad is mainly determined by Λ and is roughly given by ad/a0 ∼ 0.1 Λ for Λ ≫ 1, where a0 is the radius of a vapor molecule (Yamamoto & Hasegawa 1977). In Eq. (17), νcoll is calculated from the vapor density ρv, and tT from the cooling rate of the vapor (see Appendix D). The other parameter is a dimensionless surface tension defined by Γ = 4πa2 0γs kBTe , (18) where γs is surface tension of a condensate and a0 = (3µcma/4πρc)1/3 (i.e. equivalent radius of a sphere whose volume equals the volume of a unit cell of the condensate) with µc being molecular weight of a unit cell of a condensate, ma = 1.66 × 10−24 g is atomic mass unit, and ρc its bulk density. A degree of supercooling ∆T = Te − Tc is mainly determined by the parameter Γ and is approximately related to Γ as ∆T ∝ Γ3/2 (Yamamoto & Hasegawa 1977). Figure 8 shows a relation between Λ and the evaporation fraction η. Each of the plots indicates η calculated in §3 for a given set of values of the parameters, while Λ is calculated -- 13 -- from Eq. (D2). All panels indicate the trend that Λ increases with η. This is simply because the larger degree of evaporation of pre-existing dust is, the larger amount of the vapor is produced, which in consequence provides favorable conditions for homogeneous condensation of cosmic crystals. Note that homogeneous condensation is possible only if Λ > 1; otherwise, the vapor is too tenuous for condensation to occur. Panel (a) shows the results of the calculations for all of the parameter sets, indicating that there appear many cases of Λ > 1 for η > 10−4. Even the cases of Λ as large as 105 are realized for complete evaporation (η ≃ 1) of pre-existing dust. The contribution to the vapor production comes mainly from µm-sized dust particles if their size distribution is steeper than a−2 d . The presence of many cases of Λ > 1 implies that condensation of cosmic crystals through homogeneous nucleation behind planetesimal bow shocks is possible for η > 10−4. We note that a variety in the Λ-values suggests formation of various kinds of cosmic crystals. The panel (b) displays the dependence of η and Λ on the planetesimal radius Rp. From panel (b), one sees that the homogeneous condensation occurs hardly except for η ∼ 1 for a bow shock produced by small planetesimals of Rp = 1 km but occurs almost always for a planetesimal of Rp = 1000 km even if the evaporation is not so significant (η & 10−4). 5.2. Size and morphology of cosmic crystals Figure 9 displays typical size a∞ of condensed particles and supercooling ∆T in terms of Λ and Γ. The supercooling ∆T in the vertical axis is normalized by the equilibrium condensa- tion temperature Te. Each solid curve shows the relation between a∞ and ∆T for a constant value of Γ, and dashed lines combine points for the same value in Λ (Yamamoto & Hasegawa 1977). The grayed region indicates a parameter range expected from the planetesimal bow shock. The possible range of Λ was discussed in §5.1. The values of Γ, on the other hand, have uncertainties because of short of the experimental data for the surface tension γs of forsterite and enstatite. For forsterite, γs is measured to be 1280 erg cm−2 in vacuum for a {010} surface and larger values for other ones (de Leeuw et al. 2000), which corresponds to Γ ≃ 30 or more. For enstatite, there are no reliable data of surface tension. We assume the similar value as that of forsterite. In the calculations, we take 10 < Γ < 60 for safety. It should be noted that the sizes a∞ and the supercoolings ∆T revealed from the analyses of a variety of cosmic crystals are included in the region realized by planetesimal bow shocks. Let us discuss in more detail the formation conditions of each of the cosmic crystals shown in Fig. 9. -- 14 -- 5.2.1. Enstatite whisker and platelet The experiment by Yamada (2009) showed that formation of enstatite whiskers elon- gated toward the a-axis required the degree of supercooling of 0.15 < ∆T /Te < 0.48. They also reproduced platy-type enstatite crystals at 0.04 < ∆T /Te < 0.15. It is interesting to note that the whisker has larger Γ than the platy-type, although precise values of their sur- face tension are unknown. Typical size of the enstatite crystals is ∼ 0.1 - 1 µm, which size is similar to that of natural samples found in IDPs (Bradley et al. 1983). A set of these con- ditions is shown by the red region in Fig. 9, indicating that enstatite whiskers and platelets can be formed by planetesimal bow shocks of 103 . Λ . 104. This range of Λ is realized if the bow shocks are produced by planetesimals of intermediate size (Rp ∼ 100 km) and lead to almost complete evaporation of the original dust (η ≃ 1). If the amount of the silicate vapor is small leaving a large amount of dust particles that survived evaporation (η ≪ 1), on the other hand, the vapor will condense onto the dust surface. This case yields other types of thermally-processed particles observed in chondritic meteorites (see §6). 5.2.2. Olivine crystals with various morphologies Kobatake et al. (2008) examined supercooling ∆T required for formation of olivine crys- tals by a laboratory experiment. They showed that bulky-type olivine crystals were repro- duced at ∆T /Te . 0.12, the platy-type at 0.12 < ∆T /Te < 0.33, and the columnar-needle- type at 0.24 < ∆T /Te < 0.47. As was so for enstatite, the needle-type has larger Γ than the platy-type; the bulky type has the smallest Γ. Typical size of the condensates is a∞ ∼ µm, which size is close to those of the natural samples found in the matrix of Allende meteorite (Nozawa et al. 2009). The green region shows the supercooling ∆T /Te and the sizes a∞ for these fine olivine crystals, indicating that these particles can be formed by planetesimal bow shock of 104 . Λ . 105. This condition is realized by the bow shocks produced by relatively large planetesimals (Rp ∼ 1000 km) associated with almost complete evaporation of the original dust particles. 5.2.3. Ultra-fine particles Toriumi (1989) observed fine particles in the matrix of Allende meteorite using a SEM and a TEM and measured their sizes ad. The observed size distribution could be reproduced by a log-normal one for 1 < ad < 10 nm with its peak at a = 5 nm and by a power law for ad > 10 nm. We display the size range of the ultra-fine particles by the blue region in Fig. 9. -- 15 -- The size range suggests that Λ ≃ 10− 100 is a plausible condition for formation of ultra-fine particles. This is in agreement with the conclusion of Toriumi (1989) that ultra-fine particles seem to have been formed by condensation from a vapor far from equilibrium in the early solar nebula. The present model implies that ultra-fine particles were formed by bow shocks produced by much smaller planetesimals (Rp ∼ 1 − 10 km) than those producing µm-sized cosmic crystals, associated with almost complete evaporation. The formation condition of Λ ≃ 10 − 100 also realizes for large planetesimals (Rp & 100 km) and small evaporation fraction (η ∼ 10−4 − 10−3), however, in this case the ultra-fine particles generated from the vapor are very rare because of the tiny evaporation fraction. 5.3. Heterogeneous condensation for incomplete evaporation We discussed formation of cosmic crystals through homogeneous nucleation in §5.1 and §5.2 assuming that almost all dust particles evaporate by a planetesimal bow shock. There is an opposite case that a substantial fraction of the dust particles survives against evaporation and acts as seed nuclei and that condensation occurs through nucleation on their surfaces (heterogeneous condensation). Which type of condensation actually occurs depends on the total surface area of dust particles available for heterogeneous nucleation. We shall show below that both types of condensations can occur depending on the radii of planetesimals generating bow shocks and on the evaporation fraction. In homogeneous nucleation, condensation does not begin when the cooling vapor be- come saturated but begins effectively after the vapor becomes supersaturated to a certain degree. Namely, there arises some induction time tind after the vapor becomes saturated (Yamamoto & Hasegawa 1977). The induction time is related to the cooling timescale tT = Te/(−dT /dt)Te as tind ≃ xJ H/Te − 1 tT ∼ (0.08 − 4.0)(cid:18) Rp 1 km(cid:19) s, (19) where xJ = 2 − 70 for situations we consider in this paper (Λ = 1 − 105 and Γ = 10 − 60 as explained in §5.1 and §5.2). The time intervals required for nucleation and growth is about ten times shorter than the induction time (Yamamoto & Hasegawa 1977). Therefore, tind represents a typical timescale for dust formation through homogeneous nucleation after the vapor becomes saturated. In heterogeneous nucleation, on the other hand, we estimate its timescale by using the adhesion timescale, which provides an underestimate of the timescale of heterogeneous con- densation because it ignores the induction time for heterogeneous nucleation. The adhesion -- 16 -- timescale tad is the one during which most of the vapor molecules sticks onto the surface of dust particles. For silicate condensation, we regard SiO molecule as a key species that controls the rate of condensation (see also Appendix D). Denoting the radius of the dust particles by ad, the adhesion timescale is estimated to be: tad = adρc 3αsρd (cid:18)2πµSiOma kBTe (cid:19)1/2 ≃ 50 α−1 s (1 − η)−1(cid:18) ad µm(cid:19)(cid:18)10−10 g cm−3 ρd + ρv (cid:19) s, (20) where ρd = (1 − η)(ρd + ρv) is density of dust particles surviving in the post-shock region against evaporation, αs is sticking probability of vapor molecules onto the dust surface, and µSiO = 44 is molecular weight of SiO. One should note that, in Eq. (20), the factor 3ρd/ρcad indicates total surface area of the dust particles per unit volume and (8kBTe/πµSiOma)1/2 is mean thermal velocity of SiO molecules. Homogeneous nucleation takes place if tind < tad. This condition is satisfied when the planetesimal radius Rp is relatively small (Rp . 10− 500 km), or there are few survived dust particles because of significant evaporation (η ∼ 1). In this case, cosmic crystals condense directly from the vapor. In contrast, heterogeneous condensation becomes effective if the planetesimal radius Rp is large (Rp & 10 − 500 km) and a substantial fraction of dust particles survives against evaporation (η ≪ 1). We shall discuss generic relations between cosmic crystals and chondrules in § 6 in detail. 6. Summary and Discussion Chondritic meteorites are composed of materials that have been experienced thermal processing of various degrees in the early solar nebula. These materials include chondrules, fine-grained rims on chondrules and interchondrule matrix (Alexander 1995), and cosmic crystals discussed in the previous section. In this section, we discuss how the planetesimal bow shock scenario explains the formations of these chondritic materials. A planetesimal bow shock was originally proposed as a possible site for chondrule forma- tion (Hood 1998). Iida et al. (2001) showed that millimeter-sized dust aggregates (chondrule precursors) are heated and melt behind a shock front if the shock velocity and the pre-shock gas density are in an appropriate range. Complete evaporation hardly occurs for chondrule precursors because of their large size (see Eq. (3)). Their contribution to the vapor pro- duction is negligibly small compared with that of µm-sized dust particles for the dust size distribution steeper than a−2 d . Large molten chondrule precursor dust survives against evap- oration, cools and solidifies to form chondrules. In contrast, µm-sized particles evaporate significantly in the hot gas behind the planetesimal bow shock and produce silicate vapor. -- 17 -- The silicate vapor cools rapidly behind the bow shock and becomes supersaturated, leading to condensation to produce various kinds of materials observed in chondritic meteorites and IDPs. The condensed materials exhibit a wide variety in morphologies and sizes depending on their formation conditions such as the cooling rate and the evaporation fraction of µm-sized dust particles. The cooling rate is inversely proportional to the size of a planetesimal that produces a bow shock (see Eq. (16)), thus decreases with time on average, namely, with growth of planetesimals. The evaporation fraction η changes by many orders of magnitude in the range of the shock conditions realized in early solar nebula (Iida et al. 2001). Figure 10 summarizes condensation products in the course of the planetesimal growth. At the early stage of 1 . Rp . 10 km, the vapor produced by small planetesimals cools so rapidly that the cooling parameter is Λ . 103, which is realized for 10−2 . η . 1 (see Fig. 8b). Condensation of the vapor through homogeneous nucleation for 1 < Λ . 103 leads to formation of nm-sized ultra-fine particles as observed in the matrix (see from Fig. 9). Furthermore, the results of §5.3 indicate that the rapid cooling prevents heterogeneous con- densation on survived dust particles because of rapid consumption of the vapor by homoge- neous condensation to form ultra-fine particles. To summarize, most of the vapor condensed to the ultra-fine particles at the early stage of planetesimal growth. When planetesimals grow up to a several 100 km or more, condensation occurs through both homogeneous and heterogeneous nucleations. If almost all of the small dust particles evaporate (η ∼ 1) behind the bow shock, µm-sized euhedral silicate crystals condense through homogeneous nucle- ation. The cooling parameter is 103 . Λ . 105 for η ∼ 1 (see Fig. 8b) for planetesimals of 100 . Rp . 1000 km. This situation leads to condensation of enstatite whisker elon- gated to a-axis as found in IDPs (Bradley et al. 1983) and polyhedral olivine crystals as found in the matrix of Allende meteorite (Nozawa et al. 2009) (see Fig. 9). Bare chondrules without fine-grained rims could also be formed in this case. On the other hand, if many of the dust particles survive against evaporation (η ≪ 1) and suffer partial evaporation, the vapor condenses heterogeneously onto the survived dust particles, resulting in the formation of other kinds of meteoritic materials. The vapor condensed heterogeneously on chondrules already solidified could form fine-grained rim on their surfaces. The survived µm-sized dust particles would also be covered with materials condensed from vapor, and would accumulate as fine-grained interchondrule matrix in chondritic meteorites after that. It is worth noting that the partial evaporation of dust particles would lead to elemental fractionation. The fractionated vapor rich in volatile elements re-condensed within a short period of time (see Eq. (20)) on the survived dust particles. A fine-grained rim of a chondrule and a fine-grained interchondrule matrix thus produced would have elemental composition complementary to that of the chondrule. The composition of the whole particle should be the same as that of -- 18 -- the original dust particles before evaporation according to the present model. This is consis- tent with the genetic relationship among chondrules, interchondrule matrix, and fine-grained rims that these components either formed from a common source material, are products of the same process, or have exchanged materials during formation (Huss et al. 2005). In summary, the planetesimal bow shock model can provide a comprehensive scenario for the formation of various cosmic crystals and other materials observed in chondritic mete- orites. Because the heating events happened in a localized region of the shocked gas within a short period of time, one may expect that a series of the thermal processing, heating, evapo- ration, and condensation completed in a closed-system. The scenario is in harmony with the genetic relationship suggested by the analyses of chondritic meteorites and IDPs that these are produced in the course of the processing from a common source material together with exchanges of the materials during their formation. We are grateful to Dr. M. Nagasawa for useful discussion on the orbital evolutions of planetesimals. We acknowledge helpful comments of an anonymous referee. This study was supported partly by the Grant for the Joint Research Program of the Institute of Low Tem- perature Science, Hokkaido University. H.M. was supported by Tohoku University Global COE Program "Global Education and Research Center for Earth and Planetary Dynamics," by the "Program Research" in Center for Interdisciplinary Research, Tohoku University, and by the Grant-in-Aid for Scientific Research from JSPS (19204052). T.Y. acknowledges support by the Grant-in-Aid for Scientific Research from JSPS (21244011). A. Size Decrease of a Particle by Evaporation and the Evaporation Fraction Let us consider evaporation of a spherical dust particle of initial radius ad. The evapo- ration fraction η is given by η = 3 4πa3 d Z ad ad−∆a 4πa2da = 1 −(cid:18)1 − ∆a ad (cid:19)3 , (A1) where ∆a is decrease in radius due to evaporation. Equation (A1) indicates clearly that η depends only on ∆a/ad, the ratio of the size decrease to the initial size. Since the size decrease ∆a due to evaporation is independent of the particle radius ad except through a slight dependence of the dust temperature on ad, the increase in the particle radius ad simply causes the decrease in η according to Eq. (A1). For example, even if µm-sized dust particles evaporate almost completely (η = 0.999), chondrule-sized particles (ad = 500 µm) evaporate by only a small fraction of η ≃ 5 × 10−3. We carry out the calculations for ad = 1 µm, but one can evaluate η for other ad by using Eq. (A1). -- 19 -- With the use of the evaporation rate jevap, the size decrease ∆a by evaporation during cooling from temperature Ti to Tf is expressed by ∆a = 1 ρc Z jevapdt = 1 ρc Z Tf Ti jevap(T ) dT /dt dT. The evaporation rate as a function of temperature T behaves as jevap(T ) = const · T β exp(−H/T ), (A2) (A3) according to the Hertz-Knudsen equation (Nagahara et al. 1996; Miura et al. 2002, see), where β ∼ −1/2 is a constant. Here, H = Levap/Rgas is latent heat of evaporation in units of temperature, where Levap is that in units J mol−1 and Rgas is the gas constant in units of J K−1 mol−1. In the present case, we are concerned with evaporation of forsterite (Mg2SiO4), for which Levap = 1.58×1013 J mol−1. This leads H = Levap/6Rgas = 3.17×104 K, where the factor of 6 results from the stoichiometric coefficients of the chemical reactions at evaporation (Miura et al. 2002, see Eq. (38)). According to the measurement of evaporation rate of forsterite, jevap depends also on the partial pressure pH2 of ambient hydrogen molecule (Tsuchiyama et al. 1998), but we may take the pressure at T = Ti in Eq. (A2) because ∆a is determined by the physical conditions at T = Ti as will be seen below. Integration on the RHS of Eq. (A2) can be performed by noting that e−H/T is a rapidly varying function compared to the remaining function in the integrand. Integrating by part and remaining the term of order Ti/H ≪ 1, one obtains ∆a = jevap(Ti) ρc Ti H ∆t, where the contribution from the upper limit of the integral is negligible. Here ∆t = Ti (−dT /dt)Ti (A4) (A5) is cooling timescale of dust particles at T = Ti, for which we take dust temperature Td2 in the second stage. B. Dust Temperature behind Shock Front We give here an analytic expression that gives in good approximation of the dust tem- perature in the post-shock region after the relative velocity between the gas and dust particles is almost damped (the second stage). The dust temperature at this stage, Td2, is determined -- 20 -- by the energy balance between collisional heating by the ambient hot gas and the radiative cooling: 1 4 γ + 1 γ − 1(cid:18)γkBT ′ πµma(cid:19)1/2 n′kB (T ′ − Td2) + σSB(T 4 rad − T 4 d2) = 0, (B1) where T ′ is post-shock gas temperature, n′ is post-shock gas number density, Trad is ambi- ent radiation temperature, µ is mean molecular weight of the gas, and σSB is the Stafan- Boltzmann constant. Here, we approximated the emission and absorption coefficients to be unity (Miyake & Nakagawa 1993). Since Trad ∼ Td2 in the second stage as seen from Fig. 2(c), one obtains Td2 = γr(γkBT ′/πµma)1/2n′kBT ′/4 + 4σSBT 4 rad γr(γkBT ′/πµma)1/2n′kB/4 + 4σSBT 3 rad , (B2) from Eq. (B1) by using the approximation that (T 4 γr ≡ (γ + 1)/(γ − 1). rad − T 4 d2) ≃ 4T 3 rad(Trad − Td2), where In Eq. (B2), the post-shock gas number density n′ is given from the Rankine-Hugoniot relation and the almost isobaric condition for the post-shock gas in the one-dimensional plane-parallel geometry (Susa et al. 1998; Miura et al. 2002) by n′ ≃ 2 γ + 1 ρ0v2 s kBT ′ , (B3) where ρ0 = µman0 is the gas density in the pre-shock region and vs is the shock velocity. The post-shock gas temperature T ′ in Eq. (B2) may be obtained by using the Rankine- Hugoniot relation. However, we have to pay attention that, at high temperatures of T ′ & 2000 K, the gas cools very rapidly due to the dissociation of hydrogen molecules (Iida et al. 2001). Therefore, we set T ′ as T ′ = min(cid:20)2(γ − 1) (γ + 1)2 µmav2 s kB , 2000 K(cid:21) . (B4) The radiation temperature at the shock front is given by taking the blanket effect into account (Miura & Nakamoto 2006) as Trad =(cid:18) 2 + 3τpre 4σSB f 2 ρ0v3 s(cid:19)1/4 , (B5) where τpre is optical depth of the pre-shock region and f is the fraction of the gas energy flux that returns upstream in the form of radiation. We set f = 0.5 for simplicity. The optical depth τpre is estimated to be τpre = Ls, (B6) 3ξρ0 4adρc -- 21 -- where ξ is dust-to-gas mass ratio in the pre-shock region and Ls is dimension of the pre-shock region, in which the dust particles contribute to the blanket effect around the shock front1. We set Ls = Rp in this study (see §3.1). Equation (B2) together with Eqs. (B3), (B4), and (B5) is an analytic expression of the dust temperature in the second stage. It should be noted that these equations include all of the input parameters Ls = Rp, n0 = ρ0/µma, ξ, and vs. C. Solutions of the Equations of Expansion Integrating Eq. (11) from t = 0 to t after multiplying d R/dt = vr on both sides, we obtain the expansion velocity to be: vr = d R dt γ(γ − 1) #1/2 ="1 − R−2(γ−1) . (C1) for the initial conditions of vr = 0 and R = 1. Equation (C1) is integrated further to yield the radius R as a function of time t as t pγ(γ − 1) R =Z 1 dy p1 − y−2(γ−1) . (C2) The right-hand side of Eq. (C2) may be expressed by the hypergeometric function but numerical integration is more practical to get the results, which are shown in Fig. 6(b) by solid curves. In the limits of t ≪ 1 and t → ∞, the velocity is approximated to be: t γ (cid:20) 1 γ(γ − 1)(cid:21)1/2 (t ≪ 1), (t → ∞). (C3) vr =  Both approximations are shown in Fig. 6(b) by the dashed lines. The time at the intersection t∗, at which the two limiting approximations cross each other, is given by t∗ =(cid:18) γ γ − 1(cid:19)1/2 . (C4) 1Ls corresponds to xm in Miura & Nakamoto (2006). -- 22 -- At the intersection, the radius and expansion velocity are given by R∗ = 2γ − 1 2(γ − 1) , vr∗ =(cid:20) 1 γ(γ − 1)(cid:21)1/2 . (C5) D. Evaluation of cooling parameter Λ for multi-component evaporation In this paper, we are concerned with evaporation and condensation of forsterite (Mg2SiO4), in which Mg and SiO should be considered as vapor species (Nagahara et al. 1996). For deal- ing with nucleation of a multi-component system, we adopt the key species approximation that the rates of nucleation and grain growth are controlled by one chemical species (key species) that has the least collision frequency among the major vapor species that condense into the grain (Kozasa & Hasegawa 1987). The conditions for the key-species approxima- tion to hold were examined by Yamamoto et al. (2001) in formulating theory of nucleation involving chemical reactions. In Eq. (17), the collision frequency of vapor molecules of mass µvma and number density nv is given by νcoll = πa2 molecular weight of the vapor molecules. Following Kozasa & Hasegawa (1987), we take SiO molecules as the key species of silicate condensation. This implies that nv = nSiO and µv = µSiO = 44. Using Eqs. (16) and (12), we obtain 0αsnvp8kBTe/πµvma, where αs is sticking probability, µv is mean Λ = 4πa2 0αsnSiORp H/Te − 1 (cid:18) µ µSiO(cid:19)1/2(cid:18) Te T0(cid:19)3/2(cid:18) 1 πγ(cid:19)1/2 × 1 (0.25 − 0.35) , (D1) where µ is mean molecular weight of the gas. The value of H/Te is estimated to be H/Te−1 ≃ 20 for H ≃ 3× 104 K for forsterite; enstatite yields the similar value. The value of a0 is given by a0 = 2.6 A. In consequence, Λ is evaluated roughly to be: Λ ≃ 400(cid:18) Rp 100 km(cid:19)(cid:18) ρv 10−10 g cm−3(cid:19) (D2) for αs = 1. REFERENCES Alexander, C. M. O'D. 1995, Geochim. Cosmochim. Acta, 59, 3247 Bradley, J. P., Brownlee, D. E., & Veblen, D. R. 1983, Nature, 301, 473 -- 23 -- Ciesla, F. J., Hood, L. L., Weidenschilling, S. J. 2004, Meteorit. Planet. Sci., 39, 1809 de Leeuw, N. H., Parker, S. C., Catlow, C. R. A., & Price G. D. 2000, Phys. Chem. Minerals, 27, 332 Draine, B. T., & Salpeter, E. E. 1977, J. Chem. Phys., 67, 2230 Grossman, L. 1972, Geochim. Cosmochim. Acta, 36, 597 Hayashi, C. K., Nakazawa, K., & Nakagawa, Y. 1985, in Formation of the solar system, ed. D. C. Black & M. S. Matthews, M.S. (Univ. of Arizona Press, Tucson), 1100 Hewins, R. H., Connolly, Jr., H. C., Lofgren, G. E., & Libourel, G. 2005, in Chondrites and the Protoplanetary Disk, ed. A. N. Krot, E. R. D. Scott, & B. Reipurth (San Francisco: Astronomical Society of the Pacific), 286 Hood, L. L. 1998, Meteorit. Planet. Sci., 33, 97 Hood, L. L., Ciesla, F. J., & Weidenschilling, S. J. 2005, in Chondrites and the Protoplane- tary Disk, ed. A. N. Krot, E. R. D. Scott, & B. Reipurth (San Francisco: Astronomical Society of the Pacific), 873 Hood, L. L., Ciesla, F. J., Artemieva, N. A., Marzari, F., & Weidenschilling, S. J. 2009, Meteorit. Planet. Sci., 44, 327 Huss, G. R., Alexander, C. M. O'D., Palme, H., Bland, P. A., & Wasson, J. T. 2005, in Chondrites and Protoplanetary Disk, ed. A. N. Krot, E. R. D. Scott, & B. Reipurth (San Francisco: Astronomical Society of the Pacific), 701 Iida, A., Nakamoto, T., Susa, H., & Nakagawa, Y. 2001, Icarus, 153, 430 Kobatake, H., Tsukamoto, K., Nozawa, J., Nagashima, K., Satoh, H., & Dold, P. 2008, Icarus, 198, 208 Kozasa, T., & Hasegawa, H. 1987, Prog. Theor. Phys., 77, 1402 Miura, H., Nakamoto, T., & Susa, H. 2002, Icarus, 160, 258 Miura, H., & Nakamoto, T. 2005, Icarus, 175, 289 Miura, H., & Nakamoto, T. 2006, ApJ, 651, 1272 Miyake, K., & Nakagawa, Y. 1993, Icarus, 106, 20 Mysen, B. O., & Kushiro, I. 1988, American Mineralogist, 73, 1 -- 24 -- Nagahara, H. & Ozawa, K. 1996, Geochim. Cosmochim. Acta., 60, 1445 Nagasawa, M., Lin, D. N. C., & Thommes, E. 2005, ApJ, 635, 578 Nagashima, K., Tsukamoto, K., Satoh, H., Kobatake, H., & Dold, P. 2006, J. Crys. Growth, 293, 193 Nozawa, J., Tsukamoto, K., Kobatake, H., Yamada, J., Satoh, H., Nagashima, K., Miura, H., & Kimura, Y. 2009, Icarus, 204, 681 Susa, H., Uehara, H., Nishi, R., & Yamada, M. 1998, Prog. Theor. Phys., 100, 63 Toriumi, M. 1989, Earth Planet. Sci. Lett., 92, 265 Tsuchiyama, A., Kushiro, I., Mysen, B. O., & Morimoto, N. 1988, Proc. NIPR Symp. Antarct. Meteorites, 1, 185 Tsuchiyama, A., Takahashi, T., & Tachibana, S. 1998, Mineralogical Journal, 20, 113 Tsukamoto, K., Satoh, H., Takamura, Y., & Kuribayashi, K. 1999, Antarct. Meteorites, 24, 179 Wasson, J. T. & Rubin, A. E. 2003, Geochim. Cosmochim. Acta, 67, 2239 Weidenschilling, S. J., Marzari, F., & Hood, L. L. 1998, Science, 279, 681 Yamada, J. 2009, Master's Theses, Department of Earth and Planetary Science, Tohoku University Yamamoto, T., & Hasegawa, H. 1977, Prog. Theor. Phys., 58, 816 Yamamoto, T., Chigai, T., Watanabe, S., & Kozasa, T. 2001, A&A, 380, 373 Yurimoto, H. & Wasson, J. T. 2002, Geochim. Cosmochim. Acta, 66, 4355 This preprint was prepared with the AAS LATEX macros v5.2. -- 25 -- Fig. 1. -- Outline of the model for cosmic crystal formation. The nebular gas and precursor silicate dust come from the left side of the planetesimal and pass through the shock front produced by a planetesimal orbiting at supersonic velocity in the nebular gas. They are heated behind the shock front, and evaporation of the dust particles produce silicate vapor, which comes mainly from evaporation of µm-sized particles. The grayed region indicates existence of the vapor. Pressure, density, and temperature of the gas in the shocked region just before the expansion (t = 0) are denoted by p0, ρ0, and T0, respectively. The shocked region has higher pressure than the ambient region and expands vertically with velocity vr. R(t) is radius of the gas cloud at time t. The silicate vapor cools with expansion and becomes supercooled. Cosmic crystals condense from the cooled vapor after the vapor becomes supersaturated to a certain degree. -- 26 -- Fig. 2. -- Spatial profiles of (a) temperatures (Tg: gas, Td: dust, Trad: ambient radiation) and (b) densities (ρd: dust, ρv: silicate vapor) in the vicinity of the shock front for Rp = 100 km, n0 = 1015 cm−3, vs = 8 km s−1, ξ = 0.01, and ad = 1 µm. Panels (c) and (d) are, respectively, expansions of panels (a) and (b) in the distance scale. Td2 in panel (a) indicates the dust temperature at the beginning of the second stage (see text). -- 27 -- Fig. 3. -- Evaporation fraction η as a function of dust temperature at the second stage, Td2. We take Td2 at the time when the velocity of the dust particles relative to the gas is 1/10 times thermal velocity of the gas molecules. The symbols indicate numerical results and the curves show η calculated with use of Eq. (3) together with Eq. (4) for given ∆t, cooling timescale of the gas. Plotted in the panels are (a) numerical data for all sets of the parameters, (b) those distinguished by the Rp-values given in the panel, (c) those by n0, and (d) those by ξ. -- 28 -- Fig. 4. -- Comparison of Td2, dust temperature in the second stage, obtained by the nu- merical simulations (vertical axis) and those calculated with the use of Eq. (B2) given in Appendix B (horizontal axis). -- 29 -- Fig. 5. -- The same as Fig. 3 but the values of Td2 in the horizontal axis are replaced by those calculated by using Eq. (B2). -- 30 -- 100 analytic 10 1 10 1 analytic approx. (a) (b) 0.1 0.1 1 10 100 Fig. 6. -- Temporal variations of (a) the radius R of the gas cloud and (b) its expansion velocity vr behind a planetesimal bow shock. The curves for γ = 7/5 correspond to adiabatic expansion of a gas composed of H2 molecules. All quantities including time t are normalized (see text for details). The solid curves show exact solutions given by Eq. (C2) for R and by Eq. (C1) for vr, while the dashed curves show approximations in the two limiting cases given by Eq. (13). -- 31 -- n o i t a s n e d n o c m u i r b i l i u q e ~ non-equilibrium condensation 0.2 0.4 0.6 0.8 1 T / T0 5/3 7/5 1.1 0.4 0.35 0.3 0.25 0.2 0.15 0.1 0.05 0 0 ) 0 s t / 0 T ( e t a r g n i l o o c Fig. 7. -- Cooling rate of the vapor produced by a planetesimal bow shock during its expan- sion versus the gas temperature T . The horizontal axis is T normalized by the initial temper- ature T0 and the vertical one is the cooling rate normalized by T0/ts0, where ts0 = R0/cs0 is the sound-crossing time. The solid, dashed, and dotted curves show cooling rates for γ = 5/3, 7/5, and 1.1, respectively. The gray region indicates a range of the equilibrium condensation temperatures of silicates under the total pressure of protoplanetary disk, Te = 1300−1500 K. -- 32 -- Fig. 8. -- Relation between the cooling parameter Λ and the evaporation fraction η. Plotted are η calculated by the numerical simulations and Λ calculated from Eq. (D2): (a) for the data for all all of the parameter sets, (b) for each Rp, (c) for each n0, and (d) for each ξ. The gray region (Λ < 1) indicates the region where condensation through homogeneous nucleation does not take place during the vapor cooling. -- 33 -- Fig. 9. -- Typical radii a∞ of the particles condensed in the vapor and the degree of the supercooling ∆T /Te versus the cooling parameter Λ and the dimensionless surface tension Γ. The gray region indicates the ranges of Λ expected from the model and of possible values of Γ of the condensates. Solid curves indicate a∞ and ∆T calculated based on the homogeneous nucleation theory for Γ = 10 to 60. Dashed lines indicate those for Λ = 1 to 1010. Products of the evaporation and condensation experiments are shown by the red (enstatite crystals, Yamada 2009) and green regions (forsterite crystals, Kobatake et al., 2008). The typical size range of ultra-fine particles in the matrix of Allende meteorite (Toriumi, 1989) is shown by the blue region. -- 34 -- Fig. 10. -- Formation of chondritic materials produced by planetesimal bow shocks in the course of planetesimal growth.
1806.03900
1
1806
2018-06-11T10:54:11
The Mid Pleistocene Transition from a Budyko-Sellers Type Energy Balance Model
[ "astro-ph.EP", "math.DS" ]
A conceptual model of the Plio-Pleistocene glacial cycles is developed based on the Budyko-Sellers type energy balance model. The model is shown to admit a phenomenon like the Mid-Pleistocene transition, capturing the essence of the albedo and the temperature precipitation feedbacks.
astro-ph.EP
astro-ph
Confidential manuscript submitted to Journal XXX The Mid Pleistocene Transition from a Budyko-Sellers Type Energy Balance Model Esther R. Widiasih, Malte F. Stuecker, Somyi Baek 1,2,3 July 9, 2021 1Mathematics and Science Division, University of Hawaii - West Oahu 2Center for Climate Physics, Institute for Basic Science, Pusan National University 3School of Mathematics, University of Minnesota - Twin Cities Corresponding author: Esther R. Widiasih, [email protected] -- 1 -- 8 1 0 2 n u J 1 1 . ] P E h p - o r t s a [ 1 v 0 0 9 3 0 . 6 0 8 1 : v i X r a Confidential manuscript submitted to Journal XXX Abstract A conceptual model of the Plio-Pleistocene glacial cycles is developed based on the Budyko- Sellers type energy balance model. The model is shown to admit a phenomenon like the Mid- Pleistocene transition, capturing the essence of the albedo and the temperature precipitation feed- backs. 1 Introduction 1.1 General problem of glacial cycles and the MPT In the past 800 kyr, the Earth's climate has experienced interglacial-glacial cycles (or sim- ply glacial cycles from here on) with a 100 kyr periodicity. The Milankovitch theory proposes that the glacial cycles is driven by a linear response to the summer insolation signal at high northern latitude. However, while such signal possesses a weak 100 kyr period of the Earth's orbital eccen- tricity, it is dominated by the obliquity signal with a period of 41 kyr. This problem is also known as the 100 kyr problem. An even more fasinating problem is the change in the periodicity of the glacial cycles, that occur about 1.2 - 0.5 Mya, known as the Mid Pleistocene Transition (MPT). Prior to the MPT, the glacial cycles occur at a shorter period of 41 kyr, having a lower amplitude. After the MPT, each cycle in the glacial period lengthens to 100 kyr, having a higher amplitude and a more asymmetri- cal feature resembling a saw-tooth. Both the 100 kyr and the MPT problems call for a non-linear treatment in the modeling of the glacial cycles, having a parameter shift that induces another shift in the response variables. By design, conceptual models are highly simplified representation of the planetary climate, because these models are tools to probe big picture ideas. Numerous conceptual models have found success in exhibiting differing mechanisms for both the 100 kyr glacial cycles and the MPT. A few well studied examples are Paillard and Parrenin [2004], Maasch and Saltzman [1990], and Tziperman and Gildor [2003]. In all of these examples, the state variables are global averages. The model proposed here is based on what's known as the Budyko-Sellers model (for examples, see the discussions in North [1975]; North et al. [1981]; Sellers [1969]; Tung [2007]), a latitudi- nally averaged energy balanced model. The one used in particular is the one first written by M. Budyko Budyko [1969] to model the annual temperature distribution over hemisphere, assuming a symmetric water planet. In using such model, one may benefit by exploiting a balance principle on the latitudinal position of some significant features, such as the glacier's edge and the snowline. Indeed, the glacier's edge and the snowline positions are two of the state variables in the proposed glacial model shall represent, and the third state variable approximates the planetary average tem- perature. The model admits sustained glacial cycles (free oscillations) without any external forc- ing from orbital elements. The details of the modeling process, the limit cycle and other mathe- matical properties of the model are discussed in Walsh et al. [2016]. -- 2 -- Confidential manuscript submitted to Journal XXX 1.2 Background of the Glacial Flip Flop Model Because of the switching feature of this model, this model is called the glacial flip flop model, as it flip and flops between glaciating and de-glaciating states. The idea of a multi-state cli- mate model is not new. For example, in the conceptual model proposed by Paillard [1998], three climate regimes: i (interglacial), g (mild glacial), G (full glacial). The global circulation model ex- plored in Abe-Ouchi et al. [2013] shows two distinct climate regimes, one for a positive and the other for negative total mass balance state. These two models are clearly at the opposite ends of the spectrum in terms of model complexity, and yet, they both show transient evolution of the cli- mate, weaving in and out of these climate regimes. Each multi-state climate model of the glacial cycles discussed above is equipped with a switch mechanism. What might be the driver of the switch? A compelling discussion in Raymo et al. [2006] (par. 15) posed a conjecture that what determines the maximum ice sheet size is perhaps the albedo change causing positive and negative feedbacks affecting the temperature- precipitation feedback. The change in the interplays between the two feedbacks might also be the difference between the early and late Pleistocene. The 3 state variables in the glacial flip flop model captures exactly these interplays as the albedo feedback affects temperature and ice sheet edge maximum latitudinal position. The switching mechanism of the glacial flip flop is governed by the energy balance principle as was done in the Budyko-Sellers type models, and the mass bal- ance principle, in the same spirit as the global circulation model in Abe-Ouchi et al. [2013]. The glacial flip flop model featured here may be reminiscent of the conceptual model of the ice sheet proposed in Weertman [1976]; Källén et al. [1979], featuring the precipitation and temperature feedback. There are many similarities between the two models: the use of Budyko- Sellers type energy balance model to govern the temperature, the parameterization of the latitudi- nal extent of the some ice cover, and finally, the use of mass balance (accumulation-ablation) to drive the ice volume evolution. The difference, however, is apparent to distinguish the two models, with the main difference being the existence of the parametrization of the snow line in the glacial flip flop model. Two prior works staged the foundation of the glacial flip flop model. In the first work Widi- asih [2013], a parametrization of the albedo line in an aquaplanet is incorporated into the existing Budyko energy balance model of the spatially dependent temperature distribution. The albedo line is thought to represent a latitudinal line separating perenially snow covered region from that where the snow melts in the summer, akin to a climactic snow line. Indeed, this is in contrast to most of the Budyko-Sellers type of energy balance model where the albedo function is dependent on the temperature at that latitude. As the albedo feedback depends on the area of differing sur- face albedo, the latitudinally dependent albedo function holds the advantage in representing the yearly average energy absorbed/ reflected by the system. Assuming a spherical planet, when the spatial dependence is represented as the sine of the latitude (as was done in Tung [2007]), a snow cap extending from the pole to the snow line (or albedo line) at latitude θ covers exactly 1-sin(θ) of the planet surface area having higher albedo, while sin(θ) is the surface area with lower albedo. Therefore, the incorporation of the surface albedo feedback could be done more directly. -- 3 -- Confidential manuscript submitted to Journal XXX While the zonally averaged Budyko energy balance model of the temperature distribution is often referred to as one dimensional energy balance model because its spatial dependent variable is one dimensional (see eg. McGuffie and Henderson-Sellers [2005]), mathematically speaking, it is an infinite dimensional system of integro-partial differential equation. The second work staging the foundation of the glacial flip flop model was presented in McGehee and Widiasih [2014], and showed the reduction of that snow line-zonal temperature distribution into a manageable system of ordinary differential equations of two variables. Because of this reduction, the temperature dis- tribution could now be represented by one variable, the evolution of which captures the surface albedo feedback parametrized in the snow line, acting as a line separating lower and higher albedo surface. The addition of the third variable in the glacial flip flop model is done to capture the mass balance interplay between the accumulation zone and the ablation zone. The albedo line of the second work is driven by the temperature at the snow line that controls the maintenance of pere- nial snow cover, and thereby, captures the essence of precipitation and temperature feedback. The third variable, representing the edge of the ice sheet, models the evolution of the ice extent by cap- turing the mass balance between the accumulation zone (poleward of the snow line) and the ab- lation zone (between the snow line and the ice sheet extent). As one considers the thousand years time scale of the glacial cycles, the evolution of the annual average temperature should be consid- ered working at a relatively faster rate, while the snow line and ice line evolution should be at rela- tively slower rates (see the discussion in McGehee and Widiasih [2014]). This modeling approach allows an exploration into the conjecture on what determine the maximum ice sheet as posed by Raymo et al. [2006]. The evolution of the snow line (or albedo line) is driven by the temperature difference be- tween the snowline temperature and a critical temperature. When the snowline temperature is warmer than a critical temperature, the snowline retreats toward the pole, otherwise, it advances toward the equator. The modeling approach incorporating a critical ice-forming temperature has also been utilized by others, for example in Tziperman and Gildor [2003]. In that work, the ice- forming temperature is varied slowly from -13oC 1.5 Ma BP to -3oC at 500 kyr BP in order to simulate the deep ocean cooling. The recent work by Tzedakis et al. [2017] presented a statisti- cal model correctly predicting every complete deglaciation over the past 1 Myr. This statistical model incorporates the observations that the energy threshold to start deglaciation has increased in the past 1 Myr and that the increase in the energy threshold causes skipped deglaciation, which in turn results in larger ice sheet and longer glaciation. The shift in the ice-forming critical tem- perature modeling approach done by Tziperman and Gildor [2003] is certainly consistent with the more recent statistical finding. As the discussion of this model shows in the next section, the same shift in the ice-forming critical temperature parameter drives a bifurcation, which in turn causes an increase in the amplitude and the period of the glacial cycles. Here, we pose the main question of this work: given that the incoming solar radiation forc- ing is a function of eccentricity and obliquity, could the glacial flip flop model simulate the transi- tion from 41 kyr to 100 kyr glacial cycle period? Could the model simulate the transition even if the eccentricity is held constant? In the following sections, we discuss the details of the glacial flip -- 4 -- Confidential manuscript submitted to Journal XXX flop model, some mathematical interpretations of the model, and some simulations showing the glacial cycles and the transition from the 41kyr to 100kyr period. 2 Model description The glacial flip flop model assumes the existence of two regimes: glaciation and inter-glaciation, and the system flips and flops between the two regimes. The distinguishing features of the two regimes lies on two parameters: the ice forming critical temperature and the ablation rate. The ice forming critical temperature of glaciation regime is higher than that of the inter-glaciation regime. The three model variables, w, η , and ξ, respectively simulate the global annual temperature, the sine of the latitude of the snow/ albedo line, and the sine of the latitude of the maximum ice cover extent (ice extent from here on). Hence, variables w are real valued, while η and ξ live on the unit interval [0, 1], with 0 representing the equator and 1 the pole. Below we briefly present the formu- lation of the glacial flip flop model, a detailed explanation of the modeling can be found in Walsh et al. [2016]. The evolution of the temperature w, derived from the Budyko energy balance model, reflects the ice albedo feedback, and so, its equation of motion is coupled with the albedo line η. The mo- tion of the albedo line η reflects the precipitation and temperature feedback, hence, its evolution equation is coupled with temperature w and it is dependent on the ice forming critical temperature. The evolution of the ice extent ξ relies on the mass balance principle, hence its equation of motion keeps track of the difference between ablation rate and accumulation rate. Ablation happens on the zone between ξ and η, while accumulation takes place between η and the pole, 1. The evolu- tion of the maximum ice extent ξ depends linearly on the size of the ablation zone η − ξ and the accumulation zone 1 − η. 2.1 The variables and equations During a glacial period, that is when accumulation exceeds ablation, the ice extent is ad- vancing. The equations of motion for w, η, and ξ are: (cid:219)w = −τ(w − F(η)) = f a 1 (w, η, ξ) (cid:219)η = ρ(w − Ga(η, Tca)) = f a (cid:219)ξ =  (ba(η − ξ) − a(1 − η)) = f a 2 (w, η, ξ) 3 (w, η, ξ). (1a) (1b) (1c) During an inter-glacial period, that is when the ice extent is retreating, the equations of mo- tion for w, η, and ξ are: -- 5 -- Confidential manuscript submitted to Journal XXX (cid:219)w = −τ(w − F(η)) = f r 1 (w, η, ξ) (cid:219)η = ρ(w − Gr(η, Tcr)) = f r (cid:219)ξ =  (br(η − ξ) − a(1 − η)) = f r 2 (w, η, ξ) 3 (w, η, ξ). (2a) (2b) (2c) The two regimes have two distinct ice forming critical temperature Tca and Tcr as well as ablation rates ba and br. The value of the interglacial ice forming critical temperature as the ice extent is retreating is set to Tcr ≈ −10oC, and is close to that of today's climate. This value is used eg. in Tung [2007]. The interglacial period ice forming temperature Tcr is lower than Tca ≈ −6o to − 10oC. Note that the range of both the ice forming critical temperatures are well within that used by Tziperman and Gildor [2003]. The ablation rate or advancing ice sheet, ba is lower than br, the rate when the ice extent is retreating. The switch between the two regimes happens at the discontinuity plane, when b(η − ξ)− a(1− ξ) = 0, where b is some critical ablation rate, having the value ba < b < br. In this model, the orbital forcing enters through the function F(η) due to its dependence on the incoming solar radiation parameter Q (which depends on the eccentricity) and the distribu- tion parameter s2 (which depends on obliquity). Since the variables in the flip flop model simu- late annual averages, eccentricity and obliquity dependent orbital elements can be included in the solar insolation function, leaving out any treatment on precession. The details of the functions F, Ga, Gr, and the parameters used can be found in TABLE 1. 2.2 Some mathematical interpretations Within some reasonable parameter regimes, the glacial flip flop model is able to generate a cycle with varying amplitude and periodicity. In Figure 1, the only varying parameter is the ice forming critical temperature during the glaciation period Tca, which is increased from −9oC to −5oC. In this particular exploration, both solar orbital parameters affecting eccentricity Q and obliquity s2 are fixed constant. The amplitude and period of the cycle increases with Tca. This prompts a current work in Makarenkov and Widiasih [2018] showing the existence of a bifurcation of a limit cycles from a singularities of the nonsmooth system in equations (1) and (2). The dynamics of the cycle is controlled by the two virtual equilibria (blue and red dots in the phase plane of Figure 1). The trajectory of the glaciating regime is trying to reach the virtual equilibrium associated with the glaciating regime (the blue dot), while that of the deglaciation regime is aiming for the interglacial virtual equilibrium (the red dot). In particular, notice that in the glacial flip flop model, the parameter Tca controls the start of the deglaciation; when Tca is closer to Tcr, the deglaciation starts earlier than when Tca is at a higher value. Hence, exploration in the direction of increasing ice forming threshold Tca will be consistent with the previously cited work by Tzedakis et al. [2017] on the increasing threshold for deglaciation as well as with a similar modeling experiment done by Tziperman and Gildor [2003]. -- 6 -- Confidential manuscript submitted to Journal XXX Drawing inspirations from this, two ramp functions for Tca and Tcr increasing over the past 5 Myr may be used to capture a phenomenon like the Mid-Pleistocene Transition. Using ramp functions shown in Figure 2, phase portraits of the trajectory of the system as well as the resulting maximum ice extent ξ appears to show a transition in the period and amplitude of the system. This particular exploration shows that the cycle in the glacial flip flop system is robust to change in the ice forming critical temperature parameter, even as this parameter goes through a time dependent change. Furthermore, the cycles exist even in the absence of the orbital forcing parameters, cap- turing the climate system's internal periodic behavior, akin to that exhibited by other conceptual models eg. Maasch and Saltzman [1990]. The next exploration is aimed at obtaining a time series for the maximum ice extent ξ so its period and amplitude over the past 5 Myr mimic that of the ice volume proxy δ18O record from Lisiecki and Raymo [2005]. The results of are shown in Figure 3. To do this, eccentricity dependent parameter Q and obliquity dependent s2 are varied in time using the orbital data from Laskar et al. [2004]. The formulation for Q and s2 as functions of eccentricity and obliquity for the Budyko modeltakes a similar approach as done in McGehee and Lehman [2012]. Interestingly, the lowest mode of the singular spectrum analysis (SSA) method of ice volume proxy δ18O record from Lisiecki and Raymo [2005] shows an increasing trend in ice volume. To explore further the response of the system to the nature of the increase in the ice forming temperature threshold, a linearly transformed first SSA mode of the δ18O from the past 5 Myr is used as an ice forming critical temperature threshold Tca. 3 Discussion A Budyko-Sellers type energy balance model is used as the basis a glacial cycle model called the glacial flip flop. The conceptual model captures two important feedbacks: albedo and temperature-precipitation. A change in the ice forming critical temperature captures an MPT like phenomenon in the model. While the reasoning for the change in the critical temperature is be- yond the scope of this work involved in developing the model, the model highlights some key non-linear feed backs that should be explored further. Furthermore, the latitudinally dependent variables developed in the model provide novel tools for further conceptual explorations of the Plio-Pleistocene glacial cycles. Mathematically, the non-smooth nature of the model is also inter- esting. One might wonder what would happen to the system if it is allowed to switch between the glacial and interglacial regimes. -- 7 -- Confidential manuscript submitted to Journal XXX (a) From left to right, the phase portraits of w − η − ξ of the autonomous system with varying in equation (1a) the critical ice forming temperature T ca = −9, −7, −5oC, keeping T cr = −10oC. The mass balance is positive in the blue region, and negative in the red region. The diagonal plane along the w axis, nearly orthogonal to the page is the discontinuity plane separating the glaciation and interglaciation regimes. The red and blue dots are the virtual equilibria of the system in equations (1) and (2) (b) The corresponding global mean temperature evolution for T ca = −9, −, −5oC. (c) The corresponding ice extentξ and snow line η evolution for T ca = −9, −, −5oC. Figure 1: The behavior of the glacial flip flop system as the ice forming critical temperature Tca is varied. Keeping Tcr constant, the amplitude and the period of the cycle grow as Tca increases. -- 8 -- 0100200300400500t13.814.014.214.414.614.8Tbar100200300400500t1112131415Tbar100200300400500t468101214Tbar100200300400500t0.880.900.920.94η(blue),ξ(red)100200300400500t0.750.800.850.900.95η(blue),ξ(red)100200300400500t0.50.60.70.80.91.0η(blue),ξ(red) Confidential manuscript submitted to Journal XXX (a) The time dependent ramp functions for T ca(t) (purple) and for T cr(t) (pink), are used in this explo- ration (time t = 0 is 5 Myr). The resulting time series of the ice sheet extent variable ξ (orange) shows growth in period and amplitude. (b) The phase portrait of the glacial flip flop system with increasing ramp functions for T ca and T cr . Trajectory in yellow is from 5Myr to 4Myr, orange is from 4 Myr to 1Myr, and red is the last 1 Myr. Figure 2: Glacial flip flop model using time depenendt ramp functions. -- 9 -- 010002000300040005000t0.700.750.800.850.900.95ξ010002000300040005000t(kyr)-11-10-9-8-7Tc Confidential manuscript submitted to Journal XXX (a) The phase portrait of the system, with orbital forcing from Laskar et al. [2004], using ice forming temperature threshold T ca (purple) from LR04 record's first SSA mode and a linearly increasing T cr (pink). The LR04 record is from Lisiecki and Raymo [2005]. (b) The resulting lowest mode of the SSA reconstruction for both LR04 record and ice extent variable ξ from the experiment using the ice forming temperature function above. Figure 3: Glacial flip flop, with orbital forcing and empirical ice forming critical temperature function. -- 10 -- 10002000300040005000tkyr-11.0-10.8-10.6-10.4-10.2-10.0-9.8Tc Confidential manuscript submitted to Journal XXX References References Abe-Ouchi, A., F. Saito, K. Kawamura, M. E. Raymo, J. Okuno, K. Takahashi, and H. Blatter (2013), Insolation-driven 100,000-year glacial cycles and hysteresis of ice-sheet volume, Na- ture, 500(7461), 190. Budyko, M. I. (1969), The effect of solar radiation variations on the climate of the earth, tellus, 21(5), 611 -- 619. Källén, E., C. Crafoord, and M. Ghil (1979), Free oscillations in a climate model with ice-sheet dynamics, Journal of the Atmospheric Sciences, 36(12), 2292 -- 2303. Laskar, J., P. Robutel, F. Joutel, M. Gastineau, A. Correia, and B. Levrard (2004), A long-term numerical solution for the insolation quantities of the earth, Astronomy & Astrophysics, 428(1), 261 -- 285. Lisiecki, L. E., and M. E. Raymo (2005), A pliocene-pleistocene stack of 57 globally distributed benthic δ18o records, Paleoceanography, 20(1). Maasch, K. A., and B. Saltzman (1990), A low-order dynamical model of global climatic variabil- ity over the full pleistocene, Journal of Geophysical Research: Atmospheres, 95(D2), 1955 -- 1963. Makarenkov, O., and E. Widiasih (2018), Bifurcations of limit cycles from a fold fold singularities in a glacial cycles model, ARXIV. McGehee, R., and C. Lehman (2012), A paleoclimate model of ice-albedo feedback forced by variations in earth's orbit, SIAM Journal on Applied Dynamical Systems, 11(2), 684 -- 707. McGehee, R., and E. Widiasih (2014), A quadratic approximation to Budyko's ice albedo feed- back model with ice iine dynamics., SIADS, Vol. 13, Issue 1, DOI: 10.1137/120871286. McGuffie, K., and A. Henderson-Sellers (2005), A climate modelling primer, John Wiley & Sons. North, G. (1975), Theory of energy-balance climate models, J. Atmos. Sci, 32(11), 2033 -- 2043. North, G. R., R. F. Cahalan, and J. A. Coakley Jr (1981), Energy balance climate models, Reviews of Geophysics, 19(1), 91 -- 121. Paillard, D. (1998), The timing of pleistocene glaciations from a simple multiple-state climate model, Nature, 391(6665), 378. Paillard, D., and F. Parrenin (2004), The antarctic ice sheet and the triggering of deglaciations, Earth and Planetary Science Letters, 227(3-4), 263 -- 271. Raymo, M., L. Lisiecki, and K. H. Nisancioglu (2006), Plio-pleistocene ice volume, antarctic cli- mate, and the global δ18o record, Science, 313(5786), 492 -- 495. Sellers, W. D. (1969), A global climatic model based on the energy balance of the earth- atmosphere system, Journal of Applied Meteorology, 8(3), 392 -- 400. Tung, K.-K. (2007), Topics in mathematical modeling, Princeton University Press Princeton, NJ. Tzedakis, P., M. Crucifix, T. Mitsui, and E. W. Wolff (2017), A simple rule to determine which insolation cycles lead to interglacials, Nature, 542(7642), 427. Tziperman, E., and H. Gildor (2003), On the mid-pleistocene transition to 100-kyr glacial cycles and the asymmetry between glaciation and deglaciation times, Paleoceanography, 18(1). -- 11 -- Confidential manuscript submitted to Journal XXX Walsh, J., E. Widiasih, J. Hahn, and R. McGehee (2016), Periodic orbits for a discontinuous vec- tor field arising from a conceptual model of glacial cycles, Nonlinearity, 29(6), 1843. Weertman, J. (1976), Milankovitch solar radiation variations and ice age ice sheet sizes, Nature, 261(5555), 17. Widiasih, E. R. (2013), Dynamics of the budyko energy balance model, SIAM Journal on Applied Dynamical Systems, 12(4), 2068 -- 2092. -- 12 --
0904.0402
1
0904
2009-04-02T15:00:17
A thermodynamic basis for prebiotic amino acid synthesis and the nature of the first genetic code
[ "astro-ph.EP", "astro-ph.SR", "q-bio.BM" ]
Of the twenty amino acids used in proteins, ten were formed in Miller's atmospheric discharge experiments. The two other major proposed sources of prebiotic amino acid synthesis include formation in hydrothermal vents and delivery to Earth via meteorites. We combine observational and experimental data of amino acid frequencies formed by these diverse mechanisms and show that, regardless of the source, these ten early amino acids can be ranked in order of decreasing abundance in prebiotic contexts. This order can be predicted by thermodynamics. The relative abundances of the early amino acids were most likely reflected in the composition of the first proteins at the time the genetic code originated. The remaining amino acids were incorporated into proteins after pathways for their biochemical synthesis evolved. This is consistent with theories of the evolution of the genetic code by stepwise addition of new amino acids. These are hints that key aspects of early biochemistry may be universal.
astro-ph.EP
astro-ph
A thermodynamic basis for prebiotic amino acid synthesis and the nature of the first genetic code Paul G. Higgs and Ralph E. Pudritz Origins Institute and Department of Physics and Astronomy, McMaster University, Hamilton, Ontario. Contact - [email protected]; [email protected] Running title - Prebiotic amino acid synthesis Abstract Of the twenty amino acids used in proteins, ten were formed in Miller’s atmospheric discharge experiments. The two other major proposed sources of prebiotic amino acid synthesis include formation in hydrothermal vents and delivery to Earth via meteorites. We combine observational and experimental data of amino acid frequencies formed by these diverse mechanisms and show that, regardless of the source, these ten early amino acids can be ranked in order of decreasing abundance in prebiotic contexts. This order can be predicted by thermodynamics. The relative abundances of the early amino acids were most likely reflected in the composition of the first proteins at the time the genetic code originated. The remaining amino acids were incorporated into proteins after pathways for their biochemical synthesis evolved. This is consistent with theories of the evolution of the genetic code by stepwise addition of new amino acids. These are hints that key aspects of early biochemistry may be universal. Keywords – amino acid synthesis; meteorites; hydrothermal vents; genetic code; protein evolution; thermodynamics. 1 Introduction - Possible Mechanisms of Prebiotic Amino Acid Synthesis Among the organic molecules that are constituents of cells, amino acids stand out by virtue of their role as building blocks of proteins and their ease of chemical formation. There has been clear evidence for prebiotic formation of amino acids since the experiments of Miller (Miller, 1953; Miller and Orgel, 1974) involving electrical discharge in a mixture of atmospheric gases, which produced 10 of the amino acids used in modern proteins plus many other organic molecules. Formation of prebiotic amino acids in interstellar gas, as well as in hydrothermal vents, has since been proposed and tested. We will show that these diverse environments are remarkably consistent about which amino acids formed. The distinction between early and late amino acids is important for theories of the origin of the genetic code (i.e. the mapping between codons and amino acids – a codon is the combination of 3 bases in a nucleic acid sequence needed to specify an amino acid). Wong (1975; 2005) argues that early versions of the genetic code encoded only a few amino acids and that later amino acids were added as biosynthetic pathways evolved. With a view to determining the order of addition of amino acids to the genetic code, Trifonov (2004) collected 60 criteria by which the amino acids can be ranked. The consensus order that emerges from these criteria appears to be a useful one; however, we have reservations because diverse criteria were included that are not all consistent with one another. Our analysis below uses a similar ranking procedure but incorporates only criteria based on measurable amino acid concentrations. First, let us consider astrophysical sites for the formation of organic molecules. Interstellar molecular cloud cores and the protostellar disks that form from their collapse are a universal aspect of star and planet formation. Observations of disks around other stars reveal that active chemistry occurs near the surface regions of protoplanetary disks (Bergin et al. 2007). Less volatile gases, such as CO, NH3 and HCN, freeze out on dust grains and organic molecules can be synthesized on the surface of the grains after exposure to ultra-violet radiation. Alternatively, it has been proposed that amino acids can form during the gravitational collapse of a core, and predictions have been made regarding molecular concentrations and locations within the disk (Chakrabarti and Chakrabarti, 2000). Molecules in protoplanetary disks end up in comets and meteorites as the dust grains coalesce and are then delivered to Earth when impact events occur. Meteorites forming in the icy outer regions of the solar system can later impact the Earth if their orbits are perturbed by large outer planets. This provides a 2 way of bringing organic molecules and water to Earth that formed on grains and in meteorites at larger orbital radii - beyond the snow lines for these ices (for water, typically at or beyond 2.5 Astronomical Units). Delivery of molecules from outside Earth is only relevant for the origin of life if the rate is comparable to or exceeds the rate of formation of the same molecules by Earth- based chemistry. Several authors have tried to estimate the relative rates of these internal and external processes, although it is extremely difficult to do this with certainty (Whittet, 1997; Pierazzo and Chyba, 1999; Bernstein, 2006). Observation of the chemical composition of meteorites provides us with a time capsule to the kind of organic chemistry that was occurring in the early stages of the solar system when the meteorite parent bodies were formed. In contrast, chemistry on Earth has been dramatically changed by subsequent geological and biological activity. Thus, meteoritic compositional data and extrapolated impact rate are relevant for understanding pre-biotic conditions and suggest that meteorites contributed significantly to the organics present on Earth over 4 GA ago - at delivery rates of the order 108 kg per yr (Pierazzo and Chyba, 1999). Table 1 gives the relative concentrations of amino acids measured in meteorites. Column M1 is the Murchison meteorite (H2O extract from Table 2 of Engel and Nagy, 1982), M2 is the Murray meteorite (Cronin and Moore, 1971), and M3 is the Yamato meteorite (interior, hydrolysed sample from Shimoyama et al. 1979). The possibility that dust grains with icy mantles are a site for amino acid synthesis has been investigated experimentally (Bernstein et al. 2002; Munoz Caro et al. 2002). Several amino acids may result from the exposure of icy mantles of HCN, NH3, and H2CO to UV radiation. Column I1 shows measurements from Munoz Caro et al. (2002). We next turn to Miller’s experiments, which showed that synthesis was possible in the atmosphere of the early Earth - presumed to be reducing. Column A1 shows the results with an atmosphere of CH4, NH3, H2O and H2 (Miller and Orgel, 1972), and column A2 shows the results with an atmosphere of CH4, N2, H2O and traces of NH3 (Miller and Orgel, 1972). There have since been many claims that the early atmosphere was not strongly reducing (Kasting, 1993; Pierazzo and Chyba, 1999), but the importance of atmospheric synthesis is still maintained by some. Schlesinger and Miller (1983) have compared the quantities and the diversity of amino acids formed from various atmospheric mixtures. Although yields are lower in non-reducing atmospheres, synthesis is still possible in mixtures that are not strongly reducing, such as the experiment using proton irradiation of an atmosphere of CO, N2 and H2O 3 (Miyakawa et al. 2002), the results of which are shown in column A3. Another very recent result (Cleaves et al. 2008) on spark discharge in neutral gas mixtures dominated by CO2 and N2 shows that yields of amino acids in these conditions may be higher than previously estimated. Interestingly, two recent calculations regarding the atmosphere on early Earth (Tian et al. 2005; Schaefer and Fegley, 2007) claim that it may have been reducing after all and that the rates of organic synthesis may in fact be high. In the experiments above, amino acids are formed by Strecker synthesis (Miller, 1957), i.e. energy sources such as spark discharge or UV produce HCN and aldehydes, which dissolve in water and subsequently form amino acids. The experiments on atmospheric synthesis are relevant to us here, whatever the predominant atmospheric conditions may have been, because they illustrate the generality of the Strecker synthesis. One reason why meteorite compositions are similar to that produced in atmospheric discharge experiments is that Strecker synthesis may occur in the parent bodies of the meteorites (Wolman et al. 1972). Hydrothermal vents are local environments on Earth that are often proposed as sites for the origin of life (Baross and Hoffman, 1985; Amend and Shock, 1998). Vents are sources of reducing gases (H2, H2S etc.) that might drive synthesis of organic molecules. They are also habitats for current ecological communities where primary production is chemosynthetic not photosynthetic. Studies of amino acid formation in hydrothermal conditions have been reported. Columns H1 (Marshall, 1994) and H2 (Hennet et al. 1992) give the number of times that the amino acid was observed in greater than trace amounts in a large number of experiments. There may be some doubt as to whether the conditions used in these experiments are realistic simulations of what would occur in real hydrothermal vents. Indeed, there is still a debate as to whether amino acid synthesis can occur at all in such conditions (Miller and Bada, 1988; Bada et al. 1995). However, these are the best available experimental data that support the argument for hydrothermal synthesis, and we therefore include them in our analysis as we do not want to bias our data choice with respect on one theory or another. The final three columns in Table 1 are miscellaneous chemical synthesis experiments: S1 is shock synthesis from a gaseous mixture (Bar-Nun et al. 1970); S2 is synthesis from ammonium cyanide (Lowe et al. 1963, Exp. 2); S3 is synthesis from CO, H2 and NH3 at high temperature in the presence of catalysts (Yoshino et al. 1971). This column shows the number of positive identifications of the amino acid from many experiments. These additional experiments are included to illustrate the diversity of conditions that have been proposed as 4 relevant for prebiotic synthesis of amino acids. In choosing the data sets in Table 1, we have tried to include a wide range of points of view, and to emphasize that the nature of the amino acids that are formed in this diverse collection of data is surprisingly consistent. We have only included data on the twenty biological amino acids that are used in modern proteins. We have not included non-biological amino acids, even though some of these occur in meteorites and spark discharge experiments at relatively high concentrations. Our aim is to understand the order of addition of the biological amino acids to the genetic code, and the non-biological amino acids are not relevant for this question. Also, our argument below relies on comparison with thermodynamic data, which is not available for the non-biological amino acids. In the final section of the paper, we return to the question of why the non-biological amino acids were not incorporated into the code. Results - Consensus on Prebiotic Amino Acid Abundances We use two independent measures to ascertain amino acid frequencies; a ranking procedure (Robs), as well as relative concentrations (Crel). In our first approach, amino acids were ranked according to each criterion in Table 1. For example, for criterion M2, glycine (G) is most frequent and is given rank 1. The next two, D and E, are equally frequent, and are both given rank 2.5. The next three are A, V and P, with ranks 4, 5 and 6. All the remaining amino acids are not observed, and are given an equal bottom rank 13.5 (the average of the numbers between 7 and 20). The mean of these ranks from the 12 observations is termed Robs. There is good evidence for prebiotic synthesis of all the amino acids down as far as T on the list. The following four (KFRH) are not found in the Miller experiments or in meteorites and occur only in one or two of the other experiments. The remaining amino acids (NQCYMW) are not formed in any of the experiments. The data are not sufficient to make a definite distinction between KFRH and NQCYMW. Our ranking, Robs, is close to the ranking from Table V of Trifonov (2004), which we term Rcode (see Table 2), as it is based on criteria related to genetic code evolution. The top three are the same, and nine amino acids are in the top ten of both orders. We prefer Robs on the grounds that it is derived directly from experimental observables. We will refer to the first ten amino acids in our ranking as the early group, and the bottom ten as the late group. Our hypothesis is that the early amino acids were frequent in the environment at the time of the origin of the genetic code and the first proteins, but that the late amino acids were rare or 5 absent and were only incorporated into proteins when biochemical pathways arose to synthesize them. Amend and Shock (1998) calculated ΔGsurf, the free energy of formation of the amino +, and H2 in surface seawater conditions (18oC, 1 atmosphere). Figure 1 acids from CO2, NH4 reveals the intriguing result that Robs is strongly correlated with ΔGsurf for the 10 early amino acids (r = 0.96). The late amino acids have significantly higher ΔGsurf than the early ones. The means and standard deviations of the early and late groups are 169.3 ± 42.0 and 285.0 ± 94.1 kJ/mol. Table 2 also lists the molecular weight of each amino acid and the number of ATP molecules required to synthesize the amino acids using the biochemical pathways in E. coli (Akashi and Gojobori, 2002). Amino acid biosynthesis pathways branch off from the central pathways of sugar metabolism in the cell (glycolysis and the citric acid cycle). Akashi and Gojobori (2002) calculated the average ATP cost for growth on glucose, acetate and malate as carbon sources, assuming availability of ammonia and sulphate as sources of nitrogen and sulphur. The means and standard deviations of MW for early and late groups are 116.2 ± 20.6 and 157.3 ± 23.2 Da, and the figures for ATP cost are 18.5 ± 6.9 and 36.2 ± 17.3. It is clear that the late group are larger and more thermodynamically costly. We have omitted the two sulphur- containing amino acids, Met and Cys, from Figure 1. Several of the experiments in Table 1 do not include sulphur in the reaction mixture, so these amino acids are bound to be absent from the products. Thus, the value of Robs for Met and Cys is unclear. Nevertheless, Met and Cys are absent from the meteorites, and are late according to Rcode, so it seems reasonable to classify them in the late group. The central thermodynamic point is that the formation reactions are endergonic (ΔGsurf > 0). Those with the smallest ΔGsurf should therefore be formed most easily, as they require the least free energy input, as is seen in Figure 1. Our second measure of amino acid abundances is simply the mean value, Crel, of the relative concentrations of the amino acids for 9 data sets (excluding H1, H2 and S3). This is shown in Table 2. Figure 2 shows that Crel drops off roughly exponentially with ΔGsurf. This explains why the amino acids with high ΔGsurf are not seen in experiment: their concentration would be too low to detect. The line in Figure 2 is the exponential C G 8.15 exp( , )3.31/ = Δ− rel surf 6 which is the best exponential fit through the first nine amino acids (excluding T as it is an outlier). An exponential curve would be expected based on thermal equilibrium. Given that the ideal gas constant is 8.3 J/mol-K, 31.3 kJ corresponds to a temperature of 3.7 × 103 K. However, the exponential dependence may also be a result of the kinetics of the formation process, with progressively more complex molecules having slower formation rates, in which case the exponential decay constant cannot be interpreted as a temperature. Isotopic studies of the Murchison meteorite suggest the importance of the formation kinetics (Pizzarello et al. 2004). Another important caveat is that amino acids may differ in their long-term stability, so that relative concentrations seen in the meteorites will depend on the rates of decomposition to some degree. The free energy of formation under hydrothermal conditions, ΔGhydro (100oC, 250 atmospheres) was also calculated by Amend and Shock (1998). Formation of some of the amino acids is exergonic (ΔGhydro < 0) under hydrothermal conditions, and even the endergonic ones are less positive than they are at the surface. This was used to support the idea that the first organisms might have been deep-sea chemosynthesizers. However, we find that there is no correlation between Robs and ΔGhydro, as seen in Figure 3. The highest ranking amino acids are intermediate in ΔGhydro, the late amino acids span the full range from lowest to highest ΔGhydro, and there is no significant difference in ΔGhydro between early and late amino acids. Although we would not expect ΔGhydro to be a predictor of what amino acids are seen in meteorites or the atmospheric discharge experiments, it is puzzling that it also does not predict which amino acids are seen in the two experiments designed to simulate hydrothermal systems (H1 and H2). In fact, these experiments give results that agree fairly well with the combined ranking from the other data, rather than the calculated ΔGhydro. This may be an indication that relative concentrations in these experiments are controlled by kinetic factors and that the simplest amino acids are quickest to form, even though some of the larger amino acids have much lower ΔGhydro. Another surprise is that thermodynamic calculations in surface seawater conditions make useful predictions about the amino acids in meteorites. This may be an indication that water-based chemistry is important also in the meteorites. Indeed, it has been argued (Martins et al. 2007) that the organic content of meteorites is strongly influenced by aqueous alteration inside the meteorite parent body, and that differences in this alteration may account for the large differences in the content found among carbonaceous chondrites. 7 It is known that modern organisms are sensitive to the thermodynamic costs of amino acid synthesis. It has been observed that amino acids with lower ATP costs of synthesis have higher frequencies in highly expressed proteins (Akashi and Gojobori, 2002; Seligman, 2003). There is also some suggestion that the differences between synthesis costs in hyperthermophiles and mesophiles may cause different amino acids to be preferred in high expression proteins in hyperthermophiles (Swire, 2007). There are definitely differences in amino acid frequencies between proteins in hyperthermophiles and their mesophile relatives, but this can arise from selection for stability of proteins at high temperatures (Berezovsky et al. 2005) and does not necessarily indicate selection to economize on amino acid synthesis. Discussion - Implications for the origin of the genetic code and protein evolution The range of different specific functions to which proteins have adapted in modern organisms is astounding. As far as we know, proteins cannot achieve replication and heredity by themselves. In modern organisms, proteins are synthesized by translation, which is crucially dependent on mRNAs, rRNAs and tRNAs. This is the main reason for believing that an RNA world existed early in the history of life (Joyce, 2002) in which both genetic and catalytic roles were performed by RNA molecules. It is possible that amino acids or short peptides played some part in the RNA world, or that there were molecules with short peptide sequences attached to nucleotides, like some current coenzymes. However, it was the invention of translation and the genetic code that coupled the hereditary mechanism provided by template- directed replication of nucleic acids to the catalytic possibilities of proteins. Prior to the genetic code, proteins could not evolve. The standard genetic code is shown in Figure 4. Amino acids are coded by between 1 and 6 codons, with most amino acids being coded by either 2 or 4 codons. It is known that this standard code evolved prior to the common ancestor of all current organisms, and it has remained unchanged since then in the majority of genomes of archaea, bacteria and eukaryotes. There are numerous cases of codon reassignments that have occurred much more recently in specific lineages of mitochondrial genomes and some others (Knight et al. 2001; Sengupta et al. 2007). However, it is the early phase of code evolution leading to the establishment of the standard code that is relevant to the current paper. Early versions of the code probably used a smaller repertoire of amino acids, each coded by a larger number of codons. The codon table was divided up into progressively smaller blocks as successive amino acids were added. This 8 was suggested as far back as Crick (1968). This idea is also central to the coevolution theory for the genetic code (Wong, 1975; 2005; Di Giulio and Medugno, 1999; Di Giulio, 2005) which considers addition of amino acids to the code in an order that is determined by the way the amino acid biosynthetic pathways evolved. The coevolution theory proposes the same set of early amino acids as we do, based on the argument that these are synthesized in Miller’s experiments, and because the simplest amino acids are found at the beginnings of the biosynthetic pathways. The strong correlation between Robs and ΔGsurf found above and also the close agreement with Rcode (Trifonov, 2004) suggest that our ranking is a meaningful prediction of the relative frequencies of abiotically synthesized amino acids that would have been available to the first organisms. Our results support the coevolution theory in that they show that none of the proposed mechanisms of non-biological synthesis is able to synthesize the late amino acids; hence the late amino acids could only have been incorporated into the code after biochemical synthesis pathways arose. Conversely, however, our results show that the early amino acids can be synthesized by several non-biological means, and were likely to have been available in the environment. Hence, biosynthetic pathways were not necessarily relevant at the very early stages on code evolution when only a small number of amino acids were encoded. Another key point of the coevolution theory is that the precursor-product relationships between amino acids determine the positions in the code into which new amino acids are assigned. Specifically, new amino acids are assigned to codons that previously belonged to their precursors (Wong, 1975; Di Giulio and Medugno, 1999). In our own view, however, it is the physical properties of the amino acids that have a greater influence on which amino acids are assigned to which codons. The layout of the code is far from random. It is important to note that codons that differ by only one out of the three base positions are often assigned to amino acids with similar properties. As a result, the effects of mutations and translational errors are minimized. The canonical genetic code is better than all but a tiny fraction of randomly rearranged codes in this respect (Freeland et al. 2003). This suggests the strong role of natural selection in building up the code. Selection will favour the assignment of new amino acids to codons that previously coded for amino acids with similar properties (Higgs and Pudritz, 2007; Higgs 2009) because this will be minimally disruptive to the genes that were already encoded by the smaller amino acid set before the new one was introduced. As a result, neighbouring codons in the final code will end up with similar properties. 9 It is well known that hydrophobicity of the amino acids correlates with the column in the code rather than the row (Jungck, 1978; Taylor and Coates, 1989). A principle component analysis of amino acid physical properties (Urbina et al. 2006) shows that amino acids in the same column cluster in property space, and that this has observable consequences for the rate of sequence evolution at first and second codon positions and for the degree to which amino acid frequencies in proteins are influenced by biased mutation at the DNA level. We suggest that the current columnar structure in the current standard code has been retained from the very earliest stages of code evolution. An intriguing observation in this regard is that the five earliest amino acids according to our ranking (Gly, Ala, Asp, Glu, Val) appear on the bottom row of the genetic code table (first position base G). This suggests to us that the second base was the most important discriminator in the early code. The early code may have had a four-column structure in which all codons with the same second base would have coded for the same amino acid: NUN for Val, NCN for Ala, NAN for Asp or Glu and NGN for Gly (where N denotes any base). Selection would have favoured addition of further amino acids into particular columns of the code table according to physicochemical properties, so that hydrophobic amino acids end up in the Val column, hydrophilic amino acids end up in the Asp/Glu column etc. Higgs (2009) has considered the way selection would favour the pathways of code evolution from the four- column code to the current standard code. Another related idea is the two-out-of-three argument of Lagerkvist (1978; 1986), which was recently extended by Lehmann and Libchaber (2008). This proposes that the wobble base pair (first anticodon base with the third codon base) is much less important that the other two pairs in the codon-anticodon interaction. The usual wobble rules allow a tRNA with a U at the wobble position to translate codons ending A and G, and a tRNA with G at the wobble position to translate codons ending U and C. Hence, two tRNAs are needed to translate a four-codon family. According to the two-out-of-three argument, there is also a significant probability that the wobble-U tRNA can translate all four codons, provided that the pairing at the first and second positions is strong. Thus, when there is strong pairing at the first two positions, the four- codon box must be assigned to a single amino acid, whereas when the pairing is weak, the two tRNAs do not translate the same codons, so it is possible to split a four-codon box into two groups of two codons. This argument potentially gives an explanation of the positions of the two-codon and four-codon families in the code. 10 This leads to the question of how many amino acids can be encoded by a system with 64 codons and why there are 20 amino acids in the standard code. There should be a selective advantage of adding another amino acid to the code that comes from the increased diversity of proteins that can be made with an increased amino acid repertoire. Nevertheless, several amino acids are found to be as frequent in meteorites as the early biological ones and yet did not become incorporated into the code (Lu and Freeland, 2006; Wiltschi and Budisa, 2007). Many of the four-codon boxes are split between two amino acids. If this were done in every four- codon box, there could be up to 32 amino acids specified. On the other hand, if we accept the two-out-of-three argument, then there are 8 boxes which cannot be split. This reduces the maximum number of amino acids to 24 (8 four-codon families and 16 two-codon families). Splitting a two codon family into separate single codons is possible in the case of Met and Trp codons, but this does not increase the number of amino acids encoded in either case, and in many mitochondrial codes, Met and Trp both have two codons (Sengupta et al. 2007). None of these arguments explains why there should be some amino acids with 6 codons. Thus, the fact that the observed 20 amino acids is fewer than the maximum seems to require some explanation. It may be that if an amino acid is too chemically similar to others that are already in the code, there is little selective advantage to adding it. Furthermore, adding more amino acids requires reducing the size of the codon blocks. Hence the probability of translational error becomes larger. As the number of amino acids in the code increases, the advantage of adding a further amino acid becomes small relative to the disadvantage of increased translational error rate (Higgs, 2009). This would explain why the process of code evolution stopped with an intermediate number of amino acids incorporated. For some of the non-biological amino acids that are seen in meteorites and prebiotic synthesis experiments, chemical reasons have been proposed as to why they were avoided. For example, Weber and Miller (1981) argue that amino acids without an α hydrogen (such as α-amino isobutyric acid and isovaline) are avoided because there would be increased steric hindrance and lack of flexibility of the backbone. However, they were unable to propose a reason for avoidance of α-amino-n-butyric acid, norvaline and norleucine. These are similar in properties to valine, leucine and isoleucine, so it may simply be that the existing set of 20 gives a fairly complete coverage of physical property space already. 11 In addition to the role of physical properties and of precursor product relationships, another factor that has also been proposed to influence the initial codon assignments is that there may have been direct interactions between RNA triplets and amino acids (Yarus et al. 2005). Despite their differences, all these arguments agree that early proteins were composed of a small set of amino acids. Selection would favour the addition of new amino acids because the number of possible protein sequences increases tremendously with each amino acid in the repertoire. However, proteins composed of a restricted set of amino acids must be functional at each step of this process in order for selection to favour the use of proteins. Several studies have shown that sequences composed of small sets of amino acids can have structures and functions similar to those of 20-amino acid proteins (Davidson et al. 1995; Riddle et al. 1997; Babajide et al. 1997; Doi et al. 2005). Thus selection may favour the increasing use of more diverse and better adapted proteins and hence the RNA world would be eventually overtaken by the protein-dominated world. In summary, we have shown that in spite of the conflicting opinions on the mechanisms and locations of molecular synthesis, thermodynamics cuts across these distinctions and predicts which amino acids are formed most easily. Our results also indicate that a certain degree of universality would be expected in the types of organic molecules seen on other earth- like planets. Should life exist elsewhere, it would not be surprising if it used at least some of the same amino acids we do. Simple sugars, lipids and nucleobases might also be shared. Our analysis suggests that the first genetic code was a stripped down version of our present code, one whose simpler structure reflected the limited set of available amino acids in the prebiotic environment. Although there are countless ways that the code could have developed from those origins, the combined actions of thermodynamics and subsequent natural selection suggest that the genetic code we observe on the Earth today may have significant features in common with life throughout the cosmos. Acknowledgements This work is supported by Canada Research Chairs and the Natural Sciences and Engineering Research Council of Canada. We thank Chris McKay for useful suggestions on this work, and two anonymous referees, whose comments were extremely helpful. 12 References Akashi, H., and Gojobori, T. (2002) Metabolic efficiency and amino acid composition in the proteomes of Escherichia coli and Bacillus subtilis. Proc. Nat. Acad. Sci. USA 99:3695-3700. Amend, J.P., and Shock, E.L. (1998) Energetics of Amino Acid Synthesis in Hydrothermal Ecosystems. Science 281:1659-1662. Babajide, A., Hofacker, I.L., Sippl, M.J., and Stadler, P.F. (1997) Neutral networks in protein space: a computational study based on knowledge potentials of mean force. Folding and Design 2:261-269. Bada, J.L., Miller, S.L., and Zhao M. (1995) The stability of amino acids at submarine hydrothermal vent temperatures. Orig Life Evol Biosph. 25:111-8. Bar-Nun, A., Bar-Nun, N., Bauer, S.H., and Sagan, C. (1970) Shock synthesis of amino acids in simulated primitive environments. Science 168:470-473. Baross, J.A., and Hoffman, S.E. (1985) Submarine Hydrothermal Vents and Associated Gradient Environments as Sites for the Origin and Evolution of Life. Origins Life 15:327-345. Berezovsky, I.N., Chen, W.W., Choi, P.J. and Shakhnovitch, E.I. (2005) Entropic stabilization of proteins and its proteomic consequences. PLoS Comput. Biol. 1(4)e47. Bergin, E.A., Aikawa, Y., Blake, G.A., and van Dishoeck, E.F. (2007) The chemical composition of protoplanetary disks. In Protostars and Planets V, edited by B. Reipurth, D. Jewitt, and K. Keil. University of Arizona Press, Tucson. Bernstein, M.P., Dworkin, J.P., Sandford, S.A., Cooper, G.W., and Allamandola, L.J. (2002) Racemic amino acids from the ultraviolet photolysis of interstellar ice analogues. Nature 416:401-403. Bernstein, M.P. (2006) Prebiotic materials from on and off the early Earth. Phil. Trans. R. Soc. B 361:1689-1702. Chakrabarti, S., and Chakrabarti, S.K. (2000) Can DNA bases be produced during molecular cloud collapse? Astron. Astrophys. 354:L6-L8. Cleaves, H.J., Chalmers, J.H., Lazcano, A., Miller, S. L., and Bada, J.L. (2008) A reassessment of prebiotic organic synthesis in neutral atmospheres. Orig. Life Evol. Biosph. 38:105-115. Crick, F.H.C. (1968) The origin of the genetic code. J. Mol. Biol. 38:367-379. Cronin, C.R., and Moore, C.B. (1971) Amino acid analyses of the Murchison, Murray and Allende carbonaceous chondrites. Science 172:1327-1329. Davidson, A.R., Lumb, K.L., and Sauer, R.T. (1995) Cooperatively folded proteins in random sequence libraries. Nature Struct. Biol. 2:856-864. 13 Di Giulio, M., and Medugno, M. (1999) Physicochemical optimization in the genetic code origin as the number of codified amino acids increases. J. Mol. Evol. 49:1-10. Di Giulio, M. (2005) The origin of the genetic code: theories and their relationships, a review. BioSystems 80:175-184. Doi, N., Kakukawa, K., Oishi, Y., and Yanagawa, H. (2005) High solubility of random sequence proteins consisting of five kinds of primitive amino acids. Prot. Eng. Design and Selection 18:279-284. Engel, M.H., and Nagy, B. (1982) Distribution and enantiomeric composition of amino acids in the Murchison meteorite. Nature 296:837-840. Freeland, S.J., Wu, T., and Keulmann, N. (2003) The case for an error-minimizing standard genetic code. Origins Life Evol. Biosph. 33:457-477. Hennet, R.J.C., Holm, N.G., and Engel, M.H. (1992) Abiotic synthesis of amino acids under hydrothermal conditions and the origin of life: a perpetual phenomenon. Naturwissenschaften 79:361-365. Higgs, P.G., and Pudritz, R.E. (2007) From protoplanetary disks to prebiotic amino acids and the origin of the genetic code. In Planetary systems and the origins of life, Cambridge Series in Astrobiology, vol 3. edited by R.E. Pudritz, P.G. Higgs and J. Stone. Cambridge Uni. Press. pp62-88. Higgs, P.G. (2009) A Four-Column Theory for the Origin of the Genetic Code: Tracing the Evolutionary Pathways that Gave Rise to an Optimized Code. Biology Direct (in press). Joyce, G.F. (2002) The antiquity of RNA-based evolution. Nature 418:214-221. Jungck, J.R. (1978) The genetic code as a periodic table. J. Mol. Evol. 11:211-224. Kasting, J.F. (1993) Earth’s early atmosphere. Science 259:920-926. Knight, R.J, Freeland, S.J., and Landweber, L.F. (2001) Rewiring the keyboard: evolvability of the genetic code. Nature Rev. Genet. 2:49–58. Lagerkvist, U. (1978) “Two out of three”: an alternative method for codon reading. Proc. Acad. Nat. Sci. USA 75:1759-1762. Lagerkvist, U. (1986) Unconventional methods in codon reading. Bioessays 4:223-226. Lehmann, J., and Libchaber, A. (2008) Degeneracy of the genetic code and stability of the base pair at the second position. RNA 14:1264-1269. Lowe, C.U., Rees, M.W., and Markham, R. (1963) Synthesis of complex organic compounds from simple precursors: formation of amino acids, amino-acid polymers, fatty acids and purines from ammonium cyanide. Nature 199:219-222. 14 Lu, Y., and Freeland, S. (2006) On the evolution of the standard amino-acid alphabet. Genome Biol. 7:102. Marshall, W.L. (1994) Hydrothermal synthesis of amino acids. Geochim. Cosmochim. Acta 58:2099-2106. Martins, Z., Alexander, C.M.O’D., Orzechowska, G.E., Fogel, M.L., and Ehrenfreund, P. (2007) Indigenous amino acids in primitive CR meteorites. Meteorics Planet. Sci. 42:2125- 2136. Miller, S.L. (1953) Production of amino acids under possible primitive Earth conditions. Science 117:528-529. Miller, S.L. (1957) The mechanism of synthesis of amino acids by electric discharges. Biochim. et Biophys. Acta 23:480-489. Miller, S.L., and Orgel, L.E. (1974) The Origins of Life on the Earth. Prentice-Hall, New Jersey. Miller, S.L., and Bada, J.L. (1988) Submarine hot springs and the origin of life. Nature 334:609-611. Miyakawa, S., Yamanashi, H., Kobayashi, K., Cleaves, H.J., and Miller, S.L. (2002) Prebiotic synthesis from CO atmospheres: Implications for the origins of life. Proc. Nat. Acad. Sci. USA 99:14628-14631. Munoz Caro, G.M., Meierhenrich, U.J., Schutte, W.A., Barbier, B., Arcones Segovia, A., Rosenbauer, H., Thiemann, W.H.P., Brack, A., and Greenberg, J.M. (2002) Amino acids of ultra-violet irradiation of interstellar ice analogues. Nature 416:403-406. Pierazzo, E., and Chyba, C.F. (1999) Amino acid survival in large cometary impacts. Meteoritics Planet. Sci. 34:909-918. Pizzarello, S., Huang, Y., and Fuller, M. (2004) The carbon isotopic distribution of Murchison amino acids. Geochim. Cosmochim. Acta 68:4963-4969. Riddle, D.S., Santiago, J.V., Bray-Hall, S.T., Doshi, N., Grantcharova, V.P., Yi, Q., and Baker, D. (1997) Functional rapidly-folding proteins from simplified amino acid sequences. Nature Struct. Biol. 4:805-809. Schaefer, L., and Fegley, B. Jr, (2007) Outgassing of ordinary chondritic material and some of its implications for the chemistry of asteroids, planets and satellites. Icarus 186:462-483. Schlesinger, G., and Miller, S.L. (1983) Prebiotic synthesis in atmospheres containing CH4, CO and CO2. J. Mol. Evol. 19:376-382. Seligman, H. (2003) Cost-minimization of amino acid usage. J. Mol. Evol. 56:151-161. 15 Sengupta, S., Yang, X., and Higgs, P.G. (2007) The mechanisms of codon reassignments in mitochondrial genetic codes. J. Mol. Evol. 64:662-688. Shimoyama, A., Ponnamperuma, C., and Yanai, K. (1979) Amino acids in the Yamato carbonaceous chondrite from Antarctica. Nature 282:394-396. Swire, J. (2007) Selection on Synthesis Cost Affects Interprotein Amino Acid Usage in All Three Domains of Life. J. Mol. Evol. 64:558-571. Taylor, F.J.R., and Coates, D. (1989) The code within the codons. Biosystems 22:177-187. Tian, F., Toon, O.B., Pavlov, A.A., and De Sterck, H. (2005) A Hydrogen-Rich Early Earth Atmosphere. Science 308, 1014-1017. Trifonov, E.N. (2004) The triplet code from first principles. J. Biomol. Struct. Dynam. 22:1-11. Urbina, D., Tang, B., and Higgs, P.G. (2006) The response of amino acid frequencies to directional mutation pressure in mitochondrial genome sequences is related to the physical properties of the amino acids and the structure of the genetic code. J. Mol. Evol. 62:340-361. Weber, A.L., and Miller, S.L. (1981) Reasons for the occurrence of the twenty coded protein amino acids. J. Mol. Evol. 17:273-284. Whittet, D.C.B. (1997) Is extraterrestrial organic matter relevant to the origin of life on Earth? Orig. Life Evol. Biosph. 27:249-262. Wiltschi, B., and Budisa, N. (2007) Natural history and experimental evolution of the genetic code. Appl. Microbiol. Biotechnol. 74:739-753. Wolman, Y., Haverland, W.J., and Miller, S.L. (1972) Nonprotein amino acids from spark discharges and their comparison with the Murchison meteorite amino acids. Proc. Nat. Acad. Sci. USA 69:809-811. Wong, J.T. (1975) A coevolution theory of the genetic code. Proc. Nat. Acad. Sci. USA 72:1909-1912. Wong, J.T. (2005) Coevolution theory of the genetic code at age thirty. BioEssays 27:416-425. Yarus, M., Caporaso, J.G., and Knight, R. (2005) Origins of the genetic code: the escaped triplet theory. Annu. Rev. Biochem. 74:179-198. Yoshino, D., Hayatso, R., and Anders, E. (1971) Origin of organic matter in the early solar system – III. Amino acids: catalytic synthesis. Geochim. Cosmochim. Acta 35:927-938. 16 Table 1 Relative concentrations of amino acids observed in non-biological contexts. With the exception of columns H1, H2 and S3, the quoted figures are concentrations normalized such that Gly = 1.0. For columns H1, H2 and S3, the figures are the number of experiments in which the amino acid was observed in greater than trace amounts. M denotes meteorites, I denotes icy grains, A denotes atmospheric synthesis, H denotes hydrothermal synthesis, and S denotes other chemical syntheses (details of sources are given in the text). Robs is the mean rank derived from these observations. The final 6 amino acids are not observed S1 S2 1.000 1.000 0.473 0.097 0.000 0.581 0.000 0.000 0.006 0.000 0.000 0.154 0.000 0.002 0.001 0.002 0.000 0.000 0.000 0.002 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 S3 40 20 30 20 2 0 4 7 2 1 14 1 15 15 0 Robs 1.1 2.8 4.3 6.8 8.5 8.6 9.1 9.4 10.0 11.7 12.6 13.2 13.3 13.3 14.2 M1 M2 1.00 1.0 G - Gly 0.34 0.4 A - Ala 0.19 0.5 D - Asp 0.40 0.5 E - Glu 0.19 0.3 V - Val 0.00 0.0 S - Ser 0.13 0.0 I - Ile 0.04 0.0 L - Leu 0.29 0.1 P - Pro 0.00 0.0 T - Thr 0.00 0.0 K - Lys 0.00 0.0 F - Phe 0.00 0.0 R - Arg H - His 0.00 0.0 NQCYMW 0.00 0.0 M3 H1 A3 A2 A1 I1 1.000 1.000 1.000 1.000 1.000 18 0.380 0.293 0.540 1.795 0.155 15 0.035 0.022 0.006 0.077 0.059 10 0.110 0.000 0.010 0.018 0.000 6 0.100 0.012 0.000 0.044 0.000 1 0.003 0.072 0.000 0.011 0.018 8 0.060 0.000 0.000 0.011 0.000 8 0.035 0.000 0.000 0.026 0.000 3 0.000 0.001 0.000 0.003 0.000 9 0.003 0.000 0.000 0.002 0.000 2 0.000 0.000 0.000 0.000 0.000 7 0.000 0.000 0.000 0.000 0.000 4 0.000 0.000 0.000 0.000 0.000 0 0.000 0.000 0.000 0.000 0.000 0 0.000 0.000 0.000 0.000 0.000 0 H2 12 8 10 11 0 11 9 0 0 0 0 0 0 0 0 17 Table 2 Thermodynamic and evolutionary properties of amino acids. Robs = ranking derived from Table 1; Rcode = ranking from Trifonov (2004); Crel = mean relative concentration from Table 1; ΔGsurf and ΔGhydro = free energies of formation in kJ/mol (Amend and Shock, 1998); ATP cost of synthesis (Akashi and Gojobori, 2002); MW = molecular weight in Da. ΔGhydro ΔGsurf 14.89 80.49 113.66 -12.12 146.74 32.78 -1.43 172.13 178.00 -70.12 173.73 69.47 213.93 -96.4 -105.53 205.03 192.83 -38.75 216.50 53.51 258.56 -28.33 303.64 -114.54 409.46 197.52 350.52 154.48 201.56 83.53 223.36 44.04 224.67 60.24 -59.53 334.20 113.22 -174.71 -38.99 431.17 ATP MW 75.1 11.7 89.1 11.7 12.7 133.1 147.1 15.3 117.1 23.3 105.1 11.7 32.3 131.2 131.2 27.3 115.1 20.3 119.1 18.7 30.3 146.2 165.2 52.0 174.2 27.3 155.2 38.3 14.7 132.1 146.1 16.3 121.2 24.7 181.2 50.0 34.3 149.2 204.2 74.3 18 R G - Gly A - Ala D - Asp E - Glu V - Val S - Ser I - Ile L - Leu P - Pro T - Thr K - Lys F - Phe R - Arg H - His N - Asn Q - Gln C - Cys Y - Tyr M - Met W - Trp obs Rcode 3.5 1.1 4.0 2.8 4.3 6.0 8.1 6.8 6.3 8.5 7.6 8.6 9.1 11.4 9.4 9.9 10.0 7.3 11.7 9.4 12.6 13.3 13.2 14.4 13.3 11.0 13.3 13.0 14.2 11.3 14.2 11.4 14.2 13.8 14.2 15.2 14.2 15.4 14.2 16.5 Crel 1.0000 0.4970 0.1633 0.1153 0.0724 0.0286 0.0226 0.0116 0.0437 0.0008 0 0 0 0 0 0 0 0 0 0 Figure Captions. Figure 1 – The ranking Robs for the early amino acids (circles) correlates well with the free energy of formation in surface seawater conditions, ΔGsurf. The late amino acids (triangles) have high ΔGsurf. Figure 2 – The mean relative concentration Crel for the early amino acids decreases approximately exponentially with ΔGsurf. Figure 3 – There is no correlation between the ranking Robs and the free energy of formation in hydrothermal conditions, ΔGhydro, for either the early or late amino acids. Figure 4 – The table of assignments between codons and amino acids in the standard genetic code. 19 Fig 1 Robs 0 2 4 6 8 10 12 14 16 0 G A D S E V P I L T K F H R N Q Y W 100 200 300 400 Δ Gsurf 500 20 Fig 2 1 G 0.1 Crel 0.01 0.001 A D E V S P L I T 80 100 120 140 160 ΔG surf 180 200 220 240 21 Fig 3 Robs 0 2 4 6 8 10 12 14 G D A E V I L F Y P K W S T Q 16 -150 -100 -50 0 Δ 50 G hydro 22 N 100 H R 150 200 Fig 4 23
1010.1171
1
1010
2010-10-06T15:06:38
Polarimetric Remote Sensing of Solar System Objects
[ "astro-ph.EP", "physics.ao-ph", "physics.ins-det", "physics.optics" ]
This book outlines the basic physical principles and practical methods of polarimetric remote sensing of Solar System objects and summarizes numerous advanced applications of polarimetry in geophysics and planetary astrophysics. In the first chapter we present a complete and rigorous theory of electromagnetic scattering by disperse media directly based on the Maxwell equations and describe advanced physically based modeling tools. This is followed, in Chapter 2, by a theoretical analysis of polarimetry as a remote-sensing tool and an outline of basic principles of polarimetric measurements and their practical implementations. In Chapters 3 and 4, we describe the results of extensive ground-based, aircraft, and spacecraft observations of numerous Solar System objects (the Earth and other planets, planetary satellites, Saturn's rings, asteroids, trans-Neptunian objects, and comets). Theoretical analyses of these data are used to retrieve optical and physical characteristics of planetary surfaces and atmospheres as well as to identify a number of new phenomena and effects. This monograph is intended for science professionals, educators, and graduate students specializing in remote sensing, astrophysics, atmospheric physics, optics of disperse and disordered media, and optical particle characterization.
astro-ph.EP
astro-ph
Winner of the Basic Sciences Book Award NATIONAL ACADEMY OF SCIENCES OF UKRAINE MAIN ASTRONOMICAL OBSERVATORY NATIONAL AERONAUTICS AND SPACE ADMINISTRATION OF THE USA GODDARD INSTITUTE FOR SPACE STUDIES __________________________________________________ ɇȺɐȱɈɇȺɅɖɇȺ ȺɄȺȾȿɆȱə ɇȺɍɄ ɍɄɊȺȲɇɂ ȽɈɅɈȼɇȺ ȺɋɌɊɈɇɈɆȱɑɇȺ ɈȻɋȿɊȼȺɌɈɊȱə ɇȺɐȱɈɇȺɅɖɇȿ ɍɉɊȺȼɅȱɇɇə ȺȿɊɈɇȺȼɌɂɄɂ ɌȺ ȾɈɋɅȱȾɀȿɇɇə ɄɈɋɆȱɑɇɈȽɈ ɉɊɈɋɌɈɊɍ ɋɒȺ ʈɈȾȾȺɊȾȱȼɋɖɄɂɃ ȱɇɋɌɂɌɍɌ ɄɈɋɆȱɑɇɂɏ ȾɈɋɅȱȾɀȿɇɖ ʛ. I. ʛ˪, ʑ. K. ʟ, ʛ. ʛ. ʙ, ʓ. ʣ. ʚ˪, ʑ. ʞ. ʡ, ʑ. ʒ. ʙ, ʶ. ʛ. ʐ, ʭ. ʠ. ʲ˪, ʛ. ʛ. ʧ ʓʗʠʡʏʜʥʶʘʜʔ ʖʝʜʓʢʑʏʜʜʮ ʝʐ’ʲʙʡʶʑ ʠʝʜʮʦʜʝʷ ʠʗʠʡʔʛʗ ʞʝʚʮʟʗʛʔʡʟʗʦʜʗʛʗ ʖʏʠʝʐʏʛʗ _______________________________ ɉɊɈȿɄɌ «ɍɄɊȺȲɇɋɖɄȺ ɇȺɍɄɈȼȺ ɄɇɂȽȺ ȱɇɈɁȿɆɇɈɘ ɆɈȼɈɘ» _______________________________ ɄɂȲȼ • ȺɄȺȾȿɆɉȿɊȱɈȾɂɄȺ • 2010 M. I. Mishchenko, V. K. Rosenbush, N. N. Kiselev, D. F. Lupishko, V. P. Tishkovets, V. G. Kaydash, I. N. Belskaya, Y. S. Efimov, N. M. Shakhovskoy POLARIMETRIC REMOTE SENSING OF SOLAR SYSTEM OBJECTS _____________________________ PROJECT «UKRAINIAN SCIENTIFIC BOOK IN A FOREIGN LANGUAGE» _____________________________ KYIV • AKADEMPERIODYKA • 2010 UDK 520.85:523.4 BBK 22.63 P80 Mishchenko M. I., Rosenbush V. K., Kiselev N. N., Lupishko D. F., Tishkovets V. P., Kaydash V. G., Belskaya I. N., Efimov Y. S., Shakhovskoy N. M. Polarimetric remote sensing of Solar System objects. − K.: Akademperiodyka, 2010. − 291 p., 24 p. il. This book outlines the basic physical principles and practical methods of polarimetric remote sensing of Solar System objects and summarizes numerous advanced applications of polarimetry in geophysics and planetary astrophysics. In the first chapter we present a com- plete and rigorous theory of electromagnetic scattering by disperse media directly based on the Maxwell equations and describe advanced physically based modeling tools. This is fol- lowed, in Chapter 2, by a theoretical analysis of polarimetry as a remote-sensing tool and an outline of basic principles of polarimetric measurements and their practical implementa- tions. In Chapters 3 and 4, we describe the results of extensive ground-based, aircraft, and spacecraft observations of numerous Solar System objects (the Earth and other planets, planetary satellites, Saturn’s rings, asteroids, trans-Neptunian objects, and comets). Theo- retical analyses of these data are used to retrieve optical and physical characteristics of planetary surfaces and atmospheres as well as to identify a number of new phenomena and effects. This monograph is intended for science professionals, educators, and graduate students specializing in remote sensing, astrophysics, atmospheric physics, optics of disperse and dis- ordered media, and optical particle characterization. Ʉɧɢɝɚ ɦɿɫɬɢɬɶ ɨɫɧɨɜɧɿ ɮɿɡɢɱɧɿ ɩɪɢɧɰɢɩɢ ɿ ɩɪɚɤɬɢɱɧɿ ɦɟɬɨɞɢ ɩɨɥɹɪɢɦɟɬɪɢɱɧɨɝɨ ɞɢɫɬɚɧɰɿɣɧɨɝɨ ɡɨɧɞɭɜɚɧɧɹ ɨɛ’ɽɤɬɿɜ ɋɨɧɹɱɧɨʀ ɫɢɫɬɟɦɢ ɬɚ ɩɿɞɫɭɦɨɜɭɽ ʀɯɧɽ ɡɚɫɬɨɫɭɜɚɧɧɹ ɜ ɝɟɨɮɿɡɢɰɿ ɣ ɩɥɚɧɟɬɧɿɣ ɚɫɬɪɨɮɿɡɢɰɿ. ȼ ɩɟɪɲɨɦɭ ɪɨɡɞɿɥɿ ɤɧɢɝɢ ɩɪɟɞɫɬɚɜɥɟɧɨ ɡɚɜɟɪɲɟ- ɧɭ ɫɬɪɨɝɭ ɬɟɨɪɿɸ ɪɨɡɫɿɹɧɧɹ ɟɥɟɤɬɪɨɦɚɝɧɿɬɧɢɯ ɯɜɢɥɶ ɞɢɫɩɟɪɫɧɢɦɢ ɫɟɪɟɞɨɜɢɳɚɦɢ ɧɚ ɨɫɧɨɜɿ ɪɿɜɧɹɧɶ Ɇɚɤɫɜɟɥɚ ɣ ɨɩɢɫɚɧɨ ɫɭɱɚɫɧɿ ɮɿɡɢɱɧɨ ɨɛʉɪɭɧɬɨɜɚɧɿ ɦɟɬɨɞɢ ɬɟɨɪɟɬɢɱɧɨɝɨ ɦɨɞɟɥɸɜɚɧɧɹ. ȼ ɞɪɭɝɨɦɭ ɪɨɡɞɿɥɿ ɩɪɟɞɫɬɚɜɥɟɧɨ ɬɟɨɪɟɬɢɱɧɢɣ ɚɧɚɥɿɡ ɩɨɥɹɪɢɦɟɬɪɿʀ ɹɤ ɦɟ- ɬɨɞɭ ɞɢɫɬɚɧɰɿɣɧɨɝɨ ɡɨɧɞɭɜɚɧɧɹ, ɚ ɬɚɤɨɠ ɨɩɢɫɚɧɨ ɬɟɨɪɟɬɢɱɧɿ ɨɫɧɨɜɢ ɿ ɩɪɢɧɰɢɩɢ ɜɢɦɿ- ɪɸɜɚɧɧɹ ɩɨɥɹɪɢɡɨɜɚɧɨɝɨ ɜɢɩɪɨɦɿɧɸɜɚɧɧɹ, ʉɪɭɧɬɭɸɱɢɫɶ ɧɚ ɹɤɢɯ ɫɬɜɨɪɟɧɨ ɭɧɿɤɚɥɶɧɭ ɩɪɟɰɢɡɿɣɧɭ ɚɩɚɪɚɬɭɪɭ ɞɥɹ ɫɩɨɫɬɟɪɟɠɟɧɶ. ȼ ɬɪɟɬɶɨɦɭ ɬɚ ɱɟɬɜɟɪɬɨɦɭ ɪɨɡɞɿɥɚɯ ɩɪɨɜɟɞɟ- ɧɨ ɚɧɚɥɿɡ ɜɟɥɢɤɨɝɨ ɨɛɫɹɝɭ ɧɚɡɟɦɧɢɯ ɬɚ ɚɟɪɨɤɨɫɦɿɱɧɢɯ ɫɩɨɫɬɟɪɟɠɟɧɶ, ɧɚ ɨɫɧɨɜɿ ɹɤɨɝɨ ɜɢɡɧɚɱɟɧɨ ɨɩɬɢɱɧɿ ɬɚ ɮɿɡɢɱɧɿ ɯɚɪɚɤɬɟɪɢɫɬɢɤɢ ɩɨɜɟɪɯɨɧɶ ɿ ɚɬɦɨɫɮɟɪ ɛɚɝɚɬɶɨɯ ɬɿɥ ɋɨ- ɧɹɱɧɨʀ ɫɢɫɬɟɦɢ (Ɂɟɦɥɹ ɬɚ ɿɧɲɿ ɩɥɚɧɟɬɢ, ɫɭɩɭɬɧɢɤɢ ɩɥɚɧɟɬ, ɤɿɥɶɰɹ ɋɚɬɭɪɧɚ, ɚɫɬɟɪɨʀɞɢ, ɬɪɚɧɫɧɟɩɬɭɧɨɜɿ ɨɛ’ɽɤɬɢ ɣ ɤɨɦɟɬɢ) ɬɚ ɜɿɞɤɪɢɬɨ ɰɿɥɢɣ ɪɹɞ ɧɨɜɢɯ ɹɜɢɳ ɿ ɟɮɟɤɬɿɜ. ɉɪɢɡɧɚɱɟɧɨ ɞɥɹ ɧɚɭɤɨɜɰɿɜ, ɜɢɤɥɚɞɚɱɿɜ, ɚɫɩɿɪɚɧɬɿɜ ɬɚ ɫɬɭɞɟɧɬɿɜ, ɹɤɿ ɫɩɟɰɿɚɥɿɡɭ- ɸɬɶɫɹ ɜ ɞɢɫɬɚɧɰɿɣɧɨɦɭ ɡɨɧɞɭɜɚɧɧɿ, ɚɫɬɪɨɮɿɡɢɰɿ, ɚɬɦɨɫɮɟɪɧɿɣ ɮɿɡɢɰɿ, ɨɩɬɢɰɿ ɞɢɫɩɟɪɫ- ɧɢɯ ɿ ɜɢɩɚɞɤɨɜɢɯ ɫɟɪɟɞɨɜɢɳ ɬɚ ɨɩɬɢɱɧɿɣ ɞɿɚɝɧɨɫɬɢɰɿ ɱɚɫɬɢɧɨɤ. Approved for publication by the Scientific Council of the Main Astronomical Observatory of the National Academy of Sciences of Ukraine ¤ I. N. Belskaya, Y. S. Efimov, V. G. Kaydash, N. N. Kiselev, D. F. Lupishko, M. I. Mishchenko, V. K. Rosenbush, N. M. Shakhovskoy, V. P. Tishkovets, 2010 ISBN 978-966-360-134-2 Foreword to the project «Ukrainian Scientific Book in a Foreign Language» It is a great pleasure to introduce to the Ukrainian and international scientific communities the first volume of a new series of scientific monographs initiated by the Publications Council of the National Academy of Sciences of Ukraine (NASU). Although the literal English translation of the title of this series (“Ukrainian Scien- tific Book in a Foreign Language”) may sound a bit inelegant to a sophisticated English speaker, it makes perfect sense from the perspective of what this series is intended to accomplish. Indeed, the main idea is to provide as broad an interna- tional access to major accomplishments of leading Ukrainian scientists as possible. Of course, the “foreign language” will in most cases be the English language. Being a professional astronomer, I am especially pleased that the subject of this first volume is astrophysics of the Solar System or, more specifically, remote sens- ing of the Solar System using polarimetric techniques. The book is authored by eight Ukrainian astrophysicists and an American scientist of Ukrainian origin, all of whom are internationally recognized experts in this area of research. They rep- resent four astronomical organizations of Ukraine (the Main Astronomical Obser- vatory of NASU, the Institute of Astronomy of the Kharkiv V. N. Karazin National University, the Crimean Astrophysical Observatory, and the Institute of Radio- astronomy of NASU) as well as the Goddard Institute for Space Studies of the Na- tional Aeronautics and Space Administration of the USA. This book is the result of multi-decadal theoretical and experimental investiga- tions which have contributed profoundly to the solution of a wide range of funda- mental and applied problems of science: from the origin and evolution of the Solar System to the detection and characterization of potentially hazardous asteroids, monitoring of global climate changes on Earth, and ecological control of the ter- restrial atmosphere. I am confident that the publication of this monograph will serve to summarize the current status of polarimetric remote sensing of the Solar System as well as to facilitate further progress in this important scientific discipline. Yaroslav S. Yatskiv Head of the Scientific Publishing Council of the National Academy of Sciences of Ukraine Kyiv To our families and friends Contents EDITORIAL . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11 ACKNOWLEDGMENTS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14 INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15 1. Electromagnetic scattering by discrete random media: unified microphysical theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17 1.1. Basic assumptions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17 1.2. The macroscopic Maxwell equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19 1.3. Electromagnetic scattering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20 1.4. Far-field and near-field scattering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23 1.5. Actual observables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25 1.6. Derivative far-field characteristics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29 1.7. Foldy–Lax equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30 1.8. Multiple scattering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31 1.9. Far-field Foldy–Lax equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33 1.10. Ergodicity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36 1.11. Scattering by a small random particle group . . . . . . . . . . . . . . . . . . . . . . . . . . 37 1.12. Lorenz–Mie theory and T-matrix method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38 1.13. Spherically symmetric particles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43 1.14. General effects of nonsphericity and orientation . . . . . . . . . . . . . . . . . . . . . . 44 1.15. Mirror-symmetric ensembles of randomly oriented particles: general traits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45 1.16. Mirror-symmetric ensembles of randomly oriented particles: quantitative traits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47 1.17. Multiple scattering by random particulate media: exact results . . . . . . . 50 1.17.1. Static and dynamic light scattering . . . . . . . . . . . . . . . . . . . . . . . . . . . 51 1.17.2. Fixed configurations of randomly positioned particles: speckle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51 1.17.3. Static scattering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52 1.18. Radiative transfer theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56 1.19. Weak localization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66 1.20. Forward-scattering interference . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69 1.21. “Independent” scattering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69 1.22. Radiative transfer in gaseous media . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70 8 Contents 1.23. Energy conservation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71 1.24. Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72 1.25. Weak localization in plane-parallel discrete random media . . . . . . . . . . . 75 1.25.1. Exact backscattering direction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76 1.25.2. Angular dependence of weak localization: Rayleigh scattering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79 1.25.3. Angular dependence of weak localization: wavelength-sized scatterers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81 1.26. Near-field effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88 1.27. Scattering of quasi-monochromatic light . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94 2. Theoretical basis, methods, and hardware implementations of polarimetric remote sensing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95 2.1. Theoretical foundations of polarimetry as an efficient means of remote characterization of terrestrial and celestial objects . . . . . . . . . . . . 95 2.1.1. Retrieval requirements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97 2.1.2. Hierarchy of passive remote-sensing instruments . . . . . . . . . . . . . 99 2.1.3. Sensitivity assessment of polarimetry as a remote-sensing tool . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101 2.2. Basic concepts and principles of polarization measurements . . . . . . . . . 103 2.3. Development of observational polarimetry and polarimetric instrumentation in Ukraine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106 2.3.1. Aperture polarimeters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107 2.3.2. Panoramic polarimetry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112 2.4. Passive photopolarimeter for remote sensing of terrestrial aerosols and clouds from orbital satellites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114 2.5. Methodology of polarimetric observations . . . . . . . . . . . . . . . . . . . . . . . . . . . 117 2.5.1. Analysis of random and systematic errors . . . . . . . . . . . . . . . . . . . 117 2.5.2. Reduction of polarimetric data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120 2.5.3. Statistical analysis of polarimetric observations . . . . . . . . . . . . . 120 2.5.4. Methods of polarimetric observations of small Solar System bodies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122 2.6. Multicomponent superachromatic retarders: theory and implementation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124 2.6.1. Extension of the Pancharatnam system to an arbitrary number of components . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124 2.6.2. Practical implementation of the theory of multicomponent superachromatic retarders . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128 3. Photopolarimetric observations of atmosphereless Solar System bodies and planets and their interpretation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130 3.1. Planets and satellites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130 Contents 9 3.1.1. Terrestrial aerosols . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130 3.1.2. Imaging polarimetry of Mars with the Hubble Space Telescope . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136 3.1.3. Polarimetry of Mercury . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141 3.1.4. Optical properties of the lunar surface . . . . . . . . . . . . . . . . . . . . . . . 143 3.1.5. Phase, longitudinal, and spectral dependences of polarization for the Galilean satellites of Jupiter and Iapetus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148 3.1.6. Opposition effects for planetary satellites . . . . . . . . . . . . . . . . . . . . 160 3.1.7. Relation between observed optical effects and physical processes controlling the formation of the surface layer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161 3.2. Asteroids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163 3.2.1. Observations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163 3.2.2. Phase-angle dependence of polarization . . . . . . . . . . . . . . . . . . . . . 164 3.2.3. Spectral dependence of polarization . . . . . . . . . . . . . . . . . . . . . . . . . 166 3.2.4. Polarimetry of near-Earth asteroids . . . . . . . . . . . . . . . . . . . . . . . . . 168 3.2.5. Identification and study of asteroids with anomalous polarimetric characteristics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 172 3.2.6. Polarimetry of asteroids selected as targets of space missions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 175 3.2.7. Opposition effects of asteroids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 179 3.2.8. Determination of albedos and sizes of asteroids on the basis of polarimetric and photometric observations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 183 3.2.9. Structure of asteroid surfaces according to polarimetric data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 186 3.2.10. Asteroid composition according to polarimetric data . . . . . . . . . 188 3.2.11. Asteroid Polarimetric Database . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 189 3.3. Trans-Neptunian objects and Centaurs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 191 3.3.1. Description and results of polarimetric observations . . . . . . . . 191 3.3.2. Surface structure and composition . . . . . . . . . . . . . . . . . . . . . . . . . . . 195 3.4. Opposition phenomena exhibited by Solar System objects: a general perspective . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 198 4. Polarimetric observations of comets and their interpretation. . . . . . . . . . . . 204 4.1. Historical background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 204 4.2. Observational data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 205 4.3. Phase-angle dependence of linear polarization . . . . . . . . . . . . . . . . . . . . . . . 209 4.3.1. Negative polarization branch . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 209 4.3.2. Positive polarization branch . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 210 4.3.3. Causes of observed polarimetric differences . . . . . . . . . . . . . . . . . 212 10 Contents 4.4. Similarity and diversity of comets: classification issues . . . . . . . . . . . . . . 215 4.5. Spectral dependence of polarization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 217 4.6. Polarimetry of comets during stellar occultations . . . . . . . . . . . . . . . . . . . . . . 219 4.7. Systematic deviations of polarization parameters from “nominal” values . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 221 4.8. Circular polarization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 222 4.8.1. Measurement results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 222 4.8.2. Optical effects causing circular polarization . . . . . . . . . . . . . . . . . 225 4.9. Effects of non-steady processes on polarization . . . . . . . . . . . . . . . . . . . . . . 229 4.10. Comets with unique polarimetric properties . . . . . . . . . . . . . . . . . . . . . . . . . 231 4.11. Database of Comet Polarimetry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 237 ACRONYMS AND ABBREVIATIONS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 240 REFERENCES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 244 ABOUT THE AUTHORS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 280 INDEX . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 286 Editorial Aerosol and cloud particles exert a strong influence on the regional and global climates of the Earth and other planets. Microscopic particles forming the regolith surfaces of many Solar System bodies and cometary atmospheres have a strong and often controlling effect on many ambient physical and chemical processes. More- over, they are “living witnesses” of the history of the formation and evolution of the Solar System and can tell us much about the events that have taken place over the past ~ 5 billion years in the circumsolar part of the Universe. Thus, detailed and ac- curate knowledge of the physical and chemical properties of such particles has the utmost scientific and practical importance. More often than not it is impossible to collect samples of such particles and subject them to a laboratory test. Therefore, in most cases one has to rely on theo- retical analyses of remote measurements of the electromagnetic radiation scattered by the particles. Fortunately, the scattering and absorption properties of small parti- cles often exhibit a strong dependence on their size, shape, orientation, and refrac- tive index. This factor makes remote sensing an extremely useful and often the only practicable means of physical and chemical particle characterization in geophysics and planetary astrophysics. For a long time remote-sensing studies had relied on measurements of only the scattered intensity and its spectral dependence. Eventually, however, it has become widely recognized that polarimetric characteristics of the scattered radiation contain much more accurate and specific information about such important properties of particles as their size, morphology, and chemical composition. The progress in polarimetric remote-sensing research has always been ham- pered by the fact that the human eye is “polarization blind” and responds only to the intensity of light impinging on the retina. As a consequence, to give a simple defini- tion of polarization readily intelligible to a non-expert is almost as difficult as to describe color to a color-blind person. However, continuing progress in electro- magnetic scattering theory coupled with great advances in the polarization meas- urement capability has resulted in overwhelming examples of the immense practical power of polarimetric remote sensing which are no longer possible to ignore. As a result of persistent research efforts, polarimetry has become one of the most infor- mative, accurate, and efficient means of remote sensing. Many prominent scientists have contributed to the foundation of polarimetry as a major remote-sensing disci- pline in geophysics and planetary astrophysics. An important and often decisive role in this process has been played by the authors of this monograph. The main objective of this book is to summarize the contributions to the field of polarimetric remote sensing of Solar System bodies which we and our colleagues 12 Editorial have made over the past four decades. We believe that doing this is useful because a specialized monograph on this subject appears to be lacking and also because of the systematic and comprehensive nature of our collective research activities. Specifi- cally, our work has included the development of a complete and rigorous theory of electromagnetic scattering by disperse media and accurate, physically based mod- eling tools required for the analysis of polarimetric measurements. We have ad- vanced theoretical fundamentals and principles of measurements of polarized ra- diation and used them to design unique and precise instrumentation. Based on analyses of extensive ground-based, aircraft, and spacecraft observations, we have determined, for the first time, the optical and physical characteristics of surfaces and atmospheres of numerous Solar System objects (such as planets, satellites, Sat- urn’s rings, asteroids, trans-Neptunian objects, and comets) as well as discovered a significant number of new phenomena and effects. We have also helped develop and implement unprecedented polarimetric techniques for the remote sensing of aerosol and cloud particles in the terrestrial atmosphere from aircraft and orbiting satellites. This monograph is not intended as an exhaustive review of all the countless ap- plications of polarimetry in geophysics and astrophysics. It is also not intended as a college-level textbook similar to those by Stephens (1994) and Ɇɨɪɨɠɟɧɤɨ (2004). We believe, however, that it can form the basis of a specialized graduate-level course in polarimetric remote sensing. With this purpose in mind, we have made the book as self-contained as the practical limitations imposed by the publisher have permitted. Important supplements to this book are the monographs by Mishchenko et al. (2002a, 2006b) and Hovenier et al. (2004), where detailed theoretical deriva- tions and further information can be found. The first of them is in the public domain and is available in the PDF format at http://www.giss.nasa.gov/staff/mmishchenko/ books.html. Based on the specific character of this book, the reference list was not intended to be comprehensive. The reader will also notice that some publications included in the reference list exist only in the Russian or the Ukrainian language. As such, they may not be particularly useful to those who do not speak either language. In several cases, however, there is an official English translation of the Russian or Ukrainian original. The corresponding bibliographic information is included in the reference list to the extent possible. Following Mishchenko et al. (2002a, 2006b), we denote vectors using the Times bold font and matrices using the Arial bold font. Unit vectors are denoted by a caret, whereas dyads and dyadics are denoted by the symbol ļ. The Times italic font is reserved for scalar quantities, important exceptions being the square root of minus one, the differential sign, and the base of natural logarithms, which are de- noted by Times roman characters i, d, and e, respectively. Another exception is the relative refractive index, which is denoted by a sloping sans serif m. For the reader’s convenience, the list of most important acronyms and abbreviations is in- cluded at the end of the book. Finally, it is worth mentioning that the optical particle-characterization tech- niques developed by these authors have found extensive applications in many other areas of science and technology such as nanoscience, nanotechnology, biomedicine, 13 Editorial nanobiotechnology, biophotonics, chemistry of colloids and suspensions, solid-state physics, laser physics, etc. In fact, our publications have been cited in more than 330 different journals, including Biophysical Journal, Cancer Research, Cytometry, Journal of Colloid and Interface Science, Journal of Dairy Science, Journal of Nanoscience and Nanotechnology, Journal of Polymer Science, Metamaterials, Nanobiotechnology, Nanotechnology, Progress in Energy and Combustion Science, Solar Energy Materials and Solar Cells, etc. We view this as a convincing example of how relatively modest investments in remote-sensing and astrophysics research can generate a great practical return highly beneficial to the entire society. We also hope that this factor will help increase the usefulness of our book and broaden its readership. Consistent with the above rationale, this monograph is intended for science pro- fessionals, educators, and graduate students specializing in remote sensing, astro- physics, atmospheric physics, optics of disperse and disordered media, optical par- ticle characterization, biomedicine, and nanoscience. M. I. Mishchenko, V. K. Rosenbush, N. N. Kiselev, Editors Kyiv – New York Acknowledgments We are grateful to Academician Yaroslav S. Yatskiv for suggesting the idea of the English edition of this monograph and for his generous support during all stages of this project. We also appreciate helpful and highly professional assistance from the staff of the academic publisher “Akademperiodyka”. It is a great pleasure to thank the following colleagues for numerous discussions and productive long-term collaborations: Kirill Antoniuk, Yuri Barabanenkov, An- tonella Barucci, Matthew Berg, Anatoli Borovoi, Oleg Bugaenko, Brian Cairns, Barbara Carlson, Alberto Cellino, Jacek Chowdhary, Petr Chýlek, Carl Codan, An- thony Davis, Viktor Degtyarev, Zhanna Dlugach, Helmut Domke, Oleg Dubovik, Bryan Fafaul, Sonia Fornasier, Igor Geogdzhayev, Roald E. Gershberg, James Han- sen, Otto Hasekamp, Steven Hill, Alfons Hoeskstra, Ronald Hooker, Helmuth Horvath, James Hough, Joop Hovenier, Vsevolod Ivanov, Klaus Jockers, Michael Kahnert, Nikolai Khlebtsov, Alexander Kokhanovsky, Sergej Kolesnikov, Ludmilla Kolokolova, Miroslav Kocifaj, Theodore Kostiuk, Andrew Lacis, Kuo-Nan Liou, Pavel Litvinov, Li Liu, Kari Lumme, Andreas Macke, Daniel Mackowski, Hal Mar- ing, Pinar Mengüç, Alexander Morozhenko, Karri Muinonen, Olga Muñoz, Elena Petrova, Vilppu Piirola, William Rossow, Yuri Shkuratov, Larry Travis, Cornelis van der Mee, Feodor Velichko, Sergei Velichko, Gorden Videen, Tõnu Viik, Niko- lai Voshchinnikov, Warren Wiscombe, Thomas Wriedt, Ping Yang, Edgard Yano- vitskij, Maksim Yurkin, Nadia Zakharova, and Eleonora Zege. We are grateful to Michael A’Hearn, David Shleicher, Tony Farnham, and Rita Shulz for providing a set of narrowband spectral filters which we have used extensively in observations of comets. Last but not least, we thank our families for their unconditional and de- voted support and understanding. Our research has been sponsored by many grants from the National Academy of Sciences of Ukraine, the National Aeronautics and Space Administration of the USA, the Academy of Sciences of the Former USSR, INTAS, the US Civilian Re- search & Development Foundation, and the National Polar-orbiting Operational Environmental Satellite System program of the USA. M.I.M. acknowledges contin- ual support from the NASA Radiation Sciences Program managed by Hal Maring and from the NASA Glory Mission project. I. N. Belskaya, Y. S. Efimov, V. G. Kaydash, N. N. Kiselev, D. F. Lupishko, M. I. Mishchenko, V. K. Rosenbush, N. M. Shakhovskoy, V. P. Tishkovets Kharkiv – Kyiv – Nauchny – New York Introduction Despite the immense diversity of applications of remote sensing in geophysics and planetary astrophysics, each application has the same basic stages summarized in the following block diagram: Specific problem of geophysics or planetary astrophysics Formulation of remote-sensing problem Development of remote-sensing instrumentation Development of physically based analysis techniques Observations Data analysis The specific character of a particular geophysical or astrophysical problem ul- timately defines the complexity of the requisite remote-sensing approach. The dou- ble-headed arrows in the diagram signify the inherent interdependence of all the basic stages. For example, newly performed observations may demonstrate the need for improved instrumentation, while difficulties encountered at the data-analysis stage may ultimately necessitate a complete reformulation of the remote-sensing problem whenever the remote-sensing methodology adopted proves to be inade- quate. The main objectives of this monograph are (i) to discuss in detail all basic stages of polarimetric remote sensing and (ii) to demonstrate that polarimetry is a powerful (and often the only adequate) remote-sensing methodology which can greatly outperform remote sensing techniques based on measurements of only in- tensity. The main theme of the book is remote sensing of objects composed of (or 16 Introduction containing) small particles, such as atmospheres of the Earth, other planets, and comets as well as planetary particulate surfaces. Accordingly, in Chapter 1 we present a complete and rigorous theory of elec- tromagnetic scattering by disperse media directly based on classical electromagnet- ics. We discuss numerically exact computer solutions as well as asymptotically ac- curate analytical solutions of the Maxwell equations (such as the T-matrix approach and the microphysical theories of radiative transfer and coherent backscattering) and describe state-of-the-art physically based modeling tools. Chapter 2 begins with an illustrative theoretical analysis of polarimetry as a re- mote-sensing tool. It then outlines basic principles of polarimetric measurements and measurement-error budgeting and gives an overview of several practical hard- ware implementations of the polarimetric technique. In two final chapters, we describe the results of our extensive photopolarimetric observations of numerous Solar System objects such as planets, planetary satellites, asteroids, and comets. We demonstrate how theoretical analyses of these data can be used to retrieve optical and physical characteristics of particles constituting planetary surfaces and atmospheres as well as to identify a number of new physical phenomena and effects. 100 μm 50 nm mm1 (a) 2.7 μm (b) (c) 200 nm 1 μm (d) 0.5 (e) (f ) 100 μm (g) Plate 1.1. Examples of man-made and natural small particles. (a) 40-nm-diameter gold particles (after Khlebtsov et al. 1996). (b) Sahara desert sand (after Volten et al. 2001). (c) Dry sea-salt particles (after Chamaillard et al. 2003). (d) Fly ash particles (after Ebert et al. 2002). (e) A soot aggregate (after Li et al. 2003). (f ) A 6-mm- diameter falling raindrop. (g) Cirrus cloud crystals (after Arnott et al. 1994). 15 10 5 r e t e m a r a p e z i s e v i t c e f f E 0 0 300 100 n o i t c n u f e s a h P 10 1 m = 1.3 m = 1.425 m = 1.55 (a) 60 120 Scattering angle (deg) 180 0 60 120 Scattering angle (deg) 180 0 120 60 Scattering angle (deg) 180 −100 − 80 − 60 − 40 −20 0 00 20 40 60 80 100 10 1 Spheres ε = 1.2 ε = 1.4 ε = 1.6 ε = 1.8 ε = 2 ε = 2.2 ε = 2.4 Geometrical optics x se = 40 x se = 80 x se = 180 (b) 0.1 0.06 0 0.1 (c) 60 120 Scattering angle (deg) 180 0 60 120 Scattering angle (deg) 180 Plate 1.2. (a) Degree of linear polarization for unpolarized incident light (in %) versus scat- tering angle and effective size parameter for polydisperse spherical particles with relative re- fractive indices 1.3, 1.425, and 1.55. (b) T-matrix computations of the phase function for mi- crometer-sized polydisperse spheres and randomly oriented surface-equivalent prolate sphe- roids with aspect ratios ranging from 1.2 to 2.4 at a wavelength of 443 nm. The relative re- fractive index is fixed at 1.53 + i0.008. (c) Geometrical-optics and T-matrix phase functions for monodisperse, randomly oriented circular cylinders with surface-equivalent-sphere size parameters xse = 40, 80, and 180. The relative refractive index is 1.311 + i0.311×10–8 and is typical of water ice at visible wavelengths.
0912.2074
1
0912
2009-12-10T19:50:22
(47171) 1999 TC36, A Transneptunian Triple
[ "astro-ph.EP" ]
We present new analysis of HST images of (47171) 1999 TC36 that confirm it as a triple system. Fits to the point-spread function consistently show that the apparent primary is itself composed of two similar-sized components. The two central components, A1 and A2, can be consistently identified in each of nine epochs spread over seven years of time. In each instance the component separation, ranging from 0.023+/-0.002 to 0.031+/-0.003 arcsec, is roughly one half of the Hubble Space Telescope's diffraction limit at 606 nm. The orbit of the central pair has a semi-major axis of a~867 km with a period of P~1.9 days. These orbital parameters yield a system mass that is consistent with Msys = 12.75+/-0.06 10^18 kg derived from the orbit of the more distant secondary, component B. The diameters of the three components are dA1= 286(+45,-38) km, dA2= 265(+41,-35 km and dB= 139(+22,-18) km. The relative sizes of these components are more similar than in any other known multiple in the solar system. Taken together, the diameters and system mass yield a bulk density of p=542(+317,-211) kg m^-3. HST Photometry shows that component B is variable with an amplitude of >=0.17+/-0.05 magnitudes. Components A1 and A2 do not show variability larger than 0.08+/-0.03 magnitudes approximately consistent with the orientation of the mutual orbit plane and tidally-distorted equilibrium shapes. The system has high specific angular momentum of J/J'=0.93, comparable to most of the known Transneptunian binaries.
astro-ph.EP
astro-ph
(47171) 1999 TC36, A TRANSNEPTUNIAN TRIPLE S. D. Benecchia,*, K. S. Nolla, W. M. Grundyb, and H. F. Levisonc Accepted to Icarus 9 December 2009 a Space Telescope Science Institute, 3700 San Martin Dr., Baltimore, MD 21218 * Corresponding Author E-mail address: [email protected] b Lowell Observatory, 1400 W. Mars Hill Rd., Flagstaff, AZ 86001 c Dept. of Space Studies, Southwest Research Institute, 1050 Walnut St. #400, Boulder, CO 80302 Pages: 16 Figures: 8 Tables: 6 Proposed running head: A Transneptunian Triple Corresponding author: Susan D. Benecchi Space Telescope Science Institute 3700 San Martin Dr. Baltimore MD 21218 [email protected] phone: 410-338-4051 fax: 410-338-5090 ABSTRACT We present new analysis of HST images of (47171) 1999 TC36 that confirm it as a triple system. Fits to the point-spread function consistently show that the apparent primary is itself composed of two similar-sized components. The two central components, A1 and A2, can be consistently identified in each of nine epochs spread over seven years of time. In each instance the component separation, ranging from 0.023±0.002 to 0.031±0.003 arcsec, is roughly one half of the Hubble Space Telescope’s diffraction limit at 606 nm. The orbit of the central pair has a semi-major axis of a~867 km with a period of P~1.9 days. These orbital parameters yield a system mass that is consistent with Msys = 12.75±0.06 1018 kg derived from the orbit of the more 286−38 distant secondary, component B. The diameters of the three components are dA1= +45 km, 265−35 139−18 dA2= +41 km and dB= +22 km. The relative sizes of these components are more similar than in any other known multiple in the solar system. Taken together, the diameters and system mass 542−211 yield a bulk density of ρ= +317 kg m-3. HST Photometry shows that component B is variable € with an amplitude of ≥0.17 ± 0.05 magnitudes. Components A1 and A2 do not show variability larger than 0.08 ± 0.03 magnitudes approximately consistent with the orientation of the mutual € orbit plane and tidally-distorted equilibrium shapes. The system has high specific angular momentum of J/J’=0.93, comparable to most of the known Transneptunian binaries. € € Subject headings: Hubble Space Telescope Observations; Satellites — Composition; Kuiper Belt; Satellites of Asteroids; Photometry 1. INTRODUCTION The discovery of a sizable fraction of binaries among small bodies in the solar system, ~15%, leads to the obvious question of whether higher order systems are also common (e.g. Richardson and Walsh 2006; Noll et al. 2008). Two triple systems have been identified among Near-Earth objects: (153591) 2001 SN263 (Nolan et al. 2008) and (136617) 1994 CC (Brozovic et al. 2009) and four triple systems are found among the Main Belt asteroids: (45) Eugenia (Marchis et al. 2007), (87) Sylvia (Marchis et al. 2005), (216) Kleopatra (Marchis et al. 2008), and (3749) Balam (Marchis et al. 2008). Two multiple systems have been identified among Transneptunian objects (TNOs): the two small satellites of the Pluto/Charon binary dwarf planet (Weaver et al. 2006), and two small satellites orbiting the dwarf planet Haumea (Brown et al. 2006). (47171) 1999 TC36 orbits the Sun in the 3:2 resonance with Neptune with a heliocentric semi-major axis of 39.668 AU, an eccentricity of 0.227, and an ecliptic inclination of 8.409°. A secondary was identified in 2002 in images obtained by the Space Telescope Imaging Spectrograph (STIS; Trujillo & Brown, 2002). In the discovery images, the secondary, which we will refer to as component B, was found at a separation of 0.36 arcseconds from the apparent primary and was fainter by 2.21±0.01 magnitudes. Between 2003-2006, (47171) 1999 TC36 was observed at seven more epochs with the Advanced Camera for Surveys High Resolution Camera (HRC) leading to a determination of the orbit of the secondary with a period of ~50 days (Margot et al. 2005). 2 ACCEPTED TO ICARUS - PREPRINT There have been two arguments used to suggest that (47171) 1999 TC36 might be a triple system. Stansberry et al. (2006, 2008) found an effective surface area for the unresolved system 414.6−38.2 of dsys= +38.8 km from thermal infrared observations made with the Multiband Imaging Photometer for SIRTF (MIPS) instrument onboard the Spitzer Space Telescope. Using a system 500−200 mass of Msys = 14.4x1018 kg (Margot et al. 2005) they derived a bulk density of ρ= +400 kg m-3. For mixtures of water ice and silicate rock, a density this low requires a bulk porosity of 50-75%. Stansberry et al. noted large residuals in the PSF fitting of the primary and speculated that it € might itself be a close binary, allowing for a somewhat higher density. Since the Stansberry et al paper, a significant number of similar sized solar system objects (including other TNOs) have € been measured to also have low densities (e.g. Spencer et al. 2005; Grundy et al. 2007). Jacobson & Margot (2007) re-examined the HRC images and found that the apparent primary, component A, appeared to be elongated. With a contour-fitting routine, they found an ellipticity of 0.21±0.05 at a wide range of position angles. Attempts to fit the images with a single point-spread function (PSF) did not adequately reproduce the images of the primary. An F-Test showed with 99.9% confidence that two PSF components are required to properly model the primary. However, detailed binary fits were not presented and the relative positions and size of the components remained undetermined. No orbit solution for the central pair was given. In this paper we provide unequivocal evidence for the triple nature of (47171) 1999 TC36, including a consistent determination of the positions and relative fluxes of all three components using all available Hubble Space Telescope (HST) data. We model the images with a PSF-fitting code and find that the system is composed of two nearly equal-sized components separated by 867±11 km with a third component orbiting the central pair at a distance of 7411±12 km (these are average separations from the orbit fit, not sky plane separations). In Section 2 we define the terminology we use throughout the paper, in Section 3 we describe the observations, and in Section 4 we describe our data reduction and analysis procedures. In Sections 5 and 6 we present our orbit solution for both components and evaluation of the photometric variability of the system. In Section 7 we present a re-interpretation of system measurements in the new three-body configuration followed by our conclusions in Section 8. TERMINOLOGY 2. In discussing a newly identified component of the (47171) 1999 TC36 system we recognize the need for a consistent terminology. Throughout this paper we will refer to the apparent primary as “component A” and the previously identified secondary as “component B”. When we discuss the individual components of the apparent primary we will refer to them as “component A1” and “component A2”. 3. OBSERVATIONS In this work we present a re-analysis of all available HST images1 of (47171) 1999 TC36. (47171) 1999 TC36 was first observed with HST in two separate orbits using the STIS CCD camera in 2001. In each orbit sixteen 120 second images were obtained through the clear filter. The two orbits were separated by 9.62 hours. Beginning in 2003, the HRC was used to obtain 1 The data are available from the HST data archive at http://archive.stsci.edu ACCEPTED TO ICARUS - PREPRINT 3 images in seven HST visits, altogether spanning slightly more than three years. Each visit (a single HST orbit each) included four un-dithered individual observations, 610 seconds in duration, acquired in the F606W filter. During one of the HRC visits, two images were acquired in an F814W filter for color analysis. The circumstances of all the observations are recorded in Table 1. INSERT Table 1 HERE We also obtained data of (47171) 1999 TC36 over two consecutive nights (UT 2008-09- 25,26,27) using the Mt. Bigelow 61-inch in Arizona and three consecutive nights (UT 2008-09- 30, 10-01,02) using the CTIO 4-m telescope in Chile. The 61-inch data were collected in V and I filters with the Mont4k camera and the CTIO data were collected in a Sloan r’ filter with the Mosaic camera, a 8k x 8k CCD array operated in the 2x2 binned mode with a binned-pixel scale of 0.52 arcsec. A total of 125 exposures, 300-600 seconds in duration, were obtained. A log of observations, the geometric including comments about atmospheric conditions and circumstances during our observations, can be found in Table 1. At Mt. Bigelow both nights were photometric. At CTIO the first two nights were photometric with poor seeing [1.5-1.8 arcsec full-width-half-maximum (FWHM)], while the third night had better seeing (0.8-1.5 arcsec FWHM) with variable cirrus. The seeing at Mt. Bigelow was comparable to the first two nights at CTIO. Photometry was possible for the full five night span using on-chip reference stars calibrated on the photometric nights as described in Section 6.3. 4. HST POINT-SPREAD FUNCTION (PSF) FITTING We processed the HRC and STIS data through the standard HST pipeline (Pavlovsky et al. 2006, Dressel, L. et al. 2007) which performs basic image reduction: it flags static bad pixels, performs A/D conversion2, subtracts the bias, and dark images, and corrects for flat fielding. It also updates the header with the appropriate photometry keywords. The flat-field-calibrated images, subscripted by flt, were the ones we analyzed (for both instruments). We use the individual images instead of the distortion-corrected drizzled image in order to obtain 4 separate measurements for each HST epoch which we can examine both graphically and statistically to evaluate the significance of our result. Our fitting process includes modeling the geometric distortion in the HRC PSF. Our post-pipeline analysis process for the (47171) 1999 TC36 images is the same as was described for HRC images in Benecchi et al. (2009; Grundy et al. 2009), with some additional testing to determine the robustness of our model fits for the central pair, as we describe in section 4.3. We adapted the same fitting software for the STIS images and followed the same basic procedures with a few exceptions as noted. We performed the PSF-fitting analysis in two steps. First we obtained a best-fit model for component B. This component of the system is sufficiently far away (6-13 pixels) from component A that we could remove it from the image completely for a cleaner investigation of the central object. Next we modeled the central component as both a single and a binary. We compared the results and found that the binary fits provided a statistically significant improvement over the single object fits. For testing purposes, we also analyzed HRC images of (26308) 1998 SM165, a TNO binary with similar brightness, separation and relative magnitude of the two components. 2 A/D conversion takes the observed charge in each pixel in the CCD and converts it to a digital number with the appropriate gain setting. The gain for ACS/HRC is 2.216 e-/DN. 4 ACCEPTED TO ICARUS - PREPRINT 4.1. Component B We began our PSF-fitting by estimating the central positions of component B and component A (that we consider initially as a single object) by eye (good to about 0.5 pixels). We then sum the object flux in a 1.5 pixel radius around each position to get an initial estimate for the scaling of the PSF. We carried out calculations on sub-images 36-50 pixels on a side (the scale of the HRC is 0.025 arcsec/pixel), centered on component A. The STIS sub-images were 24 pixels on a side (the scale of STIS is 0.05 arcsec/pixel). The sub-images were defined to be large enough to encompass both components of the system and to ensure sufficient sky so that the background could be adequately computed. Bad pixels from cosmic rays and hot pixels within the sub-images were flagged and ignored in further calculations prior to determining the background. For the HRC images we multiplied by a “pixel area map” (PAM)3 correction image prior to the PSF fitting to correct the photometry for geometric distortions in the images; this correction was not applied to the STIS images. At the location of (47171) 1999 TC36 the PAM correction is a factor of 1.127. The initial estimations for position and flux were further refined by creating a series of scaled and oversampled (by a factor of 4) synthetic PSF images using the Tiny Tim software package developed for analyzing Hubble images (Krist & Hook 2004). For the HRC data, the PSFs were generated at the estimated position for each component (to account for the geometric distortion of the HRC) using a stellar spectral distribution chosen to simulate typical TNO colors [model 57 from the Bruzual-Persson-Gunn-Stryker spectrophotometric atlas (Gunn & Stryker 1983) which has a B-V=0.92, close to the median for TNOs (Hainaut & Delsanti, 2002)]. For the STIS images, where geometric distortion is not an issue, a PSF was generated with the same stellar flux distribution as the HRC, but at a single location, then shifted to the position of the object in the data. We determined each of seven components: average background, position (x1, y1) and flux (f1) of the primary and position (x2, y2) and flux (f2) of the secondary individually by minimizing the χ2 of the residual between the model and the sub-image while all other values were held fixed. Next we found values for two additional model parameters that account for small thermally-induced focus changes known as “breathing” and for small motions of the spacecraft known as “jitter”. There is some ambiguity between the focus and jitter component values – i.e jitter can be masked by a larger focus value and vice versa. We model the breathing first by adjusting the focus (z4) parameter in the input file for Tiny Tim. It is determined in an iterative fashion, bounding the value on both the positive and negative sides of zero (change of focus units in 0.05 step intervals then 0.001 step intervals as the minimum is determined) and investigating the focus values stepwise to minimize the χ2 residual. When we fit the focus value, we hold all the other parameters fixed at the best fit position, flux and background values from our initial fit. Including the focus parameter in the model has the effect of improving the χ2 residual by ~20%. Typical focus values were –0.092 to 0. Next we modeled both the jitter and the non-negligible angular diameter of the objects by convolving the model PSF with a Gaussian smoothing function (telescope jitter+object diameter), in pixel space. For the component diameters derived by Stansberry et al. (2008), the angular diameters of components A and B are 0.017 and 0.006 arcsec, respectively. For models where components A1 and A2 are considered individually the angular diameters are 0.013 and 0.012 arcsec, respectively. This translates to approximately half a pixel in the HRC and a quarter 3 http://www.stsci.edu/hst/acs/analysis/PAMS/HRCarea.gif ACCEPTED TO ICARUS - PREPRINT 5 of a pixel in STIS. We minimize the χ2 residual to determine the jitter value, where the telescope jitter value can change and the object diameter is fixed, in a step-wise fashion similar to the focus value determination. The jitter value is ~0.11 pixels. Next we implement an automated fitting process that uses the amoeba (Press et al. 1992) routine, which performs multidimensional minimization of a function containing all our position/flux and background variables using the downhill simplex method, to optimize the fits. The routine does not consider focus or jitter. We ran this automated routine iteratively with the focusing and jitter routines until the χ2 converged. Typically, 4-5 iterations were required to reach a final PSF model. From the final model parameters we extracted astrometry and photometry for both components. Photometry was extracted for the HRC data from the best-fit PSF as described in Benecchi et al. (2009). Absolute astrometry was extracted at the fitted positions using IRAF’s xytosky routine and applying the geometric distortion correction table coefficients for the HRC. It was extracted using IRAF’s xy2rd routine for the STIS images. The astrometric results for component B, relative to A1, can be found in Table 2. INSERT Table 2 HERE. 4.2. Component A In order to better evaluate the central object, component A, we subtracted the model of component B from the original image and worked only with component A. For this portion of the analysis we used a sub-image box that was 26 pixels on a side for the HRC and 20 pixels on a side for STIS. To determine if component A is really a binary, we started with the highest resolution data, HRC, and fit component A as a single object for a reference point; we fit only the (x1,y1) position, (f1) flux and background starting with the best fit results from our previous model. We ran the automatic, amoeba, fitting process and found a result, as expected, similar to our first (component A and B) fit. While a formal best fit could be found, the reduced χ2 ( 2 A) were on χν the order of 3-5 indicating a poor fit. Next we tested for a binary fit. We set the position of a secondary component a half pixel away from the single fit location and set the flux of each component to half of the total flux from the single fit and modified the factor added to the jitter to account for the diameter of the central € component assuming that the components are equal in size. We leave the focus fixed. We then ran the automatic fitting allowing any of the parameters to change (x1, y1, f1, x2, y2, f2 and background) for χ2 minimization. We find positional consistency among the results of a single visit, within 0.1 pixels (~0.002 arcseconds). The flux ratios values are the least consistent value in the fit, however the ratio is near 1. The primary effect of the flux ratio being near 1 is that it complicates the orbit fitting due to component ambiguity; we discuss this in Section 5.2. Our fitted positions and flux ratios, listed as component A2 relative to component A1, are provided in Table 3 along with the reduced χ2 for both the single ( 2 A) and binary ( 2 A1, A2) fits. The χν χν absolute astrometric results for the pair component A2 relative to A1 are found in Table 2. INSERT Table 3 HERE. We performed the same basic analysis on the STIS images, however, we skipped the step of fitting the inner component as a single since the HRC analysis confirmed the existence of two € € components and our fit for component B gave a reference point for component A as a single. The pixel scale on STIS is about double that on the HRC and as such there is more scatter in the results, in both pixel space and the residuals. The separation of components A1 and A2 with ACCEPTED TO ICARUS - PREPRINT 6 STIS is slightly less than a pixel (~0.7). Additionally, the STIS data were acquired without moving target tracking so in some cases the PSF trails by 0.5-1 pixels. When plotting the positions of the individual STIS data points we found each visit had points scattered in a region spanning about ~20° (we use the average value of all the points in each visit for the astrometric analysis). However, in each visit, the quadrant of the resulting positions was consistent with the A1-A2 orbit determined with the HRC data. Since we observe precession during the timespan of the HRC data and the STIS astrometric points were further way in time and have larger uncertainties, we did not re-fit the orbit to include the STIS data. 4.3. Variations in Initial Conditions In order to check the robustness of our PSF fits with respect to the initial guesses for component positions, and fluxes, we conducted a series of tests. These include: (1) starting the components at wide (a few pixels) separations, (2) starting the components with extreme flux values (f1~106, f2~0), (3) fitting for a range of different fixed-jitter values, and (4) re-running the focus determination. None of these changes to the inputs of the model result in χ2 minimization at significantly different positions (less than 0.01 pixels) or fluxes. 4.4. (26308) 1998 SM165 We also tested our fitting technique on another binary with similar characteristics to (47171) 1999 TC36, also observed with the HRC. (26308) 1998 SM165 is in a 2:1 mean motion resonance with Neptune, one of 17 such objects so far to be discovered. A binary companion roughly 1.9 magnitudes fainter was found in images obtained using STIS in late 2001 (Brown and Trujillo, 2002). By combining the measured absolute magnitude of the unresolved pair in the V band, HV = 6.13 (Romanishin & Tegler 2006), and thermal observations made with Spitzer, Spencer et al. (2005) found a diameter of 287 ± 36 km for the primary and 96 ± 12 km for the secondary component. Using the system mass from HST observations (Margot et al. 2005) yields 510−140 a system density of +290 kg m-3. The observations of (26308) 1998 SM165 that we used to test our PSF fitting were obtained as part of the same observing program as the (47171) 1999 TC36 HRC data and are identical in terms of filters and exposure times (obtained 2003-2006, 610 second exposures, F606W filter). Likewise, the time of the HST observations the component separations € (s26308=0.294±0.001, s47171=0.306±0.001 arcsec) and size ratios (R1_R226308=2.75±0.06, R1_R247171=2.81±0.03 assuming the same albedos for both objects and their components) are nearly identical making (26308) 1998 SM165 an ideal test case. Since the HRC images have the highest resolution of all the observations, these were the ones on which we tested our modeling. We used the same fitting sequence on images of (26308) 1998 SM165 as we did for (47171) 1999 TC36. We fit the clearly resolved secondary, subtract it out of the image then fit the central component as either a single or a binary. The binary fit for the central component of this object results in identical positions for the components within 0.04 pixels, or alternatively a flux of nearly zero for the test secondary component. We show in Figure 1 a typical sample of the fitting results for the analysis step after component B is subtracted from the image for both (47171) 1999 TC36, in two filters, and (26308) 1998 SM165. In each triple set of images, the left panel displays the observed data, the middle panel displays the best-fit model PSF and the right panel displays the residuals (observed – model). The data and model are scaled identically and the residuals are scaled to increase the ACCEPTED TO ICARUS - PREPRINT 7 dynamic range. It is clear from the left half of the figure panels A and B that component A of (47171) 1999 TC36 is not well fit with a single PSF model. However, it is also clear that if the central component of (47171) 1999 TC36 were not binary, it would be well modeled by a single PSF as we see from our model fit results for (26308) 1998 SM165 (panel C). The residuals from the two component PSF model in the right half of figure yield very similar residual patterns to those for (26308) 19998 SM165. In Figure 2, we provide a sample of the fitting results from one exposure in each of the seven HRC visits to (47171) 1999 TC36 to demonstrate the repeatability of our results. The orientations of the fields (indicated by the directional in the upper left corner of the panel for each visit) are not identical, but are similar, and it is clear that the two components move with respect to each other between observations. INSERT Figure 1 HERE INSERT Figure 2 HERE 5. ORBIT FITTING We used the component positions determined from our PSF-fitting to determine the mutual orbit of the A1/A2 pair and to re-derive the orbit of component B relative to the A1-A2 barycenter (which we define as half-way between components A1 and A2 using a mass fraction of 0.5, or equal masses). We considered each visit a single astrometric position and averaged the measurements using the scatter in the positions to provide a measure of the uncertainties. We fit the period, semi-major axis, eccentricity, inclination, and the angles: mean longitude at epoch (ε), longitude of ascending node (Ω) and longitude of periapsis (ϖ). We run a Monte Carlo simulation on the orbit parameters to derive the uncertainties on each of these parameters and calculate the system mass. 5.1. Component B The orbit of component B (Table 4) was determined from astrometric data that included two epochs of STIS and seven epochs of HRC observations. We found that the seven HRC data points that are closest in time gave a result with the smallest reduced χ2. The same basic orbit was derived when we added the STIS points farther away in time, however the reduced χ2 of the orbit increased. Such a result is expected if the system is not Keplerian and there are perturbations within the system that our model does not account for, such as the effects of the non-point mass distribution of the inner pair, which could be approximated by including a quadrupole term (J2) in the primary's gravitational field. We don't have enough observations to actually fit for the masses of all 3 bodies, or even for J2, but we note that spreading the mass of A1 and A2 over a oblate spheroid with the equatorial dimensions of their mutual orbit would produce an effective J2R2 of 37584 km2, the inclusion of which reduces our chi-squared by more than a factor of two, using Danby (1988) equations 11.15.6 (which give the precession of the secondary's orbit assuming all of the mass and angular momentum reside in the primary). With our results, we can formally rule out the direct and mirror instantaneous Keplerian orbits, however, there are clearly other influences affecting the orbit interpretation. Our values for the orbital period, semi-major axis and system mass (P=50.302±0.001 days, a=7411±12 km, Msys=12.75±0.06 x1018 kg) are within the uncertainties of those found by Margot et al. (2005; P=50.38±0.5 days, a=7640±460 km, Msys=13.9±2.5 x1018 kg); no additional details for their solution have been published. ACCEPTED TO ICARUS - PREPRINT 8 5.2. Components A1 and A2 The orbit of components A1 and A2 (Table 4) were determined using only the HRC dataset, though the STIS positions are consistent with the orbit we find. We noted earlier that the components have very similar fluxes so we allowed component identification at each visit to be a free parameter in our initial orbit fitting iteration. We used the system mass derived from the orbit of component B to constrain the range of periods and semi-major axes searched. We performed a grid search in which the data were permuted over all 64 meaningful permutations for the 7 ACS visits (swapping identities for all 7 visits at once cannot produce a meaningful new permutation, but swapping a subset of them can, which is why there are 64, not 128). For each of these 64 possible permutations of the data, we used amoeba to fit a circular orbit with initial periods ranging from 1 to 2.35 days in 0.0005 day increments. All solutions having chi-squared values below a threshold of 500 were collected and examined by hand. We then ran fits that permitted the eccentricity to be non-zero, as well as fits for the mirror orbit. The solution appearing in Table 4 emerged as the best one with a χ2 of 42.6. The next best solution had χ2 only a little worse, at 89, but a mass nearly double that of the system mass from the orbit of B, so it seemed unlikely to be correct. The next best solution with a similar system mass had a χ2 over 200, considerably worse that our preferred solution. As we found with the orbit of component B, the smallest χ2 results from observations taken most closely in time. We find evidence for orbit perturbation with the system over the 3+ years of data included in our analysis. Figure 3 plots the best fit orbits of the two components. INSERT Table 4 HERE INSERT Figure 3 HERE We calculate the system mass to be Msys = 12.75±0.06 1018 kg using the component B system parameters and Msys =14.20±0.05 1018 kg using the component A1 and A2 system parameters. The discrepancy in our results are not completely unanticipated, similar findings were reported by Buie et al. (2006) for the Pluto system before the orbit was properly modeled to account for mutual perturbations among the bodies within the system (Tholen et al. 2008). We expect a similar solution exists for (47171) 1999 TC36, but such analysis is beyond the scope of this paper. If our solution for the system mass calculated from the component B orbit was correct, and the system was not perturbed, the mass of component B would be 0.746±0.001 1018 kg. In addition, based on the orbits we fit, the mutual event season for component B is centered on 2069 and on 2075 for components A1 and A2. However, we have assumed instantaneous Keplerian orbits for this projection excluding precession and we know that mutual perturbations can be expected to shift the timing somewhat, perhaps on the order of a decade. 6. PHOTOMETRY 6.1. Previous photometric observations (47171) 1999 TC36 is a relatively bright TNO and therefore has been the target of a number of photometric studies. Attempts to determine the lightcurve of the (47171) 1999 TC36 system (combined light of all components) have found it to be inconsistent across epochs. Peixinho et al., 2002, observed (47171) 1999 TC36 for 8.362 hrs and found no apparent periodic variation. Ortiz et al., 2003, found variation with a period of 6.21±0.02 hours, but were unable to obtain repeatable observations a month later. Its phase curve changes with wavelength from 9 ACCEPTED TO ICARUS - PREPRINT B’=0.255±0.044 mag/° to V’=0.131±0.049 mag/° to I’=0.235±0.056 mag/° (Rabinowitz et al. 2007). Barkume et al. (2008) and Dotto et al. 2003 found the surface to have a moderate ~8-20% ice fraction signature when observed spectroscopically. The visible colors of the unresolved A1+A2 pair, and component B are nearly identical and red, (V-I)A=1.19±0.01 and (V- I)B=1.12±0.03 (Jacobson, private communication 2007; Benecchi et al. 2009). 6.2. HST variability HST observations are intrinsically photometric. The sparse sampling obtained by HST, while unsuitable for lightcurve measurement, can be used to search for and constrain possible variability. In Figure 4 we plot the photometry of the largest dataset in a single filter, HRC F606W. The data cover 3 years of time. The measurement uncertainties are ~0.01 magnitudes and the variation within each 30 minute visit is small (0.01-0.05 magnitudes). When all the data are considered, the variation of component B is 0.17±0.05 magnitudes. The variation of component A1+A2 is 0.08±0.03 magnitudes. We suggest that if the components are resolved, and measured on a shorter timescale, both will show lightcurves with amplitudes greater than the variability found in the HST dataset. Because of the dynamics in the system, we hypothesize that component A1 and A2 are tidally locked with rotational periods of 45.74 hours. INSERT Figure 4 HERE 6.3. Ground based photometry We obtained two nights of ground-based data of (47171) 1999 TC36 at the Mt. Bigelow 61-inch telescope and three nights at the CTIO 4-m Blanco telescope (Table 1) in September/October 2008. The data were bias-subtracted and flat-fielded with standard procedures. We performed aperture photometry on (47171) 1999 TC36 as a system as well as about a dozen field stars for comparison. (47171) 1999 TC36 did not come close to any other stars during the five nights of our observations. The seeing throughout the runs was 1.0-1.8 arcseconds (2-4 pixels) and as such we used a photometric aperture radius of 12-14 pixels. We found that the individual measurements of (47171) 1999 TC36 yield S/N~90 for the CTIO data and S/N~10 for the 61-inch data (which we binned by 3 for analysis), and performed aperture correction photometry on our object to reduce background contamination, using the two most photometrically stable comparison field stars (at the level of 0.01 magnitudes). For the 61-inch observations we photometrically calibrated our object and the field stars with the Landolt standard star, SA 113_260 (Landolt, 1992). We performed similar analysis on nights one and two of the CTIO observations using the Landolt standard star, SA 113_339 (Landolt, 1992). We used the field stars to photometrically calibrate the data from the nights of our runs that were not photometric (night one at the 61-inch and night three at CTIO). Lastly, we shifted the 61-inch data which was collected in V and I filters to the Sloan r’ filter using a combination of measured offsets from the data and synphot modeling so that all the data could be analyzed together. In Figure 5 we plot the lightcurve from each night of data in a separate panel referenced to the first observation. The variation of the data is 0.20±0.04 magnitudes. We attempted to model the data with a weighted single peaked lightcurve with periods ranging from T=3-24 hours (6-48 hour double peaked periods), but do not find any satisfactory results. Periods of ~16±2 and 23±2 hours are the most prominent, but the residuals are not clean and the 23 hour period, which is about what we expect for a tidally locked system, is too close to our observing interval of 24 hours to be confident of its validity. The uncertainty in the amplitude of the model is ~30% of the amplitude itself, which ranges between 0.03±0.01 and 0.06±0.03 magnitudes for 10 ACCEPTED TO ICARUS - PREPRINT our two best-fit periods. It may be the case that the tidally locked period is correct for components A1 and A2, and a contribution of variation from component B complicates our ability to find a clean fit, we cannot say for certain. From Leone et al. 1984 we find that the equilibrium tidally distorted shape for two ω2 /πGρ = 0.012 (where ω is 2π/T and ρ is density) will have components of equal size with a/c~0.97 (where a and c are the smallest and largest diameters of the triaxial ellipsoid). The maximum possible lightcurve amplitude for an object of this shape is 0.03 magnitudes. However, because the orbit is inclined, the amplitude is reduced to 0.02 magnitudes by the cosine of the angle of the normal to the line of sight, 40°. Since components A1 and A2 dominate the € photometric signature, the lack of a strong lightcurve is therefore not surprising. Limits for the effect of the variability of component B on the unresolved system brightness can be calculated. The average delta between components A1+A2 and component B is 2.24±0.03 magnitudes (Table 6) and the average brightness of the system is 19.50±0.01 magnitudes. If component B varies by as much as 2 times that found from the HST images, it would contribute a maximum 0.04 magnitude amplitude to the unresolved system lightcurve (MTOTAL would range between 19.50 and 19.54). The variability we find for our combined system lightcurve are consistent with this range of variability for component B, though the period and contribution of each component to the variation remains unknown. Components A1 and A2 cannot have large (>0.2 magnitude) variations as we would have measured them in our data. INSERT Figure 5 HERE 7. DISCUSSION 7.1. Comparison to other multiple systems In the context of the other known multiple systems in the solar system’s small body populations, (47171) 1999 TC36 demarks an extreme in relative component sizes. It is composed of two nearly identically sized components in close proximity with a third component approximately half the size of components A1 and A2 nearly 10 times farther away. All the other known small body multiples contain at least one relatively tiny component (Table 5, Figure 6) and components are approximately evenly spaced. In most systems the more distant of the two outer components is the larger, which we note is not true for (47171) 1999 TC36. INSERT Table 5 HERE INSERT Figure 6 HERE In absolute terms, the known multiple systems range over almost three orders of magnitude in radius and over eight orders of magnitude in volume. Despite this very large range of absolute sizes, the system architectures are remarkably similar (in the appropriately scaled units). All of the systems are tightly bound, with components orbiting at a few percent of the Hill radius or less. With only 9 known multiples in the solar system, it is too early to know whether there are clear groupings of systems in terms of their physical and/or orbital properties or, instead, if there is a continuum of such systems. If the former, the architecture of multiples may hold clues to their modes of formation. 7.2. Density and Porosity The confirmation of (47171) 1999 TC36 as a triple impacts the derivation of its physical properties, in particular its density and porosity. These issues were first discussed by Stansberry 11 ACCEPTED TO ICARUS - PREPRINT € € € € ] ) 3 et al. (2006). We revisit these calculations using the newly determined relative brightnesses of components A1 and A2 to remove this formerly speculative quantity. To calculate the density and porosity of (47171) 1999 TC36 we used an effective system 414.6−38.2 diameter of dsys= +38.8 km (Stansberry et al. 2008). We assume that the components have equal albedos, 7% (Stansberry et al. 2008), and determine their individual diameters from their observed flux differences which are found in Table 6. We average all the F606W measurements 139−18 to find ΔmAB=2.24. From this we determine the diameter of component B to be dB= +22 km. Similarly, we find ΔmA1A2=0.17 yielding the individual diameters of the two inner components, € 286−38 265−35 dA1= +45 km and dA2= +41 km. The similarity of the sizes of the components is, so far, unique among the multiple systems in the Kuiper Belt. INSERT Table 6 HERE € We calculate the density of the system as: 3M € ρ = + dB 2 + dA 2 2 4π dA1 2 ) 3 ) 3 [ ( ( ( (1) 542−211 and find ρ= +317 kg m-3. For a range of material densities 1000 < ρ0< 2000 kg m-3, we find porosities, or fractional f = 1 − ρ ρ0 , of 46-73%. (47171) 1999 TC36, however, is not the only TNO with a void space 700−210 low density. (26308) 1998 SM165 has a density of +320 kg m-3 (Spencer et al., 2006), and € 440−170 (42355) Typhon has a density of +440 kg m-3 (Grundy et al. 2008). Low densities have also been reported for some of the icy satellites (Burns, 1986; McKinnon et al., 1995) and the Trojan € asteroid (617) Patroclus (Marchis et al. 2006; Mueller et al. 2009). € € 7.3. Angular Momentum The specific angular momentum of the (47171) 1999 TC36 system can be calculated by summing the orbital and spin angular momenta of the components and normalizing by the factor 3 Reff where G is the gravitational constant, Msys is the system mass and Reff is the ʹ′ J = GM sys radius of an equivalent spherical object containing the total system mass. The total angular momentum of (47171) 1999 TC36 is dominated by the orbit angular momentum given by 1 − ei , where mi is the mass of each component, ai is measured from the J orbit = m iai 2ωi 2 ( ) i∑ system barycenter, ei is the eccentricity of the orbit, and ωi is the orbital frequency. For this calculation we assumed that all of the components have the same density. The spin component of 2 the angular momentum is given by where ni is the rotational frequency of J spin = m i ri 2n i i∑ 5 each component. For this calculation we assumed A1 and A2 to be tidally locked so that their spin period is equal to their 1.906 day mutual orbit period. The more distant component B’s spin period is unknown; we assumed a period of 8 hours, typical for small bodies. With these assumptions, the spin component contributes only 2.6% of the total angular momentum. In the € extreme case all of the components are assumed to be rotating near break-up with a period of 3 hours; the spin component of the angular momentum rises to 27.7% of the total. All of the possible cases are qualitatively the same: the system’s total angular momentum is dominated by the orbit angular momentum. Additionally, the independent contribution to orbital angular 12 ACCEPTED TO ICARUS - PREPRINT momentum is similar, though not identical, for each of the three components. Component A1 contributes 33.7%, component A2, 26.4% and component B, 39.9%. Total system angular momentum can provide clues to the mode of formation of multiple systems (Figure 7). Systems formed by spin-up and fission must have J/J’<0.39 (Dobrovolskis et al. 1997). Many near earth and main belt systems may have formed this way. Multiples formed by collision have an upper limit of J/J’ ~0.8 (Canup 2005). This assumes that ratio of the v imp / v esc < 1.3 . While this is probably a safe impactor velocity to the escape velocity is assumption for large systems thought to have formed by collision like Pluto (Canup 2005) and Haumea (Levison et al. 2008), it is less clear whether this limit applies to smaller systems like (47171) 1999 TC36 where the escape velocity, ~0.1 km/s, is likely lower than random velocities in the disk. Most Transneptunian binaries (TNBs) have J/J’ >0.8. Formation by dynamical € capture (Goldreich et al. 2002, Noll et al. 2008) allows for higher system angular momentum and thus may be more consistent with the high angular momenta found in TNBs. Like most TNBs (47171) 1999 TC36 has a relatively large angular momentum, J/J’=0.93, and we suggest that it likewise may have been formed by two successive dynamical capture events. INSERT Figure 7 HERE 7.4. Non-Kelperian Orbits A full n-body solution to the orbital dynamics of the (47171) 1999 TC36 system requires more astrometric data than is currently available and is therefore beyond the scope of this paper. However, it is possible to investigate the general dynamical behavior of this system using our estimates of system parameters as a starting point for a model of the orbital evolution. As shown in Figure 8 we found that ϖ precesses through a full 2π radians in roughly 50 years. INSERT Figure 8 HERE € Tidal Evolution 7.5. For a two-component system, the time to circularize an initially eccentric orbit can estimated from (Goldreich & Soter 1966, eq. 25): 5 4Qm2 a 3 a ⎛ ⎞ . Tcirc = ⎟ ⎜ G(m1 + m2 ) 63m1 r2 ⎝ ⎠ (2) The semi-major axis of the orbit, a, the radius of the smaller component, r2, and the component masses, m1 and m2, are available from fits to the mutual orbits (Table 4 and derived values) and the tidal dissipation coefficient, Q, is estimated to range from 10-500 for rocky and icy bodies (Goldreich & Soter 1966). However, as noted by Noll et al. (2008) this equation has limited applicability for systems with large secondaries. To apply this formulation to (47171) 1999 TC36 we ignore component B when considering the evolution of the A1 and A2 orbit or treat A1+A2 as a point mass when consider the tidal evolution of component B. To the extent that these approximations violate the assumptions leading to equation 2, these results should be used with caution. The circularization timescale for component B is Tcirc~108 years. We note that with an eccentricity of 0.2949 the system is not circularized, although given the large uncertainties in the tidal circularization formulation, this may not be constraining. We find an extremely short circularization timescale of Tcirc~100 years for the inner pair (A1 and A2). Even allowing for order of magnitude uncertainties, the inner pair would be expected to have evolved to zero 13 ACCEPTED TO ICARUS - PREPRINT eccentricity in the absence of other perturbers. The non-zero eccentricity of the A1 and A2 pair in our orbit solution points to complex dynamical interactions between the three components. Formation 7.6. The formation of multiples can be considered a special case of the more general problem of how to form binaries. For the Kuiper Belt several possible modes of formation have been proposed including collision (Canup 2005) and gravitational capture (Goldreich et al. 2002). We consider each in the following sections. Additionally, Nesvorny (2008) has suggested that TNBs may have formed directly during the gravitational collapse of the gaseous protoplanetary disk when the excess of angular momentum prevented the agglomeration of all available mass into solitary objects. The details of this model are not yet available. 7.6.1. Collision Two other multiple systems are known in the Kuiper Belt. The Pluto system consists of two large binary components, Pluto and Charon, with two more small satellites, Nix and Hydra. Nix and Hyrda orbit near the 4:1 and 6:1 mean motion resonances (formally 3.991±0.007:1 and 6.064±0.006:1), respectively and share an orbit plane with Charon (Tholen et al. 2008). These details are consistent with formation by a giant collision and reaccretion in a disk (Ward & Canup, 2006; Lithwick & Wu, 2008). It has to be noted, however, that this consistency does not constitute proof that this system arose from a collision (e.g. Goldreich et al. 2002). Haumea and its two small satellites Hi’iaka and Namaka are members of a family of collision fragments identified by their strong water ice absorption features (Brown et al. 2007; Ragozzine & Brown 2007). Dynamical evolution models find that such a family could not have been created in the massive primordial belt because the coherence of the family members would not have survived to current times. Levison et al. (2008) explored the formation of a family resulting from the collision of two scattered disk objects and found the existence of one such family consistent with their results. It seems nearly certain that this triple system arose from a low probability collision (Levison et al. 2008; Schlichting & Sari 2009). The (47171) 1999 TC36 triple system differs significantly in its architecture compared to Pluto and Haumea (Figure 6). The components are more similarly sized and there is an order of magnitude difference in their semi-major axes. The angular momentum of the (47171) 1999 TC36 system is substantially higher than for the Pluto and Haumea systems (Figure 7) and appears to be more comparable to the larger population of TNBs, as discussed in Section 7.3. 7.6.2. Gravitational capture Gravitational capture by multi-body events in a dense planetesimal disk (Goldreich et al. 2002) produces binaries with high angular momenta and with an increasing fraction at small separations as is observed (Noll et al. 2008). Equal-sized components are preferred in at least some capture models (Astakhov et al.2005, Lee et al. 2007), again in accord with observations. Based on these agreements between model and observation, we have suggested that most TNBs are formed by capture (Noll et al. 2008). Goldreich et al. (2002) specifically address the formation of multiples and predict that such systems should exist as a natural result of hardening caused by dynamical friction and ongoing capture events. For (47171) 1999 TC36 the picture is an initial binary system in the process of hardening, but securely captured, encountering another pair of objects (or a single 14 ACCEPTED TO ICARUS - PREPRINT similarly sized object in a sea of small bodies) resulting in a second component being captured and likewise hardening. Successive capture events will initially have an orbit plane oriented randomly to a preexisting pair. However, it is possible that initially different orbit planes will evolve. For example, a system that starts off with near-orthogonal orbit planes would experience strong Kozai resonances that would result in episodes of high eccentricity. Such interaction could lead to tidal damping, coalescence or perhaps loss of the component. If non-coplanar orbits are more easily disrupted, the existence of a nearly coplanar triple like (47171) 1999 TC36 may simply be a selection effect. 7.7. Are there more triples and how do you find them? Finally, we ask, how do we find additional multiples in the Kuiper Belt, assuming that such systems exist? It has been previously shown (Kern & Elliot 2006; Noll et al. 2008) that the separation distribution for TNBs increases with decreasing separation to the observation limit (0.05 arcseconds). At the current observation limits, direct imaging of close triples is unlikely if they are composed of three approximately equal sized components with two components close to contact, as is the case for (47171) 1999 TC36. Lightcurves of known binaries have the potential to reveal an additional unresolved component. This technique relies on the assumption that an unresolved pair will be tidally locked and have a lightcurve period equal to the orbit period. Typically, this will be longer than the range of rotational periods for small bodies (4-18 hrs, Lacerda & Luu, 2006). While such a method does not result in an unambiguous identification of multiples, it can identify candidates for higher resolution follow-up and provide statistical information. Ideally, a search for multiples would result from resolved lightcurves of both primary and secondary components using HST. Unresolved observations can also be used to identify multiples by fitting a complex lightcurve. However, the added complication of accounting for two variable components with different periods increases the demands on photometric precision. Indirect detection of binaries through lightcurve analysis was pioneered by Pravec et al. (2002) and is now the most productive method for identifying binaries in the Near Earth asteroid population (Pravec et al. 2006). In the transneptunian population, Sheppard & Jewitt (2004) indirectly identified a possible close binary from its lightcurve and from it estimated that ~15% of TNOs may be similarly close pairs. Alternatively, determination of a binary orbit with non-Keplerian motion may indicate an additional component in the system. While no systems have been identified by this method, both the Haumea (Ragozzine & Brown, 2009) and (47171) 1999 TC36 triple systems have non- Keplerian orbits. 8. CONCLUSIONS We present new compelling evidence that (47171) 1999 TC36 is a triple system. The component originally identified as the primary consists of two components that orbit each other with a period of TA1A2 = 1.9068±0.0001 days at a separation of aA1A2 = 867±11 km. Components 286−38 265−35 A1 and A2 have diameters of of dA1= +45 km, dA2= +41 km assuming all components share the same albedo of 7%, determined from Spitzer measurements (Stansberry et al. 2008). We have also re-determined the orbit of the more distant secondary, component B. We find an orbital period of TB = 50.302±0.001 days and a semimajor axis of aB = 7411±12 km in € € 15 ACCEPTED TO ICARUS - PREPRINT 139−18 approximate agreement with previous work. The diameter of component B is dB= +22 km for the same assumed albedo. From the component B orbit we find a system mass of Msys= 12.75±0.06 1018 kg. An orbital analysis of this system demonstrates that it precesses rapidly with ϖ completing a full 2π in approximately 50 years for all of the components. This rapid precession € limits our ability to find a unique Keplerian orbit solution (valid on the timescale of years) based on the current data. 542−211 Using the derived component diameters we find a system density of ρ= +317 kg m-3. A porosity of 46-73% is required for a range of material densities from 1000-2000 g/cm3. The relatively large uncertainty in the density is primarily due to the uncertainties in the measurements of the thermal emission by Spitzer (Stansberry et al. 2008). The system has a high specific angular momentum, J/J’=0.93 and is likely to have been formed through gravitational € capture. The (47171) 1999 TC36 system does not demonstrate a simple lightcurve over five nights of ground-based observations. The photometric data shows variability of 0.20 ± 0.04 magnitudes for the unresolved system. The HST photometry shows that component B varies by at least 0.17 ± 0.05 magnitudes and components A1 and A2 combined vary by at least 0.08 ± 0.03 magnitudes. Attempts to model the lightcurve of the unresolved system do not yield statistically significant results. ACKNOWLEDGMENTS We thank M. Buie for the use of his PSF fitting code in IDL (Grundy et al. 2009) and his advice on how to go about modifying it for the HRC images. We also thank Dr. Don McCarthy and students Nancy Thomas and Shae Hart for their contributions of the 61-inch data and Astronomy Camp for supporting their work. This work is based on observations made with the NASA/ESA Hubble Space Telescope. These observations are associated with program #9746. Analyses were supported by HST programs #11178 (PI. Grundy) and #11113 (PI. Noll). Support for these programs was provided by NASA through a grant from the Space Telescope Science Institute, which is operated by the Association of Universities for Research in Astronomy, Inc., under NASA contract NAS 5-26555. ACCEPTED TO ICARUS - PREPRINT 16 REFERENCES Astakhov, S. A., Lee, E., A. and Farrelly, D. 2005. Formation of Kuiper-belt binaries through multiple chaotic scattering encounters with low-mass intruders. Mon. Not. Roy. Astron. Soc. 360, 401-415. Barkume, K. M., Brown, M. E. and Schaller, E. L. 2008. Near-infrared spectra of Centaurs and Kuiper Belt object. Astron J. 135, 55-67. Benecchi, S. D., Noll, K. S., Grundy, W. M., Buie, M. W., Stephens, D. C., Levison, H. F. 2009. The correlated colors of Transneptunian binaries. Icarus 200, 292-303. Benner, L. A. M., Nolan, M. C., Ostro, S. J., Giorgini, J. D., Pray, D. P., Harris, A. W., Magri, C., Margot, J. 2006. Near-Earth Asteroid 2005 CR37: Radar images and photometry of a candidate contact binary. Icarus 182, 474-481. Betzler, A. S., Novaes, A. B., Celedon, J. H. Q. 2008. A Study of the Trinary NEA 2001 SN263. The Minor Planet Bulletin 35, 182-184. Brown, M. E. and Trujillo, C. A. 2002. (26308) 1998 SM165. IAUC 7807. Brown, M. E., van Dam, M. A., Bouchez, A. H., Le Mignant, D., Campbell, R. D., Chin, J. C. Y., Conrad, A., Hartman, S. K., Johansson, E. M., Lafon, R. E., Rabinowitz, D. L., Stomski, P. J., Jr., Summers, D. M., Trujillo, C. A., Wizinowich, P. L. 2006. Satellites of the Largest Kuiper Belt Objects. Astrophysical J. 639, 43-46. Brown, M. E., Barkume, K. M., Ragozzine, D. and Schaller, E. L. 2007. A collisional family of icy objects in the Kuiper belt. Nature 446, 294-296. Brozovic, M., et al. 2009. (136617) 1994 CC. IAUC 9053. Buie, M. W., Grundy, W. M., Young, E. F., Young, L. A., and Stern, S. A. 2006. Orbits and Photometry of Pluto's Satellites: Charon, S/2005 P1, and S/2005 P2. Astron J. 132, 290- 298. Burns, J.A., 1986. Some background about satellites. In: Burns, J.A., Matthews, M.S. (Eds.), Satellites. Univ. of Arizona Press, Tucson, AZ, pp. 1–38. Canup, R. M. 2005. A Giant Impact Origin of Pluto-Charon. Science 307, 546-550. Chiang E., Lithwick Y., Murray-Clay R., Buie M., Grundy W., and Holman M. 2007. A brief history of transneptunian space. In: B. Reipurth et al. (Eds.), Protostars and Planets V Univ. of Arizona Press, Tucson, pp. 895-911. Danby, J.M.A. 1988. Fundamentals of Celestial Mechanics, 2nd edition. Willmann-Bell, Inc. Richmond VA. Dobrovolskis, A. R., Peale, S. J. and Harris, A. W. 1997. Dynamics of the Pluto-Charon Binary. In: S. A. Stern and D. J. Tholen (Eds.), Pluto and Charon, Univ. of Arizona Press, Tucson, pp. 159-190. Dotto, E., Barucci, M. A., Boehnhardt, H., Romon, J., Doressoundiram, A., Peixinho, N., de Bergh, C., Lazzarin, M. 2003. Searching for water ice on (47171) 1999 TC36, 1998 SG35, and 2000 QC243: ESO large program on TNOs and centaurs. Icarus 162, 408- 414. Dressel, L., et al. 2007, "STIS Data Handbook", Version 5.0, (Baltimore: STScI). Goldreich, P., Lithwick, Y. and Sari R. 2002. Formation of Kuiper-belt binaries by dynamical friction and three-body encounters. Nature 420, 643-646. Goldreich P. and Soter S. 1966. Q in the solar system. Icarus 5, 375–389. ACCEPTED TO ICARUS - PREPRINT 17 Grundy, W. M., Stansberry, J. A., Noll, K. S., Stephens, D. C., Trilling, D. E., Kern, S. D., Spencer, J. R., Cruikshank, D. P., Levison, H. F. 2007. The orbit, mass, size, albedo, and density of (65489) Ceto/Phorcys: A tidally-evolved binary Centaur. Icarus 191, 286-297. Grundy, W. M., Noll, K. S., Virtanen, J., Muinonen, K., Kern, S. D., Stephens, D. C., Stansberry, J. A., Levison, H. F., Spencer, J. R. 2008. (42355) Typhon-Echidna: Scheduling Observations for Binary Orbit Determination. Icarus 197, 260-268. Grundy, W. M., Noll, K. S., Buie, M. W., Benecchi, S. D., Stephens, D. C., and Levison, H. F. 2009. Mutual Orbits and Masses of Six Transneptunian Binaries. Icarus 200, 627-635. Gunn, J. E., and Stryker, L. L. 1983. Stellar spectrophotometric atlas, wavelengths from 3130 to 10800 A. Astrophysical Journ. Supplement Series 52, 121-153. Hainaut, O. R. and Delsanti A. C. 2002. Colors of Minor Bodies in the Outer Solar System. A statistical analysis. Astronomy & Astrophysics 389, 641-664. Jacobson, S. and Margot, J. L. 2007. Colors of TNO Binaries and Evidence for a Triple System from HST Observations. BAAS 39, 5211. Kern, S. D., and Elliot, J. L. 2006. The Frequency of Binary Kuiper Belt Objects. Astrophysical Journ. Letters 643, 57. Krist, J. and Hook R. 2004. The Tiny Tim User's Guide: Version 6.3 (Baltimore: STScI). Lacerda, P. 2005. The Shapes and Spins of Kuiper Belt Objects. PhD. Thesis. Leiden Observatory, Leiden University, The Netherlands. Lacerda, P. and Luu, J. 2006. Analysis of the rotational properties of Kuiper belt objects. Astron. Journ. 131, 2314-2326. Landolt, A. U. 1992. UBVRI photometric standard stars in the magnitude range 11.5-16.0 around the celestial equator. Astron J. 104, 340-371, 436-491. Leone, G., Paolicchi, P., Farinella, P., and Zappala, V. 1984. Equilibrium models of binary asteroids. Astronomy & Astrophysics 140, 265-272. Lee, E. A., Astakhov, S. A., and Farrelly D. 2007. Production of trans-Neptunian binaries through chaos-assisted capture. Mon. Not. Roy. Astron. Soc.379, 229-246. Levasseur-Regourd, A. C., Hadamcik, E., Desvoivres, E., and Lasue, J. 2009. Probing the internal structure of the nuclei of comets. Planetary and Space Science 57, 221-228. Levison, H. F., Morbidelli, A., Vokrouhlický, D., Bottke, W. F. 2008. On a Scattered-Disk Origin for the 2003 EL61 Collisional Family—An Example of the Importance of Collisions on the Dynamics of Small Bodies. Astron. J. 136, 1079-1088. Lithwick, Y. and Wu, Y. 2008. On the origin of Pluto’s minor moons, Nix ad Hydra. Submitted to ApJ, AstroPH axXiv:0802.2951v1. Marchis, F., Hestroffer, D., Descamps, P., Berthier, J., Bouchez, A.H., Campbell, R.D., Chin, J.C.Y., van Dam, M.A., Hartman, S.K., Johansson, E.M., Lafon, R.E., Le Mignant, D., de Pater, I., Stomski, P.J., Summers, D.M., Vachier, F., Wizinovich, P.L., Wong, M.H., 2006. A low density of 0.8gcm3 fortheTrojanbinaryaster-oid 617 Patroclus. Nature 439, 565–567. Marchis, F., Descamps, P., Berthier, J., and Emery, J. P. 2008. (216) Kleopatra. IAUC 8980. Marchis, F., Pollock, J., Pravec, P., Baek, M., Greene, J., Hutton, L., Descamps, P., Reichart, D. E., Ivarsen, K. M., Crain, J. A., Nysewander, M. C., Lacluyze, A. P., Haislip, J. B., and Harvey, J. S. 2008. (3749) Balam. IAUC 8928. Marchis, F., Descamps, P., Hestroffer, D., and Berthier, J. 2005. Discovery of the triple asteroidal system 87 Sylvia. Nature 436, 822. ACCEPTED TO ICARUS - PREPRINT 18 Marchis, F., Baek, M., Descamps, P., Berthier, J., Hestroffer, D., and Vachier F. 2007. (54) Eugenia. IAUC 8817. Margot, J. L., Brown, M. E., Trujillo, C. A. Sari R. and Stansberry J. A. 2005. Kuiper Belt Binaries: Masses, Colors and a Density. BAAS 37, 737. McKinnon, W.B., Lunine, J.I., Banfield, D. 1995. Origin and evolution of Triton. In: Cruikshank, D.P. (Ed.), Neptune and Triton. Univ. of Arizona Press, Tucson, AZ, pp. 807–877. Mueller, M., Marchis, F., Emery, J. P., Harris, A. W., Mottola, S., Hestroffer, D., Berthier, J., and di Martino, M. 2009. Eclipsing Binary Trojan Asteroid Patroclus: Thermal Inertia from Spitzer Observations. Icarus, in press: arXiv:0908.4198. Nesvorny, D. 2008. Formation of Kuiper Belt Binaries. BAAS 40, 464. Nolan, M. C., et al. 2008. (153591) 2001 SN263. IAUC 8921. Noll, K. S., Grundy, W. M., Chiang, E. I., Margot, J. L., and Kern, S. D. 2008a. Binaries in the Kuiper Belt. In: M. A. Barucci, H. Boehnhardt, D. P. Cruikshank, and A. Morbidelli (Eds.), The Solar System Beyond Neptune, Univ. of Arizona Press, Tucson, pp. 345-364. Ortiz, J. L., Gutierrez, P. J., Casanova, V., and Sota, A., 2003. A study of short term rotational variability in TNOs and Centaurs from Sierra Nevada Observatory. Astronomy & Astrophysics 407, 1149-1155. Peixinho, N., Doressoundiram, A., and Romon-Martin. J., 2002. Visible-IR colors and lightcurve analysis of two bright TNOs: 1999 TC36 and 1998 SN165. New Astronomy 7, 359-367. Pavlovsky, C. 2006. ACS Data Handbook, Version 5.0 (Baltimore: Space Telescope Science Institute). Pravec, P., L. Sarounova, M. D. Hicks, D. L. Rabinowitz, A. Wolf, P. Scheirich, and Y. N. Krugly 2002. Two periods of 1999 HF1 - Another binary NEA candidate. Icarus 158, 276-280. Pravec, P. et al. 2006. Photometric survey of binary near-Earth asteroids. Icarus 181, 63-93. Press, W.H., Teukolsky, S.A., Vetterling, W.T. and Flannery B.P. 1992. Numerical Recipes in C. 2nd Edition. Cambridge University Press, New York. Rabinowitz, D. L., Barkume, K., Brown, M. E., Roe, H., Schwartz, M., Tourtellotte, S., and Trujillo, C. 2006. Photometric Observations Constraining the Size, Shape, and Albedo of 2003 EL61, a Rapidly Rotating, Pluto-sized Object in the Kuiper Belt. Astrophysical Journ. 639, 1238-1251. Rabinowitz, D. L., Schaefer, B. E., and Tourtellotte, S. W. 2007. The Diverse Solar Phase Curves of Distant Icy Bodies. I. Photometric Observations of 18 Trans-Neptunian Objects, 7 Centaurs, and Nereid. Astron. J. 133, 26-43. Ragozzine, D. and Brown, M. E. 2007. Candidate members and age estimates of the family of Kuiper belt object 2003 EL61. Astron J. 134, 2160-2167. Ragozzine, D. and Brown, M. E. 2009. Orbits and Masses of the Satellites of the Dwarf Planet Haumea (2003 EL61). Astron J. 237, 4766-4776. Richardson, D. C., and Walsh, K. J. 2006. Binary Minor Planets. Ann. Rev. Earth Planet. Sci. 34, 47-81. Schlichting, H. E., and Sari, R. 2009. The creation of Haumea’s collisional family. Astrophysical Journ. 700, 1242-1246. Sheppard, S. S., and Jewitt, D. 2004. Extreme Kuiper Belt Object 2001 QG298 and the Fraction of Contact Binaries. Astron J. 127, 3023-3033. ACCEPTED TO ICARUS - PREPRINT 19 Spencer, J.R., Stansberry, J.A., Grundy, W.M., Noll, K.S. 2006. A low density for binary Kuiper belt object (26308) 1998 SM165 . Bull. Am. Astron. Soc. 38, Abstract 546. Stansberry, J. A., Grundy, W. M., Margot, J. L., Cruikshank, D. P, Emery, .J. P., Rieke, G. H. , and Trilling D. E. 2006. The Albedo, Size and Density of Binary Kuiper Belt Object (47171) 1999 TC36. Astrophysical Journ. 643, 556-566. Stansberry, J., Grundy, W., Brown, M., Cruikshank, D., Spencer, J., Trilling D., and Margot, J. L. 2008. Physical properties of Kuiper Belt and Centaur objects: Constraints from the Spitzer Space Telescope. In: M. A. Barucci, H. Boehnhardt, D. P. Cruikshank, and A. Morbidelli (Eds.), The Solar System Beyond Neptune, Univ. of Arizona Press, Tucson, pp. 161-179. Stern, S.A., Mutchler, M., Weaver, H. A., and Steffl, A. J., 2007. The positions, colors, and photometric variability of Pluto’s small satellites from HST observations 2005–2006. Lunar Planet. Sci. XXXVIII, 722–723. Tholen, D. J., Buie, M. W., Grundy, W. M., and Elliott, G. T. 2008. Masses of Nix and Hydra. Astron J. 135, 777-784. Trujillo, C. A. and Brown, M. E. 2002. IAUC 7787. Ward, W. R. and Canup, R. M. 2006. Forced Resonant Migration of Pluto’s Outer Satellites by Charon. Nature 313, 1107-1109. Weaver, H. A., et al. 2006. Discovery of two new satellites of Pluto. Nature 439, 943-945. ACCEPTED TO ICARUS - PREPRINT 20 FIGURES Figure 1. Sample fitting results for (47171) 1999 TC36 in both F606W (panel A) and F814W (panel B) filters and (26308) 1998 SM165 (panel C). Each triplet of images from left to right includes: observed data, PSF model and residual image (observed-model) with the left row showing the results of fitting the object with a single component (‘X’ marks the best fit location) and the right fitting it with two components. Numbers (1, 2) are plotted at the pixel location where the position and flux of the components minimize the χ2 residual. The χ2 residual for the binary fit of (47171) 1999 TC36 is a factor of 2-4 times better than that of the single fit, independent of filter. For (26308) 1998 SM165 the binary fit results in two components with identical locations with a nearly identical χ2 to the single fit. The image range is 13000 counts. ACCEPTED TO ICARUS - PREPRINT 21 Figure 2. Sample results of binary fits. The images are ordered by time. From left to right, the panels are: observed data, Tiny Tim model and residual image. Numbers (1, 2) are plotted at the pixel location where the position and flux of the component minimize the χ2 residual. The orientation of all the images is similar, but not identical, so a North-East indicator has been added for each visit. In most cases, 1 corresponds to the brighter of the two components. ACCEPTED TO ICARUS - PREPRINT 22 Figure 3. Component orbits. Component B has a period of 50.302±0.001 days and component A2 has a period of 1.9068±0.0001 days (shown to scale and blown up for visibility). The panel with both orbits (lower left) is plotted relative to the A1-A2 barycenter, while the zoomed panel (upper right) is plotted relative to A1 (this conserves the correct scaling of the orbits and uncertainties). Each black point with error bars (often difficult to discern) is an average of the points collected during a single HST orbit. The model is a weighted fit to the points, and the blue lines drawn between the black data points and the red model points indicate the residuals at the time of observation. A filled black circle represents A1 in the upper right panel and the A1-A2 barycenter in the lower left panel. North is up, and East is to the left. ACCEPTED TO ICARUS - PREPRINT 23 Figure 4. Resolved variability for the inner pair (top) and component B (bottom). The photometry is geometrically corrected and plotted by HST visit, on a log scale to facilitate better visibility. Time zero is defined as the time of the first HST observation (offset by 0.01 days since log 0 is undefined) and average values for each visit are plotted in red. The top panel of each plot zooms in on each visit to demonstrate the scatter in the points, which are used within each visit to calculate the uncertainty on the average values. The unresolved component A1+A2 has a peak-to- peak variation of 0.08±0.03 magnitudes and component B has a peak-to-peak amplitude of 0.17±0.05 magnitudes. ACCEPTED TO ICARUS - PREPRINT 24 Figure 5. Ground-based lightcurve observations of (47171) 1999 TC36. Data were collected from the Mt. Bigelow 61-inch and CTIO 4-m telescopes over 5 nights. The data have been plotted so that t0 is the time of the first observation. The 61-inch data were obtained using the Mont4k camera and the CTIO data were obtained using the Mosaic camera. The 61-inch data are normalized to the Sloan r’ filter and the uncertainties in the CTIO data are about the size of the data points. Attempts to model the data with a lightcurve do not yield any statistically significant results. The 61-inch data were contributed by Dr. Don McCarthy and students Nancy Thomas, Shae Hart of the University of Arizona Astronomy Camp. ACCEPTED TO ICARUS - PREPRINT 25 Figure 6. Comparison of Solar System multiples. All eight known multiples (n>2) in the solar system’s small body populations are represented in this plot showing the relative sizes and separations of the components. The semimajor axis relative to the system barycenter, a, scaled to the Hill radius, RHill, is plotted on the x-axis with a logarithmic scale. The relative component mass is plotted on the y-axis, also with a logarithmic scale, as the cube of the ratio of the component radius, r, to the cube of the effective system radius, Reff where Reff is the radius of a sphere having the same volume as all of the components. Primaries were forced to have a minimum a/RHill of 10-4 and offset for clarity. Components falling in the stippled area have less than 1% of the volume (and mass) of a body having the effective system radius. (47171) 1999 TC36 stands out from the other TNO triple systems in that all three components contribute a significant fraction of the total system mass. ACCEPTED TO ICARUS - PREPRINT 26 Figure 7. Comparison of angular momenta for Solar System multiples. Normalized angular momentum plotted vs. semi-major axis scaled by the Hill radius, RHill, of the system for all binaries and triples in the Kuiper Belt with determined orbits (Noll et al. 2008; Grundy et al. 2009) and all other solar system triple systems. Where rotational periods are known, we use them, where they are unknown we assume a period of 8 hours. Dobrovolskis et al. 1997 find that objects with J/J’ > 0.39 cannot have formed through rotational fission. Canup (2005) suggest that such systems are likely to be formed by a single catastrophic collision, however, if J/J’ > 0.8 additional considerations exist and formation by collision is limited. Systems with J/J’>0.8, including (47171) 1999 TC36, were most likely to formed through three-body gravitational interactions/capture as suggested by Goldreich et al (2002) (Chiang et al. 2007) or via the Nesvorny (2008) formalism. ACCEPTED TO ICARUS - PREPRINT 27 Figure 8. Orbit evolution. The evolution of our nominal orbit (Table 4) for both components is displayed. The orbit is stable and ϖ precesses through a full 2π radians in roughly 50 years. ACCEPTED TO ICARUS - PREPRINT 28 TABLE 1. OBSERVATIONS Instrument Comments Site UT Date R(AU) ∆(AU) PA(°)a TimeL (min)b STIS not tracked 2001-12-08 31.386 31.147 1.749 259.05 HST not tracked STIS HST 1.756 259.18 2001-12-09 31.385 31.316 — ACS HST 2003-06-30 31.205 31.165 1.868 259.19 ACS — 2003-07-09 31.203 31.013 1.842 257.93 HST — ACS HST 1.695 255.79 2003-07-25 31.198 30.756 ACS — 2003-08-23 31.190 30.380 1.134 252.66 HST ACS — 2004-05-26 31.111 31.661 1.551 263.32 HST ACS — 2005-08-12 30.996 30.363 1.482 252.52 HST — 1.885 256.82 2006-07-07 30.919 30.880 ACS HST 0.531 247.84 Bigelow Mont4k 2008-09-26 30.763 29.800 Photometric 2008-09-27 30.763 29.796 0.504 247.81 Bigelow Mont4k Photometric Photometric CTIO Mosaic 0.428 247.73 2008-09-30 30.763 29.787 2008-10-01 30.762 29.784 0.404 247.70 CTIO Mosaic Photometric 2008-10-02 30.762 29.782 0.381 247.68 CTIO Mosaic Cirrus a solar phase angle b light travel time TABLE 2. ASTROMETRY ∆Dec_Bc ∆RA_Bc 0.0019±0.0088 0.3812±0.0084 0.0385±0.0114 0.3750±0.0082 0.2715±0.0009 0.1839±0.0005 -0.1081±0.0007 0.2278±0.0003 -0.1576±0.0014 -0.0296±0.0003 0.3146±0.0015 0.0447±0.0014 -0.1431±0.0005 0.1474±0.0013 -0.2734±0.0021 -0.0249±0.0031 0.2583±0.0019 0.2702±0.0019 Decb -08:02:43.9 -08:02:32.5 -05:11:28.2 -05:13:14.5 -05:18:29.9 -05:33:53.5 -04:21:34.9 -03:36:42.1 -02:27:58.6 RAa Hour Date 2001/12/08 21.71786 00:10:33.41 2001/12/09 07.34166 00:10:32.99 2003/06/30 01.70880 00:36:49.57 2003/07/09 19.41372 00:37:00.11 2003/07/25 17.93292 00:36:53.80 2003/08/23 13.35984 00:35:33.71 2004/05/26 18.32280 00:43:06.70 2005/08/12 23.45220 00:53:20.19 2006/07/06 12.96780 01:02:14.64 a HH:MM:SS.SS b DD:MM:SS.S c arcseconds ∆RA_A2c -0.0247±0.0182 0.0324±0.0113 -0.0238±0.0008 -0.0210±0.0007 0.0236±0.0013 0.0228±0.0025 -0.0192±0.0018 0.0223±0.0028 -0.0174±0.0011 ∆Dec_A2c 0.0367±0.0124 0.0244±0.0218 -0.0084±0.0010 0.0154±0.0008 0.0040±0.0017 -0.0214±0.0021 -0.0241±0.0012 0.0167±0.0028 0.0258±0.0004 ACCEPTED TO ICARUS - PREPRINT 29 TABLE 3. A1,A2 MODEL FITTING RESULTS f2/f1 Julian date Rootname Δya Δxa 2 A χν 2452820.55979 0.162 -1.031 0.86 4.79 j8rl05abq 3.07 2452820.56740 0.174 -1.025 0.83 j8rl05acq 2452820.57500 0.184 -0.949 0.76 2.68 j8rl05adq 2452820.58260 0.227 -1.003 0.65 2.88 j8rl05aeq € € 2452830.29750 0.720 -0.749 1.07 3.05 j8rl06x0q 3.15 2452830.30510 0.743 -0.714 0.97 j8rl06x1q 2452830.31271 0.749 -0.685 0.92 3.48 j8rl06x2q 2452830.32031 0.768 -0.668 0.93 3.13 j8rl06x3q 2452846.23580 0.151 0.907 1.02 3.04 j8rl07h8q 3.06 2452846.24340 0.083 0.991 0.89 j8rl07h9q 2452846.25101 0.031 1.013 0.87 2.48 j8rl07haq 2452846.25861 0.012 0.914 1.08 2.29 j8rl07hbq 5.76 2452875.04526 -0.785 0.718 0.72 j8rl08b9qb 2452875.05286 -0.864 0.822 0.65 4.60 j8rl08baq 2452875.06046 -0.882 0.912 0.75 4.34 j8rl08bbq 2452875.06807 -0.992 0.926 0.71 5.14 j8rl08bcq 4.35 2453152.25204 -0.508 -1.057 1.23 j8rl09p9q 2453152.25965 -0.458 -1.071 1.03 3.97 j8rl09paq 2453152.26725 -0.386 -1.103 1.05 3.57 j8rl09pbq 4.05 2453152.27485 -0.369 -1.176 1.25 j8rl09pcq 2453595.46545 0.682 0.826 0.87 4.08 j8rl19akq 2453595.47306 0.707 0.953 2.30 5.83 j8rl10alqb 2453595.48129 0.547 0.720 0.90 2.10 j8rl10amq 2.21 2453595.48889 0.501 0.784 1.55 j8rl10anq 2453923.02892 1.066 -0.381 1.34 4.72 j8rl04r8q 2453923.03652 1.107 -0.412 1.03 5.13 j8rl04r9q 2453923.04413 1.075 -0.361 1.01 4.44 j8rl04raq 3.99 2453923.05173 1.050 -0.316 1.07 j8rl04rbq a measured in pixels cosmic ray poorly placed, fitting difficult b 2 A1,A2 χν 2.76 1.01 1.31 1.20 0.94 1.31 1.99 1.63 1.62 1.79 1.23 1.44 3.86 1.33 1.10 1.24 1.17 1.47 1.22 1.46 1.06 3.56 1.36 1.34 1.86 1.69 1.53 1.61 ACCEPTED TO ICARUS - PREPRINT 30 TABLE 4. ORBITAL ELEMENTS AND DERIVED PARAMETERS A2 Solutionc,d B Solutiond Parameter Fitted orbital elements 50.302±0.001 1.9068±0.0001 P Period (days) a 867±11 7411±12 Semi-major axis (km) e 0.101±0.006 0.2949±0.0009 Eccentricity 79.3±0.2 88.9±0.6 Inclination (°) i ε 184.4±1.6 281.1±0.3 Mean longitudea at epochb (°) Longitude of Ascending Nodea (°) Ω 330.0±1.0 325.2±0.1 ϖ Longitude of periapsisa (°) 47.7±6.3 292.1±0.2 Derived Parameters System Mass (x1018 kg) Msys 14.20±0.05 12.75±0.06 α Orbit pole right ascensiona (°) 240.055±1.149 236.837±0.162 δ Orbit pole declinationa (°) 11.167±0.175 1.023±0.389 a Referenced to J2000 equatorial frame. b The epoch is Julian date 2453880 (2006 May 24 12:00 UT). c excludes the STIS observations as discussed in Section 4.2 d Best fits to the data. The chi-square values are relatively large because precession is not included in our model and the data span a long enough timespan that precession is observed. ACCEPTED TO ICARUS - PREPRINT 31 Ref (1) (2) (3) (4) (5) (6) (7) (8) (9) TABLE 5. MULTIPLE SYSTEMS ρ System Regiona To P2 P1 d2 d1 a1 a2 d0 Msys a (kg m-3) (hr) (days) (days) (km) (km) (km) (km) (km) (kg) (AU) 0.83 0.2 0.1 0.05 0.65 1.2 0.5 1000 2.39 1.64 1.4x1011 NEA (136617) 1994 CC 3.4 6.13 1.92 1.2 0.5 1000 4 17 2.8 1.98 1.2x1013 (153591) NEA 2001SN263 3 2.80 110 1.39 1.5 310 20 1200 7 2.23 1.3x1014 MBA (3749) Balam 2 12.7 5.7 4.76 6 1120 700 1184 214.6 2.72 5.8x1018 (45) MBA Eugenia 1.4 5.39 4.2 5 3 118 650 380 3500 2.79 2.3x1018 MBA (216) Kleopatra 5.18 3.64 1.38 18 7 1200 706 1356 287 3.49 1.5x1019 (87) MBA Sylvia 50.3 45.7c 1.90 135 582 259 281 7411 867 39.2 1.3x1019 KB (47171) 1999 TC36 25.49 153.3 6.38 88 1920 19571 49240 2412 1212 39.5 1.3x1022 Plutob KB 49.1 3.91 KB 34.7 320 160 3000 25657 49880 1436 43.1 4.4x1021 Haumea Note: In the column headings, subscripts 0, 1 and 2 denote the primary, 1st component, and 2nd component, respectively. a is semi-major axis, d is the diameter of the component, P is the orbital period and T is the rotational period of the primary. a NEA=Near Earth Asteroids, MBA = Main Belt Asteroids, KB = Kuiper Belt b fourth component a3=65210 km, d3=72 km, P3=38.85 days. c assumed to be synchronous with orbital period. References: (1) IAUC 9053; (2) Nolan et al. 2008, Betzler et al. 2008; (3) IAUC 8928, IAUC 7827, CBET 1297; (4) IAUC 8817 (5) IAUC 8980; (6) Marchis et al. 2005; (7) This work; (8) Tholen et al. 2008; (9) Ragozzine et al. 2009; Rabinowitz et al. 2006. Values in bold/italics are come from the Johnston “Asteroids with Satellites” archive (http://www.johnstonsarchive.net/astro/asteroidmoons.html). System mass is calculated from the values provided in the table. TABLE 6. PHOTOMETRY ∆mA1A2 Julian Date VisitID Filter NOBS MTOTAL Δ mAB 19.571±0.009 0.28±0.07 2.17±0.02 2452820.57120 05 F606W 4 0.0±0.0 2.16±0.04 19.546±0.013 4 F606W 2452830.30890 06 2452846.24720 07 F606W 4 19.520±0.011 0.0±0.0 2.20±0.03 2452875.05666 08 F606W 4 19.409±0.007 0.37±0.03 2.22±0.01 19.556±0.014 0.14±0.05 2.24±0.03 4 F606W 2453152.26345 09 2453595.46925 10 F606W 2 19.429±0.014 0.3±0.3 2.27±0.03 2453595.48509 10 F814W 2 18.388±0.014 0.2±0.2 2.30±0.03 19.488±0.002 0.11±0.07 2.36±0.12 4 F606W 2453923.04032 04 ACCEPTED TO ICARUS - PREPRINT 32
0903.0652
1
0903
2009-03-03T23:19:34
A Search for Multi-Planet Systems Using the Hobby-Eberly Telescope
[ "astro-ph.EP", "astro-ph.SR" ]
Extrasolar multiple-planet systems provide valuable opportunities for testing theories of planet formation and evolution. The architectures of the known multiple-planet systems demonstrate a fascinating level of diversity, which motivates the search for additional examples of such systems in order to better constrain their formation and dynamical histories. Here we describe a comprehensive investigation of 22 planetary systems in an effort to answer three questions: 1) Are there additional planets? 2) Where could additional planets reside in stable orbits? and 3) What limits can these observations place on such objects? We find no evidence for additional bodies in any of these systems; indeed, these new data do not support three previously announced planets (HD 20367b: Udry et al. 2003, HD 74156d: Bean et al. 2008, and 47 UMa c: Fischer et al. 2002). The dynamical simulations show that nearly all of the 22 systems have large regions in which additional planets could exist in stable orbits. The detection-limit computations indicate that this study is sensitive to close-in Neptune-mass planets for most of the systems targeted. We conclude with a discussion on the implications of these non-detections.
astro-ph.EP
astro-ph
A Search for Multi-Planet Systems Using the Hobby-Eberly Telescope1 Robert A. Wittenmyer2,3, Michael Endl2, William D. Cochran2, Harold F. Levison4, Gregory W. Henry5 [email protected] ABSTRACT Extrasolar multiple-planet systems provide valuable opportunities for testing the- ories of planet formation and evolution. The architectures of the known multiple- planet systems demonstrate a fascinating level of diversity, which motivates the search for additional examples of such systems in order to better constrain their formation and dynamical histories. Here we describe a comprehensive investiga- tion of 22 planetary systems in an effort to answer three questions: 1) Are there additional planets? 2) Where could additional planets reside in stable orbits? and 3) What limits can these observations place on such objects? We find no evidence for additional bodies in any of these systems; indeed, these new data do not support three previously announced planets (HD 20367b: Udry et al. 2003, HD 74156d: Bean et al. 2008, and 47 UMa c: Fischer et al. 2002). The dynamical simulations show that nearly all of the 22 systems have large regions in which additional planets could exist in stable orbits. The detection-limit computations indicate that this study is sensitive to close-in Neptune-mass planets for most of the systems targeted. We conclude with a discussion on the implications of these non-detections. Subject headings: stars: planetary systems -- extrasolar planets 9 0 0 2 r a M 3 . ] P E h p - o r t s a [ 1 v 2 5 6 0 . 3 0 9 0 : v i X r a 1Based on observations obtained with the Hobby-Eberly Telescope, which is a joint project of the Uni- versity of Texas at Austin, the Pennsylvania State University, Stanford University, Ludwig-Maximilians- Universitat Munchen, and Georg-August-Universitat Gottingen. 2McDonald Observatory, University of Texas at Austin, Austin, TX 78712 3Department of Astrophysics, School of Physics, University of NSW, 2052, Australia 4Department of Space Studies, Southwest Research Institute, Boulder, CO 80302 5Center of Excellence in Information Systems, Tennessee State University, 3500 John A. Merritt Blvd., Box 9501, Nashville, TN 37209, USA -- 2 -- 1. Introduction About 12% (N = 31) of known planetary systems contain more than one planet. Now that radial-velocity precision at the 1-2 m s−1 level is being achieved by several planet search programs (Butler et al. 2006; Lovis et al. 2006), Neptune-mass planets are becoming de- tectable. Recent discoveries of "super-Earths" (m sin i < ∼ 10 M⊕) by the High-Accuracy Radial Velocity Planet Search (HARPS) instrument (Bouchy et al. 2008; Mayor et al. 2009; Udry et al. 2007; Bonfils et al. 2007) suggest that super-Earths may be common. The presence of close-in giant planets ("hot Jupiters") inferred by precision radial- velocity surveys has emphasized the importance of post-formational dynamical evolution processes such as planetary migration. The core-accretion model of planetary formation (Lissauer 1995; Pollack et al. 1996) posits that rocky cores form in the outer regions of the protoplanetary disk and experience runaway gas accretion once they reach a mass of ∼10 Earth masses. These giant planets then migrate inward to become hot Jupiters (Bodenheimer et al. 2000). Alternatively, the disk-instability model suggests that such plan- ets form by direct gravitational collapse of the protoplanetary disk (Boss 1995, 1998). Multi- planet systems can be formed by this method (Boss 2003), though subsequent evolution can easily eject planets, resulting in a wide variety of system end-states (Levison et al. 1998). The discovery of additional multi-planet systems will provide valuable added constraints to these two models of planet formation. Trilling et al. (1998) have proposed that gas giant planets migrating inward can overflow their Roche lobes and be stripped of their gaseous envelopes. Under the core-accretion model of planet formation, a Neptune-mass rocky core would then remain in a close orbit, and the detection of such objects would lend support to that theory. Alternatively, the nondetection of close-in, low-mass (mp < 15 M⊕) planets would tend to favor the disk-instability model, in which gas giant planets have no solid cores. Hence, an intensive effort to characterize the population of detectable planets around nearby stars will be extremely valuable for understanding the processes of planet formation and evolution. The architectures of multi-planet systems can shed light on their formation and dy- namical history. Chatterjee et al. (2008) performed simulations of systems with three giant planets and found that at least one planet would be ejected before the system stabilised. When two planets remained (80% of cases), their median eccentricities were e ∼0.4. Sim- ilarly, randomly generated planetary systems simulated by Juri´c & Tremaine (2008) typi- cally retained 2-3 giant planets after 108 yr. That all five planets (Fischer et al. 2008) in the 55 Cancri system have relatively low eccentricities (e < 0.2) suggests that systems with inactive dynamical histories (i.e. free of major perturbation events) may be able to retain several giant planets in nearly circular orbits. -- 3 -- The final configuration of a planetary system is dependent on the post-formation mi- gration and dynamical interaction processes. Mandell et al. (2007) showed that the mi- gration of a Jupiter-mass planet through a disk of planetesimals can result in the forma- tion of an interior terrestrial-mass planet. Simulations of known multi-planet systems by Barnes & Quinn (2004) and Barnes & Raymond (2004) suggest that planetary systems are "packed" -- that is, they contain the maximum number of planets that is dynamically possi- ble. Barnes & Raymond (2004) investigated the dynamically stable regions of the HD 74156 system. Those authors used the results to predict that an additional planet, between planets b and c, could be present. The detection by Bean et al. (2008) of such an object lends sup- port to the "packed planetary systems" hypothesis (Barnes et al. 2008), which would imply that multiple-planet systems are common. However, our own results (see § 3) do not support this hypothesis. A series of papers by Ida & Lin (Ida & Lin 2004a,b) predicts a paucity of planets of 10-100 Earth masses within ∼ 1 AU (the "planet desert"). Their core-accretion simulations also predict an abundance of close-in (a < ∼ 0.1 AU) planets with masses below about 10 M⊕. Ida and Lin further suggest that the distribution of planetary mass vs. semimajor axis will constrain the dominant formation processes of planets. In a subsequent paper, Ida & Lin (2008) show that the frequency of giant planets depends sensitively on the Type I migration rate, which must be slowed by a factor C1 ∼0.03-0.1 in order to reproduce the distribution of detected planets. In this work, we describe an intensive three-year radial-velocity campaign to search for additional planets in known planetary systems (§ 2). Section 3 gives the results of the orbit fits and the search for new planets, with discussion about a few of the interesting systems. Section 4 describes the dynamical simulations used to determine the regions in each system where additional planets could reside in stable orbits. The detection limits, which determine the sensitivity of this survey, are presented in § 5. Finally, Section 6 assesses the impact of these new data and analyses on the theories of planet formation and the population-level statistics of extrasolar planets. This work thus presents a three-fold approach to the question of planetary system architecture: 1) Are additional planets present in these known planetary systems? 2) Where could additional objects reside in stable orbits? 3) What limits can be placed on such objects? 2. Observational Data Twenty-two targets were chosen for this project from the list of ∼150 planet hosts known in 2004 September. A majority of the observational data were obtained at McDonald Ob- -- 4 -- servatory with the 9.2 m Hobby-Eberly Telescope (HET: Ramsey et al. 1998) using its High Resolution Spectrograph (HRS) (Tull 1998). The targets were selected according to the fol- lowing criteria: 1) HET observability, with declination between -11o and +72o, and 2) Either a long-period (P > ∼ 1 yr) planet such that inner planets may be dynamically stable, or a very short-period (P < ∼ 10 days) hot Jupiter which would allow for previously undetected outer planets, and 3) The orbital solution for the known planet in each system has RV residuals of 10-20 m s−1, so that an additional planet may be present but undetected. The targets and their stellar parameters are listed in Table 1. Except where noted, masses are obtained from Takeda et al. (2007), [Fe/H], Tef f , and V sin i from Valenti & Fischer (2005), and the chromospheric emission ratio log R′ HK (Noyes et al. 1984) computed from measurements of the Ca II S-index obtained with the 2.7m telescope using the techniques developed by Paulson et al. (2002). The uncertainties on the stellar masses given in Takeda et al. (2007) are asymmetric about the central value; for the purposes of Table 1 and the determination of planetary parameters, the adopted stellar mass uncertainty was taken to be the larger of the two. All of the HET observations for this program were performed at a spectral resolution of 60,000, with the 316 gr/mm cross-disperser and a central wavelength of 5936A. An iodine cell temperature-controlled at 70◦ C was used as the velocity metric (Marcy & Butler 1992). This setup, identical to that used for the ongoing planet search program (Cochran et al. 2004; Endl et al. 2008), places the iodine region (∼5000-6000 A) almost entirely onto the blue CCD, which is cosmetically superior to the red CCD. For each target, an iodine-free template spectrum was obtained near the beginning of the first season in which is was observable. We determined precise radial velocities following the general recipe outlined by Butler et al. (1996), using an advanced version of our own code "Austral" (Endl et al. 2000). We observed each target with the HET in queue mode using a random observing interval of 2-10 days between visits. Each visit consisted of one spectrum, except for seven bright targets (HD 3651, HD 19994, HD 38529, HD 74156, 47 UMa, HD 128311, HD 136118) for which 3 consecutive spectra were obtained in each visit. HET data consisting of multiple exposures per visit were binned using the weighted mean value of the velocities in each visit. We adopted the quadrature sum of the rms about the mean and the mean internal error as the the error bar of each binned point. This procedure was done for HD 3651, HD 19994, HD 38529, HD 74156, 47 UMa, HD 128311, and HD 136118. Targets were observed with the HET from 2004 December through 2007 November. During the three years of this study, supplemental observations were also made using the 2.7m Harlan J. Smith telescope at McDonald Observatory. All available published radial-velocity data were also gathered from the literature for the purpose of fitting orbits to the known planets. Those data are summarized in Table 2. All radial-velocity data obtained from McDonald Observatory are given in Tables 10-45. -- 5 -- 3. Refined Planetary System Parameters 3.1. Orbit Fitting Results Available published data were combined with velocities from the HET and the 2.7m to fit Keplerian orbits using GaussFit (Jefferys et al. 1987), which is a generalized least- squares program used here to solve a Keplerian radial-velocity orbit model. The GaussFit model has the ability to allow the offsets between data sets to be a free parameter. This is important because the radial velocities cited in published works, and those computed from HET and 2.7m data, are not absolute radial velocities, but rather are measured relative to the iodine-free stellar template. The Geneva planet-search group, however, makes use of a simultaneous thorium-argon calibration rather than an iodine absorption cell (Baranne et al. 1996). Each data set thus has an arbitrary zero-point offset which must be accounted for in the orbit-fitting procedure. The best-fit Keplerian orbital solutions and planetary parameters are shown in Table 3. A summary of the fit results for each individual data set is given in Table 4. In computing the planetary minimum mass M sin i and semimajor axis a, the stellar masses listed in Table 1 were used. The addition of a large amount of new data and the use of multiple independent data sets in fitting Keplerian orbits have generally improved the precision of the derived planetary parameters by a factor of 2-4 over the published results summarized in the Catalog of Nearby Exoplanets (Butler et al. 2006). In particular, the precision of the orbital periods have been improved by the addition of new data, due to the increased number of orbits now observed. Our parameters generally agree within 2σ of previously published estimates. In this section, we highlight interesting results from the combined fits. For each object, we searched for periodic signals in the residuals to the known planet's orbit using a Lomb-Scargle periodogram (Lomb 1976; Scargle 1982). To assess the statistical significance of those periods, the false alarm probabilities (FAP) were calculated using the bootstrap randomization method detailed by Kurster et al. (1997). The bootstrap method randomly shuffles the velocity observations while keeping the times of observation fixed. The periodogram of this shuffled data set is then computed and its highest peak recorded. In this way, we can determine the probability that a periodogram peak of a given power level will arise by chance, without making any assumptions about the error distribution of the data. All bootstrap FAP estimates result from 10000 such realizations. Those results are shown in Table 5. -- 6 -- HD 20367. A planet orbiting HD 20367 was first announced in a conference proceedings (Udry et al. 2003), but has not yet appeared in a refereed journal. The Geneva planet search group website1 lists the planet's period as 469.5 days, with an eccentricity of 0.32 and M sin i=1.17 MJup. Eighty-one observations of HD 20367 were obtained with the HET over three observing seasons, as well as 19 observations from the 2.7m, but period searches of these data give no indication of such a signal. Figure 1 shows the radial-velocity data from HET and the 2.7m telescopes, and the periodogram of those data. The Geneva group's solution has been overplotted. The highest peak, at 5.58 days, has a bootstrap FAP of 8.5%. The dominant periodicity of 5.58 days, which was evident early in the observation campaign, prompted a photometric investiga- tion to search for transits and to rule out stellar rotation. We obtained 132 observations of HD 20367 from 2006 September to 2007 January with the T10 0.8m automated photometric telescope (APT) at Fairborn Observatory in southern Arizona. The T10 APT and its pre- cision photometer are very similar to the T8 APT described in Henry (1999). The precision of a single observation is typically around 0.001 mag. The results indicate a stellar rotation period of 5.50±0.02 days, with a photometric amplitude of 0.0055±0.0003 mag (Figure 2). From these observations, we conclude that the 5.6-day radial-velocity periodicity is caused by starspots rotating into and out of view. This is consistent with the estimate of Prot =6 days reported by Wright et al. (2004), and the high level of chromospheric activity for this star (log R′ HK = −4.50). The literature contains conflicting age estimates for HD 20367: Holmberg et al. (2007) estimate an age of 4.4+1.6 −2.1 Gyr, whereas Wright et al. (2004) report an age of 0.9 Gyr. Based on the rapid rotation rate, and the high level of chromospheric emission, the younger age estimate is favored. The lack of any Keplerian signal in the 100 observations presented here leads us to conclude that there is not convincing evidence for the existence of HD 20367b. HD 74156. For HD 74156, we fit the two planets at 51 and 2473 days using ELODIE and CORALIE data from Naef et al. (2004), and 82 independent HET visits. This system warrants closer scrutiny in light of the report by Bean et al. (2008) of a third planet, with a period of 346 days and a radial-velocity semi-amplitude K = 10.5 m s−1. That result was obtained using the same HET spectra as considered in this work, but velocities were derived using an independent method described in Bean et al. (2007). Here, we further investigate the possibility of an additional planet in the HD 74156 system. Applying our orbit-fitting methods as described above to the velocities for HD 74156 given in Bean et al. (2008), a periodogram peak is evident near 346 days, and we obtain a three-planet Keplerian orbit fit 1http://obswww.unige.ch/∼udry/planet/hd20367.html -- 7 -- which is consistent with that of Bean et al. (2008). This indicates that the fitting method used here is not responsible for our non-detection of HD 74156d. It is possible that the HET velocities derived by Bean et al. (2008) are of superior quality to those presented here. However, the rms of the HET data about a two-planet fit reported by Bean et al. (2008) is 8.5 m s−1, whereas we obtain an rms of 8.3 m s−1 for those data. These results suggest that there is no significant difference in quality between the two extant sets of HET velocities for HD 74156. The uncertainties quoted by Bean et al. (2008) are generally smaller than ours by a factor of 2-3. We repeated the fitting procedure, reducing the HET uncertainties by a factor of 2 and 3, but there was no significant change in the residuals: no signal is evident at periods near 346 days. Since the total rms scatter about our two-planet fit is 11.5 m s−1, and the semi-amplitude of planet d is K = 10.5m s−1, it is possible that a third planetary signal may have been lost in the noise. To test this possibility, we performed the following Monte Carlo simulations. From each of the two data sets considered in the fits described here, we generated 1000 simulated sets of velocities consisting of three Keplerian signals plus a Gaussian noise term. This noise was equivalent to the mean uncertainty of each data set (ELODIE+CORALIE: 10.8 m s−1, HET: 8.3 m s−1) added in quadrature to a stellar jitter of 4 m s−1 (the jitter estimate used in Bean et al. 2008). The parameters of the three simulated planets were those from Bean et al. (2008). These simulated datasets retained the times of observation and the error bars of the originals. We then fit the simulated data with a two-planet model exactly as described above, and examined the residuals of the two-planet fit by the periodogram method, to determine whether the signal of planet d was recovered. The criteria for recovery were that the period of the second planet had to be detected correctly and with a FAP of less than 0.1%. This FAP was computed using the analytic FAP formula of Horne & Baliunas (1986). Of the 1000 trials, only 11 did not result in a successful recovery of the signal of the second planet. The correct period was recovered 995 times, and the FAP exceeded 0.1% only 6 times; the worst FAP was 0.3%. These results indicate that our method should have been able to detect the signal of HD 74156d, had it been present with the parameters given by Bean et al. (2008). In Bean et al. (2008), the iodine-free stellar template spectrum was obtained at a re- solving power of R =120,000, rather than the R =60,000 which is standard for targets in this paper. We obtained an R =120,000 template spectrum on 2007 Nov 12, but the velocities computed using this template resulted in a 2-planet fit with a slightly higher rms (HET: 8.9 m s−1) than the original R =60,000 template (HET: 8.3 m s−1). All analysis for HD 74156 in this paper refers to velocities obtained using the R =60,000 template. A periodogram of the residuals to our 2-planet fit is shown in the left panel of Figure 3, -- 8 -- and those residuals are phased to the 346.6 day period in the right panel. The window function (grey dotted line) has a broad peak near 346 days due to the 1-year observing window. The phase gaps (right panel) are expected since the trial period is close to 1 year. No clear Keplerian signal is evident despite the large number of data points (N = 177). We conclude from these data that there is not sufficient evidence for a third planet in the HD 74156 system. 47 UMa (=HD 95128). In Wittenmyer et al. (2007a), we performed these fits to an earlier set of data from McDonald Observatory. Those results did not provide convincing evidence for the outer planet reported by Fischer et al. (2002a) at P∼2594 days; rather, we obtained a best-fit 2-planet model with P2 ∼6900 days. Here we include an additional 14 epochs from HET, and the best-fit 2-planet model now calls for P2 ∼9660 days. As in previous attempts to fit a second planet, the parameters e2 and ω2 needed to be held fixed, at the values proposed by Fischer et al. (2002a): e2 = 0.005 and ω2 =127o. The rms about a single-planet model is 10.2m s−1, compared to 8.6m s−1 when a second planet is included. Considering the continued ambiguity in the parameters for a second planet, and the ever- lengthening period of such an object, we use the one-planet fit for all further analysis in this work. HD 114783. Vogt et al. (2002) reported the planet orbiting HD 114783, and recently, Wright et al. (2007) proposed an outer companion with a period of at least 8 yr. Here, we combine the Keck data given in Butler et al. (2006) with HET observations. A single-planet fit has a total rms of 6.25 m s−1 and χ2 ν=4.91, whereas a two-planet fit reduces the rms to 4.42 m s−1 and the χ2 ν to 1.81. The data considered in Wright et al. (2007) were of insufficient duration to establish a solution for the outer planet, but the combination of data allows for a Keplerian fit to converge. Although a 2-Keplerian model can be fit to these data, it is of limited utility: the outer planet has a 50% uncertainty in period (P2 = 5098 ± 2576 days). Our results support those of Wright et al. (2007), that a second object is likely present, although there is not yet a sufficient time baseline of observations to establish its nature. The 1-planet fit was used to derive the parameters given in Table 3, and was also used for the detection-limits determination in § 5. HD 128311. The inner planet (P ∼450 days) in the HD 128311 system was first dis- covered by Butler et al. (2003), who noted a linear trend in the residuals to the fit, as well as the extremely high activity level. Those authors estimated the stellar jitter at 30 m s−1, and expressed concern that the planetary signal may have its origin in the stellar veloc- ity jitter. Additional data proved that the inner planet was indeed real, and Vogt et al. (2005) reported a second planet at the 2:1 mean-motion resonance (MMR). They published a solution consisting of two superposed Keplerian orbits, noting that preliminary dynamical -- 9 -- tests showed the system to be unstable, and that the system was likely in a protected 2:1 resonance. Go´zdziewski & Konacki (2006), in their dynamical analysis of available radial- velocity data, suggested that the observed signal could be attributed to a 1:1 resonance, i.e. a pair of Trojan planets. In this work, we fit a double Keplerian model to the combined Keck and HET data. Convergence is achieved, with a total rms of 16.9 m s−1 about both data sets (Keck -- 15.8 m s−1, HET -- 17.9 m s−1). The residuals show a strong periodicity near 11.5 days, with bootstrap FAP less than 0.01%. Photometry of HD 128311 by G. Henry in Vogt et al. (2005) indicates a stellar rotation period of 11.53 days with a photometric amplitude of 0.03 mag. Hence, it is quite clear that the residual signal is caused by stellar rotation in this highly active star. HD 130322. HD 130322 is host to a hot Jupiter in a 10.7-day period, discovered with the CORALIE observations of Udry et al. (2000). Four data sets are available for this object: CORALIE (Udry et al. 2000), Keck (Butler et al. 2006), HET, and 2.7m. Fitting all four sets together results in a total rms of 14.8 m s−1, but removing the CORALIE data drops the rms to 9.3 m s−1. In addition to the large scatter about the fit, a highly significant periodicity remains at 35 days (FAP<0.01%), which vanishes when the CORALIE data are removed. Due to these irregularities, we elect to exclude those data from the fits. The precision of the derived orbital parameters is not significantly affected by this removal, since the CORALIE data span only 167 days. For all further analysis in this work, we refer to the fit which excluded the CORALIE data. As given in Table 5, a residual period is present at P ∼438 days (FAP=0.16%). However, the HET velocities obtained using a second iodine- free template spectrum show no such periodicity. Those results show a residual period at 2.518 days, with a bootstrap FAP of 0.35%. A second planet can be fitted at this shorter period, and preliminary dynamical tests show that it would remain stable for at least 107 yr; however, the disagreement between the two templates makes it imprudent for us to claim a detection at this time. 4. Dynamical Mapping With the increasing availability of computing power and planetary systems, many in- vestigators have undertaken N-body simulations of known planetary systems in an effort to characterise regions in which additional bodies could be found. Menou & Tabachnik (2003) performed a comprehensive test-particle analysis of 85 systems to determine the extent of the habitable zones in the presence of the known planet(s). Due to disruptions from the known giant planet's "zone of influence," they found that only one-fourth of the systems had dynamical habitability comparable to our own Solar system. In addition to test particles, -- 10 -- massive "test planets" have also been used to test observational claims for new planets and to probe known multiple-planet systems for additional regions of stability (Rivera & Lissauer 2000, 2001; Raymond & Barnes 2005; Rivera & Haghighipour 2007). Likewise, in this sec- tion we perform test-particle and massive-body simulations on the systems targeted by the intensive radial-velocity monitoring described in § 3. 4.1. Test Particle Simulations We performed test particle simulations using SWIFT2 (Levison & Duncan 1994) to in- vestigate the dynamical possibility of additional low-mass planets in each of the systems considered here. SWIFT is a numerical integration package which is designed to solve the equations of motion for gravitational interactions between massive bodies (star, planets) and massless test particles. Neptune-mass planets can be treated as test particles (1 Neptune mass = 0.054 MJup) since the exchange of angular momentum with jovian planets is small. We chose the regularized mixed-variable symplectic integrator (RMVS3) version of SWIFT for its ability to handle close approaches between massless, non-interacting test particles and planets. This version is most efficient when the gravitational interactions are dominated by a single body (the central star). A symplectic integrator has the advantage that errors in energy and angular momentum do not accumulate. Particles are removed if they are (1) closer than 1 Hill radius to the planet, (2) closer than 0.01 AU to the star, or (3) farther than 10 AU from the star. A planetary-mass object passing within 1 Hill radius of another planet, or within 0.01 AU (2 R⊙) of the star's barycenter, is unlikely to survive the encounter. Since the purpose of these simulations is to determine the regions in which additional planets could remain in stable orbits, we set the outer boundary at 10 AU because the current repository of radial-velocity data cannot detect objects at such distances. For each planetary system, 390 test particles were placed in initially circular orbits, spaced every 0.005 AU in the region between 0.05-2.0 AU. We have chosen to focus on this region because the duration of our high-precision HET data is currently 3-4 years for the objects in this study. The test particles were coplanar with the existing planet, which had the effect of confining the simulation to two dimensions. The initial orbital positions of the particles were randomly distributed in orbital phase with respect to the existing planets. The method used here are the same as Wittenmyer et al. (2007b), in which we performed test- particle simulations for six highly eccentric planetary systems. Input physical parameters (Table 3) for the known planet in each system were obtained from our Keplerian orbit fits 2SWIFT is publicly available at http://www.boulder.swri.edu/∼hal/swift.html. -- 11 -- combining published velocity data and new observations from McDonald Observatory. The planetary masses were taken to be their minimum values (sin i = 1). By choosing the minimum mass for the planets, the regions of dynamical stability shown by the test-particle results are larger. Since the system inclinations are almost certainly not edge-on, and hence the true planetary masses are higher, we expect the actual regions of stability to be smaller than shown here. The systems were integrated for 107 yr, following Barnes & Raymond (2004) and allowing completion of the computations in a reasonable time. We observed that nearly all of the test-particle removals occurred within the first 106 yr; after this time, the simulations had essentially stabilized to their final configurations. 4.2. Test Particle Results The results of the test-particle simulations are shown in Figures 4-13. The survival time of the test particles is plotted against their initial semimajor axis. Two systems targeted by the radial-velocity observations were not included in these simulations: HD 20367, because there is no evidence for a planet, and HD 128311, since the Keplerian orbit solution ob- tained in § 3.1 results in an unstable system. As shown in Figure 4, the short-period planet HD 3651b sweeps clean the region inside of about 0.5 AU. However, a small number of test particles remained in low-eccentricity orbits near the 1:3 and 2:1 mean-motion resonances (MMR). Since these regions lie within the orbital excursion of HD 3651b, these appear to be protected resonances. The eccentricity of the test particles in the region of the 1:3 MMR oscillated between 0.00 and 0.31 with a periodicity of about 1.2 × 105 yr, while those in the 2:1 resonance remained at e < ∼ 0.07 throughout the simulation. All particles beyond about 0.6 AU also remained in stable orbits, which is not surprising given the low mass of the planet. In simulations by Mandell et al. (2007) and Raymond et al. (2006), a migrating Jupiter-mass planet captured planetesimals into low-order resonances, and these accreted into terrestrial planets during the 200 Myr run. The architecture of the HD 3651 system, with a 0.2 MJup planet at 0.3 AU, is similar to the configuration modeled by Mandell et al. (2007). Given the stable regions evident near the 1:3 and 2:1 resonances for HD 3651b, it is possible that terrestrial-mass planets were captured into these regions during the migration process. The detection limits for HD 3651 (§ 5) complement the dynamics well, and the current data can place upper limits of 1-2 Neptune masses (17-34 Earth masses) on such objects. For most of the systems, the test-particle results give few surprises. Broad stable regions exist interior and exterior to HD 8574b, with the inner 0.47 AU retaining 100% of particles. For HD 10697 and HD 23596, particles remained in the inner 1.35 AU and 1.4 AU, respec- -- 12 -- tively. The HD 19994 system, shown in Figure 5, proved to be quite interesting. One would expect any particles in orbits which cross that of the planet to be removed straightaway, but a few particles remained near the 1:1 resonance with the planet, in the range 1.29-1.33 AU. Laughlin & Chambers (2002) investigated the possibility of planets in a 1:1 resonance, and concluded that such configurations are indeed possible, In the "eccentric resonance," one 1:1 configuration described by Laughlin & Chambers (2002), one planet is in a nearly circular orbit while the other is in a highly eccentric orbit. Though the orbits cross, the longitudes of pericenter are sufficiently different to avoid close encounters. In the HD 28185 system (Fig. 6), no stable regions exist exterior to the planet out to the maximum separation tested (a = 2.0 AU). Figure 7 shows the results for the HD 38529 and HD 40979 systems. There is a broad region of stability between the widely-separated planets HD 38529b and c, consistent with the results of Barnes & Raymond (2004). The outer planet does not fall within the range of Fig. 7, but has an orbital excursion of 2.43 -- 4.99 AU. For HD 74156, the recently-announced planet d (Bean et al. 2008) in a 346-day period between planets b and c, was not included in the simulation. Only those particles in a narrow strip near 1.25 AU survived the full 10 Myr; planet d would fall within the stable region. The 47 UMa system (Figure 10) included only the inner planet (a = 2.11 AU) for this experiment. The parameters of an outer planet are highly uncertain (Wittenmyer et al. 2007a; Naef et al. 2004), and such an object would be too distant to affect the inner 2 AU explored here. A large region interior to the planet is stable for the full duration, including the habitable zone. This result is consistent with that of Jones et al. (2001), who also found the 47 UMa habitable zone to be stable for an Earth-mass planet at 1 AU. With an M sin i of 6.9 MJup, HD 106252b clears out all particles outside of a ∼0.7 AU. For the HD 108874 system, no test particles survive between the two planets (Figure 11), but those in the innermost 0.3 AU remain stable. Particles interior to HD 114783b were stable to about a ∼0.7 AU. As expected for the HD 130322 hot-Jupiter system, all particles with a > 0.15 AU survived (Fig. 12). In the HD 178911B system (Fig. 13, some particles remained in the inner 0.1 AU despite the large mass (M sin i=6.95 MJup) and relative proximity (a = 0.34 AU) of the planet. 4.3. Massive Body Simulations Regions stable for massless test particles may not be stable for massive bodies. Alterna- tively, regions unstable for test particles may be able to host a massive planet. In the latter case, the existing planet(s) may adjust their orbits in response to the perturbation induced -- 13 -- by the introduced planet. For these reasons, it is important to also consider the effect of massive "test planets" in order to obtain a more complete dynamical picture of the systems under consideration. In this section, we explore the effect of inserting massive bodies into a known planetary system. SWIFT's RMVS3 integrator cannot handle close encounters, when massive bodies are closer to each other than 3 Hill radii. For the massive-body simulations, we use the Mercury orbital integrator (Chambers 1999), which has a hybrid feature that switches from an MVS integration to a Bulirsch-Stoer method when objects are within 3 Hill radii of each other. General relativistic effects have not been included. For these tests, fictitious planets were placed in each system on initially circular orbits at 0.05 AU intervals from 0.05-2.00 AU. The masses of the bodies were set at a Saturn mass (=0.3 MJup); this is comparable to the mass detectable by the radial-velocity survey, and is the mass used by Raymond & Barnes (2005) in a similar investigation. These simulations ran for 106 yr, and we observed that unstable configurations usually resulted in system destruction within 105 yr. Figure 14 shows a histogram of the survival times for the unstable trials. 4.4. Massive Body Results The results of the massive-body simulations are shown in Figure 15. The filled circles indicate test planets which remained throughout the 106 yr integration. For most of the systems, the regions stable for test particles are also stable for Saturn-mass planets. For HD 3651 and HD 80606, some test planets which crossed orbits with the known planet sur- vived. The 2:1 resonance of HD 3651b (a ∼ 0.45 AU) retained the Saturn-mass planet for 106 yr, although its eccentricity varied chaotically, reaching e ∼0.22. The HD 80606 system gave the most unexpected result: Saturn-mass planets remained in the region a ≤0.15 AU, despite crossing orbits with HD 80606b. The test planets at 0.05, 0.10, and 0.15 AU reached maximum eccentricities of 0.13, 0.26, and 0.57, respectively. For the test planets at 0.05 and 0.10 AU, the oscillations in eccentricity were regular in period and constant in amplitude, whereas for a = 0.15 AU, the oscillations varied in period and increased in amplitude toward the end of the 106 yr simulation (Figure 16). For the two cases in which the test plan- ets exhibited irregular variations in eccentricity, the simulations were continued for 107 yr, anticipating the eventual destruction of the system. The test planet at 0.45 AU in the HD 3651 system caused the ejection of HD 3651b after 1.8×106 yr. Likewise, the test planet at 0.15 AU in the HD 80606 system was ejected after 5.7×106 yr. -- 14 -- 5. Detection Limits 5.1. Methods In Wittenmyer et al. (2006), we described a detection-limits algorithm implemented on the sample of constant stars from the long-term planet search at McDonald Observatory. This approach was based on that used by Endl et al. (2002) to derive detection limits from their survey with the ESO Coude Echelle Spectrometer. In brief, we add a Keplerian signal to the existing velocity data, then attempt to recover that signal using a Lomb-Scargle periodogram. The mass of the simulated planet is increased until 99% of the injected signals are recovered with FAP<0.1%. For the constant stars in Wittenmyer et al. (2006), the null hypothesis is that no planets are present, and so the detection-limit algorithm can be applied directly to the velocity data. In the case of the known planet hosts, this null hypothesis no longer applies, and it would not do to "pre-whiten" those data by removing the known planet's orbit as if its parameters were known perfectly. The presence of an additional planet will act to modify the fitted parameters of the known planet. If two or more planets are present, and only one has been fitted, then part of the signal from the additional planets can be absorbed into the orbital elements of the 1-planet fit. To approach this task with the maximum rigor, these effects must be accounted for. Hence, the detection-limit algorithm was modified in the following way: the test Keplerian signal was added to each of the original data sets, then these modified data sets were fitted for the known planet(s) using GaussFit. A residuals file was generated and then subjected to the periodogram search as described above. This process of fitting and removing the known planet occurred for every injected test signal. This method has the advantage of being essentially identical to the planet-search method described in § 3.1. 5.2. Results All data used in the fits for each planet host were subjected to the limits-determination routine, using 100 trial periods at even steps in the logarithm between 2 days and the total duration of observations. The results are plotted in Figures 17-27; planets with masses above the lines were recovered in 99% of trials (solid and dotted lines), or 50% of trials (dashed lines), and hence can be ruled out by the data at those confidence levels, respectively. To match the parameter space specifically targeted in this study, and to match that of the test- particle simulations, the detection-limits plots show the inner 2 AU only. For the eccentric trials (solid lines), the eccentricity of the injected test signals was chosen to be the mean eccentricity of the surviving test particles from the N-body simulations described in § 4.2. -- 15 -- This approach was chosen because the dynamical simulations demonstrated that objects placed in circular orbits do not stay that way; the eccentricity of an undetected low-mass planet is expected to be influenced to nonzero values by the known giant planet. It is important to note that the limits presented here represent the companions that can be ruled out by the data with 99% confidence. Lower-mass planets could have been detected in this survey, but not necessarily at all (or 99% of all) possible configurations. It is likely that a particular combination of parameters for a simulated planet makes that signal fiendishly difficult to recover by this method, owing to the known planet's radial-velocity signal and the sampling of the data. This is particularly important for simulated eccentric planets, where the velocity signal becomes markedly non-sinusoidal. The 50% limits are also shown to illustrate the effect of relaxing the recovery criteria in order to reduce the impact of especially unfortunate configurations. Table 6 summarizes the results of the detection limits computations. The mean detection limits shown in Table 6 show that we could have detected 99% of planets with M sin i ∼1.6 Neptune masses at 0.05 AU, and M sin i ∼2.4 Neptune masses at 0.1 AU. The tightest limits were obtained for HD 3651, HD 108874, and 47 UMa, in which we are able to rule out Neptune-mass planets within 0.1 AU at the 99% level. For all of the systems, the limits shown in Figures 17-27 exhibit some "blind spots" evident where the periodogram method failed to recover the injected signals with FAP<0.1%. Typically this occurs at certain trial periods for which the phase coverage of the observational data is poor, and often at the 1- month and 1-year windows. Using this method of fitting the known planet for each injected trial signal, such regions of ignorance are also present at periods close to that of the existing planet. 6. Discussion The aim of this project has been to intensively monitor known planetary systems in search of additional planets. However, in the sample of 22 planet hosts, the results have been quite the opposite. These new data cast doubt on the existence of two of the previously known planets, HD 20367b and 47 UMa c (Wittenmyer et al. 2007a). The announcement by Bean et al. (2008) of a third planet in the HD 74156 system, one of the targets of this study, prompted a detailed investigation; at present we cannot confirm this object. These results suggest that systems with multiple giant planets are considerably more rare, or harder to detect, than anticipated at the outset of this project. In this section, we will explore some reasons why no new multiple-planet systems were -- 16 -- detected. Four possibilities are: 1) Biases in the target selection conspire against detection of weak signals, 2) There exist fundamental physical differences between single- and multiple- planet systems, 3) We did not obtain a sufficient quantity of high-quality data, and 4) Apparent single-planet systems may contain terrestrial-mass planets below the detection threshold. 6.1. Biases in the Sample As with any scientific experiment, it is important to determine whether the sample selection resulted in unforeseen biases which affected the results. The target-selection process for this study, described in § 2, included an intentional bias in favor of planet hosts with "large" (10-20 m s−1) radial-velocity scatter about the orbital solution. The reasoning for this choice is straightforward: if a single planet can be fit with minimal scatter, there is little room for additional undetected planets to hide in the residuals. An unintended consequence of this selection criterion is that the excess scatter may be intrinsic to the star rather than indicative of additional planets. The achievable velocity precision improves with the number and strength of photospheric lines (Butler et al. 1996). Stars with higher temperatures or lower metallicities would have fewer and weaker lines, and result in lower velocity precision. In rapidly rotating stars, the spectral lines are broadened, which also degrades the radial- velocity precision. Fischer & Valenti (2005) showed that the probability of a given star hosting a planet is positively correlated with its metallicity. In addition, those authors suggested that among planet host stars, metal-rich stars are more likely to host multiple planets. To check for these sorts of biases, we can perform a Kolmogorov-Smirnov (K-S) test to determine the probability that two samples are drawn from the same distribution. Comparing our sample of 22 planet host stars with other planet hosts not targeted (N = 200), the K-S test shows no significant differences in Tef f (P = 0.698), [Fe/H] (P = 0.841), or V sin i (P = 0.323). A comparison of the mean and median values of these quantities is shown in Table 7. The uncertainties are too large to make statistically meaningful comparisons, but the K-S test results suggest that there are no significant differences between the 22 planet hosts targeted here and those planet hosts not chosen. 6.2. Fundamental Differences In this section, we ask the question, "Is there something special about the multi-planet systems"? Physical differences between single and multiple planet systems could arise either from the host star or from the processes of formation and dynamical evolution. Table 8 gives -- 17 -- statistics on the planetary and stellar parameters for single and multiple-planet systems. Only those planets detected by radial-velocity with M sin i <13 MJup were considered in the compilation of these statistics. Table 9 shows the results of K-S tests on the planetary and stellar characteristics listed in Table 8. None of the parameters tested showed statistically significant differences between single and multiple planet systems. There are hints from the data in Table 8, and the K-S test results in Table 9 that planets in multiple systems have larger a and smaller M sin i than those in single-planet systems. Both of these trends would work against the radial-velocity detection of planets in multiple systems. As the semimajor axis a increases by a factor of N, the velocity semiamplitude K decreases by √N , and as the planet mass decreases by a factor of N, K also drops by a factor of N. It is also possible that a tendency toward lower mass and larger semimajor axis in multi-planet systems is the result of a selection effect. Once a single planet is found, follow-up observations may reveal longer-period (larger a) planets, and intensive monitoring programs such as this work may then find lower-mass planets. We can test whether a selection effect is at work by computing the statistics in Table 8 for the first planet discovered in the known multi- planet systems. These results are also given in Table 9; by comparing only the first planet found in the multiple systems with single planets, any significant difference between the distributions vanishes. Recently, Wright et al. (2008) have presented a detailed investigation of multiple-planet systems, and they find that planets in multiple systems tend to have lower eccentricities than single planets. We discuss this possibility in § 6.5. Wright et al. (2008) also note that the orbital distances of planets in multiple systems are more evenly distributed in log-period, whereas single planets are more frequent at a ∼0.05 AU and near 1 AU. 6.3. Observing Strategy In considering whether there are important differences between the objects targeted in this work and known multi-planet systems, we can focus the comparison on the type of planetary system this survey was aimed at finding. The original motivation for this work was to investigate the possibility that systems containing a Jovian planet also contain Neptune- mass planets (1 Neptune mass=0.054 MJup). At this writing, there are four such systems: 55 Cnc, GJ 876, µ Ara (=HD 160691), and GJ 777A (=HD 190360). With a sample size of only four, a meaningful statistical comparison of the host stars is not possible, but one can look at the characteristics of the body of radial-velocity data for these systems. In so doing, we ask whether those data are of exceptional quality or quantity which facilitated the detection of the additional low-mass planets in those systems. The recent detection of a fifth planet in the 55 Cnc system by Fischer et al. (2008) used 636 measurements, binned into 250 Lick visits and 70 Keck visits. The detection of the fourth planet by McArthur et al. (2004) -- 18 -- used 138 HET observations combined with 143 Lick data points (Marcy et al. 2002) and 48 data points from Naef et al. (2004). For µ Ara, the Neptune-mass planet was discovered using the HARPS spectrograph, which consistently delivers velocity precision of ∼1 m s−1 (Santos et al. 2004b; Pepe et al. 2007). The fourth planet in the µ Ara system (Pepe et al. 2007) was discovered using a total of 86 HARPS measurements combined with data from CORALIE and the AAT. The 18M⊕ planet GJ 777Ac was discovered by Vogt et al. (2005) using 87 Keck velocities, and Rivera et al. (2005) found the 7.5M⊕ GJ 876d after 155 Keck observations. All four of these systems appear to have required an unusually large amount of the highest-quality data from Keck and HARPS, with a mean of 107 data points. By contrast, the targets in this work each received an average of 53 HET visits. It is possible that the number of visits required to detect a hot Neptune was underestimated. 6.4. Swarms of Earths Another possibility is that multiple-planet systems are indeed common, but, like our own Solar system, contain many terrestrial-mass objects which are undetectable by current radial- velocity surveys. Core-accretion simulations by Ida & Lin (2004a) predict a preponderance of 1-10M⊕ planets inside of 1 AU, and a "planet desert" in the range of 10-100 M⊕, arising due to rapid gas accretion by cores once they reach about 10 M⊕. The current survey is not sensitive to the terrestrial-mass objects, but planets within the "desert" could have been detected. Interestingly, Schlaufman et al. (2009) show that the presence of the planet desert could be confirmed by a radial-velocity survey with 1 m s−1 precision and ∼700 observations, which is similar in scope to the present work. Of course, many more than 22 systems need to be studied before conclusions can be made, but the characterization of hundreds of new systems by the Kepler spacecraft (Borucki et al. 2003) will help to define the upper and lower mass boundaries of the planet desert. Ida & Lin (2004a) note that the lower mass boundary would indicate the core mass required for rapid gas accretion, while the upper mass boundary would give insight into the mechanism by which gas accretion stops. Kepler discoveries of short-period super-Earths with masses 1-10M⊕ would lend further support to the core-accretion mechanism. Simulations of planetesimal formation and migration also provide support for the ex- istence of terrestrial-mass planets in systems with a gas giant planet. The GJ 876 sys- tem (Rivera et al. 2005), which contains two giant planets and an interior "super-Earth" (M sin i=7.5 M⊕), is thought to have originated by the shepherding of material as the giant planets migrated inward (Zhou et al. 2005). 200 Myr simulations by Raymond et al. (2006) and Mandell et al. (2007) resulted in the formation of planets with 1-5 Earth masses interior -- 19 -- and exterior to the migrating hot Jupiter. Those models included only Type II migration, in which the migrating giant planet opens a gap in the protoplanetary disk. The models of Fogg & Nelson (2007) consider the effects of Type I migration, in which the giant planet does not open a gap in the disk and inward drift is driven by differential torques on the planet. Inclusion of Type I migration did not alter the general outcome, that planets of several Earth masses are shepherded inward by the hot Jupiter, and some remain exterior to it. These models indicate that the inner regions of planetary systems may be populated with terrestrial-mass planets which would remain wholly undetectable by current radial-velocity surveys. Although this work achieved detection limits of 15-30 Earth masses, rocky planets in the range of 1-5 Earth masses could easily have been missed. 6.5. Broader Implications for Planetary Systems We now take a step back and look at the bigger picture of planetary system formation and evolution. Based on the target selection and the resulting detection limits, this survey was most sensitive to systems with two giant planets (larger than Saturn mass). More specifically, our "key demographic" is a system with a "cold" Jupiter (a ∼1 AU) and a close-in planet with M sin i > ∼ 1-2 Neptune masses (0.05-0.1 MJup). The detection limits given in § 5 exclude such configurations at the 99% level for all of the planetary systems considered here. Systems containing a long-period, massive planet could also have been detected by trends or curvature in the velocity residuals; no such trends were present for any of the targets. This survey was much less sensitive to planetary systems like our own, with multiple terrestrial-mass planets and long-period giants, for the reasons discussed in § 6.4. Planetary systems with architectures like our own Solar system may yet be common, but we will need to wait for the results from Kepler to begin making quantitative statements. The results of this work are most useful in assessing the frequency of planetary sys- tems in which extensive migration has occurred, to bring two gas giant planets interior to the "snow line." In the core-accretion theory of giant planet formation (Pollack et al. 1996; Lissauer 1995), surface-density enhancement by ices facilitates the formation of ∼10-15 M⊕ cores. The snow line, beyond which ices are present in the protoplanetary disk, has been estimated to lie at 1.6-1.8 AU in a minimum-mass solar nebula (Lecar et al. 2006). Per- haps the extensive migration required to construct systems with multiple giant planets with a < ∼ 2 AU is uncommon; the typical timescale in which a system is undergoing migration may be short. In other words, migration may be fast, a hypothesis which has led to theoretical scenarios in which the observed planets are the last of many "batches" of planets which mi- grated onto the host star (Trilling et al. 2002; Ida & Lin 2004a; Narayan et al. 2005). Type I -- 20 -- migration, in which a net viscous torque on the protoplanet changes its orbit (Ward 1997), results in very fast migration with a timescale proportional to M −1 planet. When a planet is massive enough (0.3-1.0 MJup: Armitage 2007) to clear a gap in the disk, the slower Type II migration begins. The results of this work, showing a deficit of systems with multiple giant planets inside of 2-3 AU, suggest that they are dominated by Type I migration and rapidly accrete onto the star. Tanaka et al. (2002) showed that the Type I migration timescale is inversely proportional to the disk mass: planets in more massive disks migrate faster. If we make the reasonable assumption that multiple giant planets form from unusually massive disks, then Type I migration works against these planets surviving the migration if they remain below the gap-opening mass. To generate systems with multiple giant planets inside of 2-3 AU, migration should then be rapid enough to bring them there, but not so fast as to send the planets into the star. The results presented here suggest that such a scenario is uncommon. In addition to migration, the dynamical history of planetary systems is an important factor in producing the observed architectures. The eccentricity distribution of extrasolar planets suggests that dynamically active histories are common. Interactions between giant planets can result in the ejection of one while imparting a significant eccentricity on the remaining planet (Rasio & Ford 1996; Ford et al. 2005; Malmberg & Davies 2008). Systems containing a single giant planet on a moderately eccentric orbit may be the result of such encounters, and thus less likely to host the sort of planets this survey was seeking. The median eccentricity of the planets targeted in this work is 0.29, compared to a median e of 0.15 for all other planets. Comparing the distributions by the K-S test gives a probability of 0.048, indicating a marginally significant difference between the two. Fischer et al. (2008) use the relatively low eccentricities (e < 0.2) of the five 55 Cnc planets to suggest that a benign dynamical history allowed so many planets to remain. The GJ 876, HD 37124, HD 73526, and GJ 581 systems also have multiple planets with e < 0.2, but counterexamples are found in HD 160691, HD 74156, and HD 202206 (emax=0.57, 0.64, and 0.44, respectively). An uneventful dynamical history contributes to a planetary system's observed end state, but comprises only a part of the picture in combination with its formation history. A primary goal of the search for extrasolar planets is to estimate how common the architecture of our own Solar system might be. If the processes of planet formation and migration form many systems similar to our own, it becomes more likely that Earth-like planets may be present. The results of this work indicate that planetary systems like our own may be common if 1) terrestrial-mass planets are present but undetected, or 2) Type I migration timescales are so short that multiple giant planets rarely end up within 2-3 AU. Conversely, our Solar system may be rare if the dynamical history of most planetary systems results in many ejections and high eccentricities. -- 21 -- 7. Summary We have carried out an intensive radial-velocity campaign to monitor 22 known plan- etary systems for additional planets. No new planets were found, and these new data do not support the proposed planets HD 20367b, HD 74156d, and 47 UMa c. We have used test particles and Saturn-mass bodies to probe 20 planetary systems for regions in which additional planets could exist. The massive-body results are consistent with the test-particle results: each of these systems has regions, sometimes quite large, where additional planets may remain in stable orbits. Finally, we show that this campaign could have detected 99% of planets with M sin i < ∼ 2.6 Neptune masses within 0.10 AU. This material is based on work supported by the National Aeronautics and Space Admin- istration under Grants NNG04G141G, NNG05G107G issued through the Terrestrial Planet Finder Foundation Science program and Grant NNX07AL70G issued through the Origins of Solar Systems Program. We are grateful to the HET TAC for their generous allocation of telescope time for this project. Much of the computing for the dynamical simulations used the Lonestar cluster at the Texas Advanced Computing Center. This research has made use of NASA's Astrophysics Data System (ADS), and the SIMBAD database, operated at CDS, Strasbourg, France. The Hobby-Eberly Telescope (HET) is a joint project of the Uni- versity of Texas at Austin, the Pennsylvania State University, Stanford University, Ludwig- Maximilians-Universitat Munchen, and Georg-August-Universitat Gottingen The HET is named in honor of its principal benefactors, William P. Hobby and Robert E. Eberly. REFERENCES Baranne, A., et al. 1996, A&AS, 119, 373 Barnes, R., Go´zdziewski, K., & Raymond, S. N. 2008, ApJ, 680, L57 Barnes, R., & Quinn, T. 2004, ApJ, 611, 494 Barnes, R., & Raymond, S. N. 2004, ApJ, 617, 569 Bean, J. L., McArthur, B. E., Benedict, G. F., & Armstrong, A. 2008, ApJ, 672, 1202 Bean, J. L., McArthur, B. E., Benedict, G. F., Harrison, T. E., Bizyaev, D., Nelan, E., & Smith, V. V. 2007, AJ, 134, 749 Bodenheimer, P., Hubickyj, O., & Lissauer, J. J. 2000, Icarus, 143, 2 -- 22 -- Bonfils, X., et al. 2007, A&A, 474, 293 Borucki, W. J., et al. 2003, Proc. SPIE, 4854, 129 Butler, R. P., Marcy, G. W., Vogt, S. S., Fischer, D. A., Henry, G. W., Laughlin, G., & Wright, J. T. 2003, ApJ, 582, 455 Boss, A. P. 2003, ApJ, 599, 577 Boss, A. P. 1998, ApJ, 503, 923 Boss, A. P. 1995, Science, 267, 360 Bouchy, F., et al. 2008, arXiv:0812.1608 Butler, R. P., et al. 2006, ApJ, 646, 505 Butler, R. P., Marcy, G. W., Williams, E., McCarthy, C., Dosanjh, P., & Vogt, S. S. 1996, PASP, 108, 500 Chambers, J. E. 1999, MNRAS, 304, 793 Chatterjee, S., Ford, E. B., Matsumura, S., & Rasio, F. A. 2008, ApJ, 686, 580 Cochran, W. D., et al. 2004, ApJ, 611, L133 Cochran, W. D., Hatzes, A. P., & Hancock, T. J. 1991, ApJ, 380, L35 Endl, M., Cochran, W. D., Wittenmyer, R. A., & Boss, A. P. 2008, ApJ, 673, 1165 Endl, M., Kurster, M., & Els, S. 2000, A&A, 362, 585 Endl, M., Kurster, M., Els, S. H. A. P., Cochran, W. D., Dennerl, K., Dobereiner, S. 2002, A&A, 392, 671 Fischer, D. A., & Valenti, J. 2005, ApJ, 622, 1102 Fischer, D. A., et al. 2008, ApJ, 675, 790 Fischer, D. A., Marcy, G. W., Butler, R. P., Laughlin, G., & Vogt, S. S. 2002, ApJ, 564, 1028 Fogg, M. J., & Nelson, R. P. 2007, A&A, 472, 1003 Ford, E. B., Lystad, V., & Rasio, F. A. 2005, Nature, 434, 873 -- 23 -- Go´zdziewski, K., & Konacki, M. 2006, ApJ, 647, 573 Henry, G. W. 1999, PASP, 111, 845 Holmberg, J., Nordstrom, B., & Andersen, J. 2007, A&A, 475, 519 Horne, J. H., & Baliunas, S. L. 1986, ApJ, 302, 757 Ida, S., & Lin, D. N. C. 2008, ApJ, 673, 487 Ida, S., & Lin, D. N. C. 2004, ApJ, 604, 388 Ida, S., & Lin, D. N. C. 2004, ApJ, 616, 567 Jefferys, W. H., Fitzpatrick, M. J., & McArthur, B. E. 1987, Celestial Mechanics, 41, 39 Jones, B. W., Sleep, P. N., & Chambers, J. E. 2001, A&A, 366, 254 Juri´c, M., & Tremaine, S. 2008, ApJ, 686, 603 Kurster, M., Schmitt, J. H. M. M., Cutispoto, G., & Dennerl, K. 1997, A&A, 320, 831 Latham, D. W., Stefanik, R. P., Mazeh, T., Mayor, M., & Burki, G. 1989, Nature, 339, 38 Laughlin, G., & Chambers, J. E. 2002, AJ, 124, 592 Lecar, M., Podolak, M., Sasselov, D., & Chiang, E. 2006, ApJ, 640, 1115 Levison, H. F., & Duncan, M. J. 1994, Icarus, 108, 18 Levison, H. F., Lissauer, J. J., & Duncan, M. J. 1998, AJ, 116, 1998 Lissauer, J. J. 1995, Icarus, 114, 217 Lomb, N. R. 1976, Ap&SS, 39, 447 Lovis, C., et al. 2006, Nature, 441, 305 Malmberg, D., & Davies, M. B. 2008, arXiv:0811.3420 Mandell, A. M., Raymond, S. N., & Sigurdsson, S. 2007, ApJ, 660, 823 Marcy, G. W., Butler, R. P., Fischer, D. A., Laughlin, G., Vogt, S. S., Henry, G. W., & Pourbaix, D. 2002, ApJ, 581, 1375 Marcy, G. W., & Butler, R. P. 1992, PASP, 104, 270 -- 24 -- Mayor, M., et al. 2009, A&A, 493, 639 Mayor, M., & Queloz, D. 1995, Nature, 378, 355 Mayor, M., Udry, S., Naef, D., Pepe, F., Queloz, D., Santos, N. C., & Burnet, M. 2004, A&A, 415, 391 Menou, K., & Tabachnik, S. 2003, ApJ, 583, 473 McArthur, B. E., et al. 2004, ApJ, 614, L81 Naef, D., Mayor, M., Beuzit, J. L., Perrier, C., Queloz, D., Sivan, J. P., & Udry, S. 2004, A&A, 414, 351 Naef, D., et al. 2001, A&A, 375, L27 Narayan, R., Cumming, A., & Lin, D. N. C. 2005, ApJ, 620, 1002 Noyes, R. W., Hartmann, L. W., Baliunas, S. L., Duncan, D. K., & Vaughan, A. H. 1984, ApJ, 279, 763 Paulson, D. B., Saar, S. H., Cochran, W. D., & Hatzes, A. P. 2002, AJ, 124, 572 Pepe, F., et al. 2007, A&A, 462, 769 Perrier, C., Sivan, J.-P., Naef, D., Beuzit, J. L., Mayor, M., Queloz, D., & Udry, S. 2003, A&A, 410, 1039 Pollack, J. B., Hubickyj, O., Bodenheimer, P., Lissauer, J. J., Podolak, M., & Greenzweig, Y. 1996, Icarus, 124, 62 Ramsey, L. W., et al. 1998, Proc. SPIE, 3352, 34 Rasio, F. A., & Ford, E. B. 1996, Science, 274, 954 Raymond, S. N., & Barnes, R. 2005, ApJ, 619, 549 Raymond, S. N., Mandell, A. M., & Sigurdsson, S. 2006, Science, 313, 1413 Rivera, E., & Haghighipour, N. 2007, MNRAS, 374, 599 Rivera, E. J., & Lissauer, J. J. 2000, ApJ, 530, 454 Rivera, E. J., & Lissauer, J. J. 2001, ApJ, 558, 392 Rivera, E. J., et al. 2005, ApJ, 634, 625 -- 25 -- Santos, N. C., Israelian, G., & Mayor, M. 2004, A&A, 415, 1153 Santos, N. C., et al. 2004, A&A, 426, L19 Santos, N. C., Mayor, M., Naef, D., Pepe, F., Queloz, D., Udry, S., & Burnet, M. 2001, A&A, 379, 999 Scargle, J. D. 1982, ApJ, 263, 835 Schlaufman, K. C., Lin, D. N. C., & Ida, S. 2009, ApJ, 691, 1322 Takeda, G., Ford, E. B., Sills, A., Rasio, F. A., Fischer, D. A., & Valenti, J. A. 2007, ApJS, 168, 297 Tanaka, H., Takeuchi, T., & Ward, W. R. 2002, ApJ, 565, 1257 Trilling, D. E., Benz, W., Guillot, T., Lunine, J. I., Hubbard, W. B., & Burrows, A. 1998, ApJ, 500, 428 Trilling, D. E., Lunine, J. I., & Benz, W. 2002, A&A, 394, 241 Tull, R. G. 1998, Proc. SPIE, 3355, 387 Udry, S., et al. 2007, A&A, 469, L43 Udry, S., Mayor, M., & Queloz, D. 2003, Scientific Frontiers in Research on Extrasolar Planets, 294, 17 Udry, S., et al. 2000, A&A, 356, 590 Valenti, J. A., & Fischer, D. A. 2005, ApJS, 159, 141 Vogt, S. S., Butler, R. P., Marcy, G. W., Fischer, D. A., Henry, G. W., Laughlin, G., Wright, J. T., & Johnson, J. A. 2005, ApJ, 632, 638 Vogt, S. S., Butler, R. P., Marcy, G. W., Fischer, D. A., Pourbaix, D., Apps, K., & Laughlin, G. 2002, ApJ, 568, 352 Ward, W. R. 1997, Icarus, 126, 261 Wittenmyer, R. A., Endl, M., & Cochran, W. D. 2007, ApJ, 654, 625 Wittenmyer, R. A., Endl, M., Cochran, W. D., & Levison, H. F. 2007, AJ, 134, 1276 Wittenmyer, R. A., Endl, M., Cochran, W. D., Hatzes, A. P., Walker, G. A. H., Yang, S. L. S., & Paulson, D. B. 2006, AJ, 132, 177 -- 26 -- Wright, J. T., Upadhyay, S., Marcy, G. W., Fischer, D. A., Ford, E. B., & Johnson, J. A. 2008, arXiv:0812.1582 Wright, J. T., et al. 2007, ApJ, 657, 533 Wright, J. T., Marcy, G. W., Butler, R. P., & Vogt, S. S. 2004, ApJS, 152, 261 Zhou, J.-L., Aarseth, S. J., Lin, D. N. C., & Nagasawa, M. 2005, ApJ, 631, L85 Zucker, S., et al. 2002, ApJ, 568, 363 This preprint was prepared with the AAS LATEX macros v5.2. -- 27 -- Table 1. Stellar Parameters Star Spec. Type Distance K0V F8 G5IV F8V G0 F8V G5 G4IV F8V G0V G0 G5 F7V G0V G0 G5 K0 K0 K0V F9V HD 3651 HD 8574 HD 10697 HD 19994 HD 20367 HD 23596 HD 28185 HD 38529 HD 40979 HD 72659 HD 74156 HD 80606 HD 89744 47 UMa HD 106252 HD 108874 HD 114783 HD 128311 HD 130322 HD 136118 HD 178911B G5 HD 190228 G5IV (pc) 11.1±0.1 44.2±1.6 32.6±0.9 22.4±0.4 27.1±0.8 52.0±2.3 39.6±1.7 42.4±1.7 33.3±0.9 51.4±2.7 64.6±4.6 58±20 39.0±1.1 14.1±0.1 37.4±1.3 68.5±5.8 20.4±0.4 16.6±0.3 29.8±1.3 52.3±2.3 47±11 62.1±3.1 Mass (M⊙) [Fe/H] Tef f (K) V sin i (km s−1) logR′ HK 0.882±0.026 1.122±0.022 1.112±0.026 1.365±0.042 1.04±0.06a 1.159±0.062 0.98±0.05c 1.477±0.052 1.154±0.028 1.068±0.022 1.238±0.044 0.958±0.072 1.558±0.048 1.063±0.029 1.007±0.024 0.950±0.036 0.853±0.034 0.828±0.012 0.836±0.018 1.191±0.026 1.014±0.057 1.821±0.050 5221±44 0.24±0.03 6050±44 -0.03±0.03 5680±44 0.17±0.03 0.27±0.03 6188±44 -0.09±0.10b 5998±75 5904±44 0.33±0.03 0.12±0.10b 5546±75 5697±44 0.51±0.03 0.15±0.03 6089±44 5920±44 -0.02±0.03 6068±44 0.11±0.03 5573±44 0.47±0.03 6291±44 0.26±0.03 0.04±0.03 5882±44 5870±44 -0.07±0.03 5551±44 0.19±0.03 5135±44 0.21±0.03 4965±44 0.08±0.03 -0.02±0.03 5308±44 6097±44 -0.11±0.03 5668±44 0.34±0.03 -0.24±0.03 5348±44 1.1 4.5 2.5 8.6 3.0 4.2 3.0 3.9 7.4 2.2 4.3 1.8 9.5 2.8 1.9 2.2 0.9 3.6 1.6 7.3 1.9 1.9 -4.99±0.05 -4.88±0.04 -5.07±0.15 -4.93±0.04 -4.50±0.05 -4.96±0.05 -5.37±0.40 -5.01±0.03 -4.59±0.01 -5.02±0.09 · · · · · · -5.03±0.04 -5.03±0.07 -4.91±0.14 · · · · · · · · · -4.76±0.02 -4.91±0.04 -4.83±0.02 -4.98±0.02 aMass obtained from Holmberg et al. (2007). b[Fe/H], Tef f , and V sin i obtained from Holmberg et al. (2007). cMass obtained from Santos et al. (2004a). -- 28 -- Table 2. Summary of Published Radial-Velocity Data Star Reference N < σ > (m s−1) RMS about fit (m s−1) HD 3651 HD 8574 HD 8574 HD 10697 HD 19994 HD 23596 HD 28185 HD 38529 HD 40979 HD 72659 HD 74156 HD 80606 HD 89744 47 UMa 47 UMa HD 106252 HD 106252 HD 108874 HD 114783 HD 128311 HD 130322 HD 130322 HD 136118 HD 178911B HD 178911B HD 190228 Butler et al. (2006) Perrier et al. (2003) Butler et al. (2006) Butler et al. (2006) Mayor et al. (2004) Perrier et al. (2003) Santos et al. (2001) Butler et al. (2006) Butler et al. (2006) Butler et al. (2006) Naef et al. (2004) Naef et al. (2001b) Butler et al. (2006) Fischer et al. (2002a) Naef et al. (2004) Perrier et al. (2003) Butler et al. (2006) Vogt et al. (2005) Butler et al. (2006) Vogt et al. (2005) Udry et al. (2000) Butler et al. (2006) Butler et al. (2006) Zucker et al. (2002) Butler et al. (2006) Perrier et al. (2003) 163 41 26 59 48 39 40 162 65 32 95 61 50 91 44 40 15 49 54 76 118 12 37 51 14 51 3.4 10.3 10.4 2.7 6.7 9.1 6.5 5.3 9.1 3.2 10.8 13.7 11.2 5.7 7.3 10.7 11.4 3.4 2.7 3.3 12.4 2.7 16.1 10.4 2.7 8.7 6.6 13.1 23.0 6.8 8.1 9.2 10.0 13 23 4.2 10.6 17.7 16.0 7.4 7.4 10.5 9.1 3.7 4.7 18.0 16.1 11.0 22.0 11.0 7.7 8.0 Table 3. Keplerian Orbital Solutions Planet Period (days) T0 (JD-2400000) e ω K (degrees) (m s−1) M sin i (MJup) a (AU) χ2 ν rms m s−1 HD 3651 b HD 8574 b HD 10697 b HD 19994 b HD 23596 b HD 28185 b HD 38529 b HD 38529 c HD 40979 b HD 72659 b HD 74156 ba HD 74156 c HD 80606 b HD 89744 b 47 UMa bb HD 106252 b HD 108874 b HD 108874 c HD 114783 b HD 128311 b HD 128311 c HD 130322 bc HD 136118 b HD 178911B b HD 190228 b 62.218±0.015 227.0±0.2 1075.2±1.5 466.2±1.7 1561±12 385.9±0.6 53932.6±0.6 53981.0±3.2 51480±18 53757±72 53163±22 53793.6±8.8 14.3098±0.0005 54012.64±0.16 2140.2±5.7 264.15±0.23 3383±100 52256.4±6.4 53919.0±2.7 51572±52 51.645±0.003 53788.59±0.09 2473±13 53415±13 0.596±0.036 0.297±0.026 0.099±0.007 0.063±0.062 0.266±0.014 0.092±0.019 0.257±0.015 0.341±0.005 0.252±0.014 0.271±0.022 0.627±0.009 0.432±0.013 111.429±0.001 53421.923±0.004 0.9324±0.0006 256.78±0.05 1076.6±2.3 1531.0±4.7 395.8±0.6 1624±23 493.7±1.8 454.2±1.6 923.8±5.3 10.7085±0.0003 1187.3±2.4 71.484±0.002 1136.1±9.9 51505.5±0.4 49222±347 53397.5±4.7 54069±17 52839±44 53806±14 53835±11 56987±41 53995.0±2.3 52999.5±5.3 53808.1±0.3 53522±12 0.673±0.007 0.012±0.023 0.482±0.011 0.082±0.021 0.239±0.031 0.144±0.032 0.345±0.049 0.230±0.058 0.011±0.020 0.338±0.015 0.114±0.003 0.531±0.028 242.5±4.5 26.6±5.4 111.2±6.3 346±55 272.6±3.3 351.9±8.2 92.5±3.9 17.8±1.2 323.4±4.1 241±8 176.5±1.2 258.6±2.7 300.4±0.3 195.1±1.0 147±117 292.8±1.8 232±10 27±10 86±11 63±16 28±15 145±77 319.9±2.1 168.2±1.6 101.2±2.1 15.9±0.7 58.3±1.8 115.4±1.1 29.3±2.1 127.0±2.0 158.8±4.2 56.1±0.9 173.2±1.2 119.4±2.2 42.4±1.1 109.6±2.3 116.5±3.3 470.6±1.8 271.6±4.0 46.6±1.1 138.8±2.0 37.0±0.8 18.2±0.7 31.9±0.9 46.5±4.5 78.8±2.6 108.3±2.0 210.7±2.5 343.3±1.0 91.4±3.0 0.229±0.008 1.80±0.06 6.21±0.15 1.37±0.12 7.71±0.39 5.59±0.33 0.839±0.030 13.38±0.39 4.01±0.13 3.15±0.14 1.80±0.06 8.06±0.37 3.91±0.19 8.44±0.23 2.45±0.10 6.92±0.16 1.29±0.06 0.99±0.06 1.10±0.06 1.45±0.13 3.24±0.10 1.04±0.03 11.60±0.25 7.03±0.28 5.93±0.20 0.295±0.003 0.757±0.005 2.131±0.018 1.305±0.016 2.772±0.062 1.032±0.019 0.131±0.002 3.712±0.048 0.846±0.007 4.511±0.114 0.292±0.004 3.850±0.054 0.447±0.011 0.918±0.010 2.100±0.022 2.611±0.026 1.038±0.014 2.659±0.060 1.160±0.019 1.086±0.008 1.745±0.017 0.0896±0.0006 2.333±0.020 0.339±0.006 2.604±0.032 3.82 2.21 3.39 5.27 0.88 2.28 6.32 6.32 4.44 1.00 1.60 1.60 1.41 2.58 3.61 1.42 0.88 0.88 4.91 21.38 21.38 4.29 1.82 1.80 0.78 6.3 14.2 8.1 14.0 8.7 9.5 11.8 11.8 20.3 6.6 11.5 11.5 13.3 15.2 10.2 12.2 4.1 4.1 6.3 16.9 16.9 8.9 16.5 9.1 7.4 -- 2 9 -- aResults from two-planet fit. bResults from one-planet fit. cResults for HD 130322 exclude data from Udry et al. (2000). -- 30 -- Table 4. Summary of Radial-Velocity Data Star N RMS about fit (m s−1) ∆T (days) Source HD 3651 HD 3651 HD 3651 HD 3651 (total) HD 8574 HD 8574 HD 8574 HD 8574 HD 8574 (total) HD 10697 HD 10697 HD 10697 HD 10697 (total) HD 19994 HD 19994 HD 19994 HD 19994 (total) HD 20367c HD 20367 HD 20367 (total) HD 23596 HD 23596 HD 23596 HD 23596 (total) HD 28185 HD 28185 HD 28185 (total) HD 38529 HD 38529 HD 38529 HD 38529 (total) HD 40979 HD 40979 HD 40979 HD 40979 (total) HD 72659 HD 72659 HD 72659 (total) HD 74156 HD 74156 HD 74156 (total) HD 80606 HD 80606 HD 80606 HD 80606 (total) HD 89744 163 35 4 202 41 44 16 26 128 59 32 40 131 48 56 12 116 81 19 100 39 63 6 108 40 34 74 162 73 7 242 65 91 4 160 32 53 85 95 82 177 61 23 46 130 50 6.5 5.1 9.3 6.3 14.8 8.7 13.4 20.7 14.2 6.5 8.8 9.7 8.1 14.8 12.5 18.5 14.0 12.9 10.5 12.4 9.4 8.5 5.8 8.7 10.4 8.5 9.5 13.0 8.9 9.2 11.8 22.8 18.9 9.6 20.3 4.1 7.8 6.6 13.8 8.3 11.5 18.6 6.1 5.3 13.3 16.2 Butler et al. (2006) HETa 2.7mb Perrier et al. (2003) HET 2.7m Butler et al. (2006) Butler et al. (2006) 2.7m HET Mayor et al. (2004) HET 2.7m HET 2.7m Perrier et al. (2003) HET 2.7m Santos et al. (2001) HET Butler et al. (2006) HET 2.7m Butler et al. (2006) HET 2.7m Butler et al. (2006) HET Naef et al. (2004) HET Naef et al. (2001b) HET Butler et al. (2006) Butler et al. (2006) 7376 3609 4057 3367 974 3603 2971 3745 3588 3593 3408 2893 -- 31 -- Table 4 -- Continued Star N RMS about fit (m s−1) ∆T (days) Source HD 89744 HD 89744 HD 89744 (total) 47 UMa 47 UMa 47 UMa 47 UMa 47 UMa (total) HD 106252 HD 106252 HD 106252 HD 106252 HD 106252 (total) HD 108874 HD 108874 HD 108874 (total) HD 114783 HD 114783 HD 114783 (total) HD 128311 HD 128311 HD 128311 (total) HD 130322 HD 130322 HD 130322 HD 130322 (total) HD 136118 HD 136118 HD 136118 HD 136118 (total) HD 178911B HD 178911B HD 178911B HD 178911B (total) HD 190228 HD 190228 HD 190228 HD 190228 (total) 33 9 92 91 44 43 77 255 40 43 15 12 110 49 40 89 54 34 88 76 78 154 12 30 5 47 37 64 4 108 51 40 14 105 51 42 8 101 12.9 19.0 15.2 11.1 11.8 11.4 7.0 10.2 14.8 8.2 12.2 16.1 12.2 3.4 4.8 4.1 6.6 5.8 6.3 15.8 17.9 16.9 8.3 8.7 13.3 8.9 21.6 18.3 14.9 16.5 11.5 5.4 7.5 9.1 8.8 9.8 9.3 9.2 HET 2.7m Fischer et al. (2002a) Naef et al. (2004) 2.7m HET Perrier et al. (2003) HET Butler et al. (2006) 2.7m Vogt et al. (2005) HET Butler et al. (2006) HET Vogt et al. (2005) HET Butler et al. (2006) HET 2.7m Butler et al. (2006) HET 2.7m Zucker et al. (2002) HET Butler et al. (2006) Perrier et al. (2003) HET 2.7m 2943 7673 3682 2850 3208 3335 2496 3450 3392 3776 a9.2 m Hobby-Eberly Telescope. bMcDonald Observatory 2.7 m Harlan J. Smith Telescope. cNo planet was fit. -- 32 -- Table 5. Results of Periodogram Analysis Star Period (days) FAP HD 3651 HD 8574 HD 10697 HD 19994 HD 20367 HD 23596 HD 28185 HD 38529 HD 40979 HD 72659 HD 74156 HD 80606 HD 89744 47 UMaa 47 UMab HD 106252 HD 108874 HD 114783 HD 128311 HD 130322 HD 136118 HD 178911B HD 190228 44.17 2272.73 26.68 54.88 5.58 25.13 4.76 294.12 2.26 6.99 80.39 357.14 23.27 2380.95 2.91 322.58 12.39 8.44 0.707 0.687 0.028 0.399 0.085 0.141 0.224 0.023 0.795 0.758 0.035 0.616 0.075 0.045 0.341 0.126 0.857 0.925 11.21 <0.0001 0.002 0.014 0.925 0.777 438.60 442.48 7.88 2.57 aResiduals from one-planet fit. bResiduals from two-planet fit. -- 33 -- Table 6. Companion Limit Summary Star Eccentricity M sin i 0.05 AU (MJup) M sin i 0.1 AU (MJup) Median K Median K 99% recovery 50% recovery m s−1 m s−1 HD 3651 HD 3651 HD 8574 HD 8574 HD 10697 HD 10697 HD 19994 HD 19994 HD 20367 HD 23596 HD 23596 HD 28185 HD 28185 HD 38529 HD 38529 HD 40979 HD 40979 HD 72659 HD 72659 HD 74156 HD 74156 HD 80606 HD 80606 HD 89744 HD 89744 47 UMa 47 UMa HD 106252 HD 106252 HD 108874 HD 108874 HD 114783 HD 114783 HD 128311 HD 130322 HD 130322 HD 136118 HD 136118 HD 178911B HD 178911B HD 190228 HD 190228 Mean (99% recovery) Mean (50% recovery) 0.20 0 0.10 0 0.04 0 0.09 0 0 0.10 0 0.09 0 0.12 0 0.11 0 0.10 0 0.15 0 0.31 0 0.01 0 0.02 0 0.15 0 0.15 0 0.11 0 0 0.02 0 0.11 0 0.07 0 0.16 0 0 0 0.025 0.024 0.124 0.124 0.059 0.059 0.116 0.117 0.098 0.081 0.078 0.083 0.080 0.078 0.075 0.135 0.123 0.057 0.054 0.080 0.074 0.119 0.104 0.176 0.168 0.039 0.039 0.091 0.087 0.035 0.034 0.056 0.056 0.102 0.147 0.147 0.125 0.120 0.061 0.066 0.080 0.077 0.087±0.036 0.063±0.027 0.041 0.040 0.142 0.143 0.094 0.094 0.173 0.166 0.122 0.091 0.092 0.129 0.129 0.123 0.124 0.201 0.193 0.085 0.085 0.109 0.105 0.184 0.160 0.197 0.197 0.067 0.067 0.179 0.173 0.059 0.055 0.083 0.080 0.166 0.231 0.231 0.224 0.226 0.130 0.124 0.114 0.110 0.131±0.052 0.090±0.036 4.8 4.4 14.1 14.1 7.4 7.4 16.2 16.2 12.3 8.5 8.5 12.3 11.7 8.9 8.9 17.8 17.0 8.1 8.1 10.7 10.2 18.7 15.5 18.7 18.7 6.1 6.1 12.9 12.3 5.8 5.6 8.5 8.1 16.2 22.5 22.5 16.2 16.2 10.2 9.7 8.1 7.7 · · · 3.2 · · · 10.2 · · · 5.6 · · · 10.7 9.3 · · · 6.4 · · · 9.7 · · · 5.8 · · · 12.3 · · · 5.6 · · · 7.4 · · · 10.2 · · · 12.9 · · · 4.6 · · · 9.3 · · · 3.5 · · · 6.1 12.9 · · · 13.5 · · · 12.9 · · · 7.0 · · · 5.8 -- 34 -- Table 6 -- Continued Star Eccentricity M sin i M sin i Median K Median K 0.05 AU 0.1 AU 99% recovery (MJup) (MJup) m s−1 50% recovery m s−1 Table 7. Comparison of Stellar Characteristics Quantity Targets Non-Targets Units [F e/H] (mean) [F e/H] (median) Tef f (mean) Tef f (median) (B − V ) (mean) (B − V ) (median) V sin i (mean) V sin i (median) 0.12±0.18 0.13 5741±361 5697 0.67±0.11 0.63 3.72±2.50 2.48 0.07±0.23 0.14 5608±496 5704 0.74±0.20 0.69 2.75±1.72 2.40 dex dex K K mag mag km s−1 km s−1 Table 8. Characteristics of Single and Multiple Planet Systems Quantity Single Multiple Units a (mean) a (median) e (mean) e (median) M sin i (mean) M sin i (median) Star mass (mean) Star mass (median) [F e/H] (mean) [F e/H] (median) Tef f (mean) Tef f (median) (B − V ) (mean) (B − V ) (median) 0.95±1.05 0.49 0.24±0.23 0.18 2.72±3.16 1.60 1.14±0.41 1.07 0.09±0.21 0.14 5640±473 5724 0.73±0.18 0.68 AU AU 1.19±1.38 0.63 0.19±0.17 0.16 1.93±2.38 MJup 1.03 MJup M⊙ M⊙ dex dex K K mag mag 1.06±0.32 1.04 0.05±0.30 0.14 5532±529 5584 0.77±0.22 0.72 -- 35 -- Table 9. K-S Tests on Single and Multiple Planet Systems Quantity K-S Probabilitya a a (first planet) M sin i M sin i (first planet) e Star mass [F e/H] Tef f (B − V ) 0.004 0.249 0.015 0.349 0.125 0.644 0.841 0.135 0.383 aProbability that the two samples are drawn from the same distribution. -- 36 -- Fig. 1. -- Left panel: Radial-velocity data for HD 20367. Filled circles: HET, open circles: 2.7m. The Geneva group's orbital solution for the proposed planet is shown as a solid line. Right panel: Lomb-Scargle periodogram of the velocities. The 5.5-day stellar rotation period is evident, but no other periodocities are significant. -- 37 -- Fig. 2. -- Photometric observations of HD 20367 phased to the stellar rotation period of 5.50 days. Two cycles are shown for clarity. -- 38 -- 346 days Fig. 3. -- Left panel: Periodogram of the residuals of a 2-planet fit for the HD 74156 system. The window function is shown as a grey dotted line, and the 346-day period of planet d is marked. Right panel: The residuals to the 2-planet fit, phased to a period of 346.6 days (Bean et al. 2008). For clarity, two cycles are shown, and the error bars have been omitted. A reference error bar representing the mean uncertainty of 9.65 m s−1 is shown. -- 39 -- Fig. 4. -- Left panel: Survival time as a function of initial semimajor axis for test particles in the HD 3651 system after 107 yr. The filled regions indicate test particles which survived. The orbital excursion of HD 3561b is indicated by the horizontal error bars at the top. Particles were placed on initially circular orbits with 0.05 < a < 2.00 AU. Right panel: Same, but for the HD 8574 system. Fig. 5. -- Same as Fig. 4, but for the HD 10697 (left) and HD 19994 (right) systems. -- 40 -- Fig. 6. -- Same as Fig. 4, but for the HD 23596 (left) and HD 28185 (right) systems. Fig. 7. -- Same as Fig. 4, but for the HD 38529 (left) and HD 40979 (right) systems. -- 41 -- Fig. 8. -- Same as Fig. 4, but for the HD 72659 (left) and HD 74156 (right) systems. HD 72659b, with an orbital excursion of 3.48-6.48 AU, is off the plot. The recently-announced planet HD 74156d, between planets b and c, was not included in the simulation, but would reside in the narrow stable strip. Fig. 9. -- Same as Fig. 4, but for the HD 80606 (left) and HD 89744 (right) systems. -- 42 -- Fig. 10. -- Same as Fig. 4, but for the 47 UMa (left) and HD 106252 (right) systems. Only 47 UMa b was considered in the simulations. An outer body would be too distant to affect the region under consideration. Fig. 11. -- Same as Fig. 4, but for the HD 108874 (left) and HD 114783 (right) systems. -- 43 -- Fig. 12. -- Same as Fig. 4, but for the HD 130322 (left) and HD 136118 (right) systems. Fig. 13. -- Same as Fig. 4, but for the HD 178911B (left) and HD 190228 (right) systems. -- 44 -- Fig. 14. -- Histogram of the survival times for the unstable test configurations (N = 352). Twenty realizations survived longer than 105 yr. -- 45 -- Fig. 15. -- Survival of Saturn-mass planets for 106 yr on initially circular orbits in 20 plane- tary systems. The orbital excursions of the existing planets are indicated by the horizontal error bars. Open circles represent unstable locations, filled circles were stable for 106 yr. Fig. 16. -- Left panel: Behaviour of the semimajor axis (top) and eccentricity (bottom) of a Saturn-mass test planet starting at a = 0.10 AU in the HD 80606 system over a 106 yr period. Right panel: Same, but for an object starting at a = 0.15 AU, which was then ejected at t = 5.7 × 106 yr. -- 46 -- Fig. 17. -- Left panel: Detection limits for additional planets in orbits with e = 0.20 in the HD 3651 system (solid line). This value represents the mean eccentricity of surviving test particles from the dynamical simulations discussed in § 4. Planets in the parameter space above the solid line are excluded at the 99% confidence level. Limits for planets in circular orbits are shown as dotted (99% recovery) and dashed (50% recovery) lines. Right panel: Same, but for HD 8574 (solid line: e = 0.10). Fig. 18. -- Left panel: Same as Fig. 17, but for HD 10697 (solid line: e = 0.04). Right panel: HD 19994 (solid line: e = 0.09). -- 47 -- Fig. 19. -- Left panel: Same as Fig. 17, but for HD 20367. These results were obtained without attempting to fit an existing planet, as no planet was confirmed in this system. Right panel: HD 23596 (solid line: e = 0.10). Fig. 20. -- Left panel: Same as Fig. 17, but for HD 28185 (solid line: e = 0.09). Right panel: HD 38529 (solid line: e = 0.12). -- 48 -- Fig. 21. -- Left panel: Same as Fig. 17, but for HD 40979 (solid line: e = 0.11). Right panel: HD 72659 (solid line: e = 0.10). Fig. 22. -- Left panel: Same as Fig. 17, but for HD 74156 (solid line: e = 0.15). Right panel: HD 80606 (solid line: e = 0.31). -- 49 -- Fig. 23. -- Left panel: Same as Fig. 17, but for HD 89744 (solid line: e = 0.01). Right panel: 47 UMa (solid line: e = 0.02). Only 47 UMa b was included in the limits computations. Fig. 24. -- Left panel: Same as Fig. 17, but for HD 106252 (solid line: e = 0.15). Right panel: HD 108874 (solid line: e = 0.15). -- 50 -- Fig. 25. -- Left panel: Same as Fig. 17, but for HD 114783 (solid line: e = 0.11). Right panel: HD 128311. Only circular orbits are considered since no test-particle simulations were conducted for this system. Fig. 26. -- Left panel: Same as Fig. 17, but for HD 130322 (solid line: e = 0.02). These results do not include data from Udry et al. (2000). Right panel: HD 136118 (solid line: e = 0.11). -- 51 -- Fig. 27. -- Left panel: Same as Fig. 17, but for HD 178911B (solid line: e = 0.07). Right panel: HD 190228 (solid line: e = 0.16). -- 52 -- Table 10. HET Radial Velocities for HD 3651 JD-2400000 Velocity (m s−1) Uncertainty (m s−1) 53581.21162 53600.79860 53604.79357 53606.78360 53608.77426 53615.96471 53628.74240 53669.61203 53678.79142 53682.78611 53687.77875 53691.76158 53694.75466 53696.76029 53955.83593 53956.83044 53957.82392 53973.80980 53976.78586 53978.97197 53985.95982 53987.95527 53989.74009 54003.70817 54005.68492 54056.78111 54062.55312 54064.54902 54130.55508 54282.92879 54352.96182 54394.64607 54399.61380 54414.77832 54423.75714 14.4 0.4 -6.7 -9.3 -10.9 -18.5 3.8 -12.2 -10.0 -18.1 17.0 12.3 16.7 15.5 12.3 7.3 10.6 -4.3 -13.2 -2.2 -15.2 -13.6 -20.6 13.2 17.2 -9.9 19.0 13.0 15.4 -1.7 -9.8 -1.9 -9.2 -1.0 -14.9 3.7 4.7 3.5 3.9 3.6 3.6 3.3 3.6 3.6 3.3 3.7 3.9 3.7 3.4 3.8 3.9 3.5 4.7 3.4 5.9 4.4 3.0 2.8 4.4 3.6 3.5 3.4 3.3 3.4 4.4 3.1 3.8 3.5 3.6 4.4 Table 11. 2.7m Radial Velocities for HD 3651 JD-2400000 Velocity (m s−1) Uncertainty (m s−1) 53633.86853 53654.79777 53690.69920 54020.84477 -3.3 6.6 -3.4 0.1 5.0 6.6 6.5 5.8 -- 53 -- Table 12. HET Radial Velocities for HD 8574 JD-2400000 Velocity (m s−1) Uncertainty (m s−1) 53601.81736 53604.80314 53605.82189 53607.81271 53609.79513 53612.79858 53633.96072 53653.69022 53663.88291 53665.63807 53668.64072 53687.81363 53695.79449 53696.78211 53703.77271 53705.75396 53936.90653 53936.90653 53969.80550 53975.81126 53987.99307 53989.98424 53997.96536 54000.73798 54013.69475 54018.90830 54043.84966 54049.61516 54057.78673 54067.55166 54071.76076 54106.65631 54110.66352 54121.63331 54306.89112 54327.84854 54344.80292 54352.76329 54367.73942 54402.85369 54402.86084 54404.84838 54419.81583 54434.54809 -40.9 -51.4 -42.2 -45.9 -48.9 -46.1 -30.7 -22.6 -27.5 -16.9 -25.3 -6.7 -2.5 6.2 34.9 37.0 38.7 42.1 61.9 64.7 55.7 66.3 29.8 18.7 4.2 -2.5 -31.4 -43.5 -38.9 -37.1 -50.1 -14.8 -12.7 -3.8 -30.4 -31.8 -26.1 -24.5 0.4 49.0 42.4 52.9 65.1 85.3 6.3 6.8 6.9 6.7 7.3 7.3 7.8 9.2 9.0 8.9 8.7 9.5 11.0 10.6 9.6 9.8 7.5 11.1 7.4 10.6 9.0 9.6 8.1 10.0 8.7 7.8 10.2 9.9 9.9 8.9 9.7 9.7 10.5 9.3 8.2 7.3 7.9 9.3 8.2 8.9 8.1 8.8 10.8 9.3 -- 54 -- -- 55 -- Table 13. 2.7m Radial Velocities for HD 8574 JD-2400000 Velocity (m s−1) Uncertainty (m s−1) 52116.95398 52141.96262 52219.89758 52249.70181 52331.61330 52493.90858 52540.91557 52658.62787 52932.83883 53015.71309 53564.94976 53632.92472 53635.90969 53691.75684 53970.92894 54018.87142 24.0 70.0 -10.8 -30.2 1.7 -17.4 -13.9 -18.5 -45.8 29.6 2.1 -62.8 -22.1 -4.6 95.0 3.7 9.6 8.0 10.0 9.6 7.7 9.7 8.9 9.3 8.9 9.2 10.8 8.6 8.4 10.4 13.5 9.9 -- 56 -- Table 14. HET Radial Velocities for HD 10697 JD-2400000 Velocity (m s−1) Uncertainty (m s−1) 53581.90281 53606.84709 53653.91334 53663.69183 53665.67577 53681.83488 53681.83752 53690.81667 53694.60078 53696.79921 53701.77012 53703.79160 53923.95145 53954.87796 53956.86965 53958.88031 53969.83080 53971.83685 53984.79870 53988.80905 53990.98655 53999.74552 54041.64127 54042.65929 54049.62382 54056.63399 54069.56942 54071.57888 54105.67044 54108.67092 54130.60755 54135.60424 54330.86457 54344.80964 54346.79984 54352.78595 54357.77232 54366.75077 54419.83538 54424.59411 22.3 -21.0 -49.4 -43.8 -51.5 -67.6 -67.4 -60.4 -77.7 -73.3 -82.7 -70.7 -65.8 -47.2 -47.3 -46.6 -69.9 -61.7 -34.4 -26.3 -37.3 -28.7 -10.3 -10.5 -19.4 -3.7 14.7 6.6 33.5 23.4 41.9 60.0 130.5 108.3 124.2 136.5 114.1 119.1 117.8 121.5 9.3 8.7 7.9 8.9 8.4 8.4 8.6 10.3 10.0 9.2 10.4 10.0 8.3 8.5 8.8 9.0 9.4 8.8 8.3 8.5 9.1 7.8 8.3 8.4 8.7 10.6 9.7 9.8 8.9 9.2 10.4 10.3 8.7 8.8 8.4 9.3 8.2 9.1 11.5 9.1 -- 57 -- Table 15. 2.7m Radial Velocities for HD 10697 JD-2400000 Velocity (m s−1) Uncertainty (m s−1) 51066.97570 51152.79209 51211.61496 51239.60083 51449.91000 51503.72131 51529.67754 51558.57566 51775.92530 51811.88858 51859.67414 51917.68431 51946.64764 51987.56487 52116.96710 52247.79070 52306.67720 52493.92186 52539.87531 52577.87821 52897.88453 52932.87904 53017.69801 53215.85694 53215.87160 53320.75657 53564.96191 53566.91539 53635.89769 53690.71397 53968.92927 54018.85867 76.4 108.3 98.3 105.0 -20.3 -69.6 -72.8 -97.3 -83.4 -86.5 -64.9 -19.2 3.0 10.3 84.9 109.6 101.0 3.3 -31.9 -63.3 -68.5 -69.5 4.1 99.7 89.8 129.6 3.4 16.1 -51.9 -94.3 -88.3 -61.1 6.0 6.3 6.5 12.8 5.3 10.7 7.6 6.7 5.6 6.8 5.8 7.0 7.5 8.8 6.7 5.9 5.7 6.2 5.7 6.2 8.2 6.8 6.5 15.3 6.7 10.1 8.1 6.8 8.8 7.5 10.9 6.9 -- 58 -- Table 16. HET Radial Velocities for HD 19994 JD-2400000 Velocity (m s−1) Uncertainty (m s−1) 53605.28626 53608.94050 53612.92809 53627.89670 53633.88721 53655.82407 53663.80786 53665.80005 53669.87714 53675.84787 53680.83658 53685.73881 53689.81169 53691.72151 53694.72971 53696.79236 53701.77903 53703.78435 53743.59301 53749.65649 53771.60345 53964.96481 53966.95936 53985.91398 53987.90646 53989.91975 53996.96612 53998.88994 54000.87587 54003.87611 54008.85651 54018.91598 54047.74879 54050.73465 54055.73818 54061.72068 54065.69886 54067.70983 54069.67713 54071.73275 54084.64261 54105.58129 54122.62277 54130.61594 54330.98292 54352.90449 -22.1 -39.5 -26.6 -21.8 -3.1 11.6 5.6 10.1 9.5 5.1 20.6 18.7 29.1 34.3 20.8 24.2 36.5 36.2 46.2 58.6 42.0 7.0 -8.4 -30.2 -35.7 -31.4 -18.2 -11.4 -28.4 -16.3 -16.0 -25.5 -10.0 -10.5 -25.0 -27.8 -10.8 -9.9 -11.2 -13.2 -0.1 23.5 15.8 22.3 20.4 -15.2 5.1 6.4 6.2 6.0 6.7 6.3 7.5 7.1 7.2 7.6 7.5 7.3 7.7 7.8 8.6 8.2 8.6 9.1 8.0 6.9 9.0 6.6 6.2 7.1 7.7 6.6 7.5 7.9 8.1 7.4 7.3 7.6 8.5 8.3 7.8 7.5 8.4 8.9 9.0 5.5 8.3 8.4 14.8 8.4 5.3 6.2 -- 59 -- Table 16 -- Continued JD-2400000 Velocity (m s−1) Uncertainty (m s−1) 54362.88624 54374.94279 54391.80959 54396.88855 54400.87136 54402.77756 54415.83289 54419.82914 54425.79502 54428.70304 4.0 12.9 9.0 20.4 12.8 -20.6 5.2 -20.9 -35.2 -4.9 6.4 6.6 8.3 7.2 7.6 7.6 9.1 9.0 8.4 8.0 Table 17. 2.7m Radial Velocities for HD 19994 JD-2400000 Velocity (m s−1) Uncertainty (m s−1) 53635.94301 53655.86581 53690.87053 53747.67357 54020.88376 54310.93898 54346.88504 54377.85942 54404.79890 54404.80349 54460.79143 54496.57966 -2.0 -33.6 34.6 52.8 -33.9 22.2 2.7 -19.8 -4.2 -9.8 14.2 -23.3 9.7 18.0 8.0 10.1 8.5 10.4 11.4 9.6 8.9 10.1 12.5 8.5 -- 60 -- Table 18. HET Radial Velocities for HD 20367 JD-2400000 Velocity (m s−1) Uncertainty (m s−1) 53581.93777 53592.91499 53594.92230 53605.87357 53607.87923 53608.87356 53610.85696 53612.85169 53633.80562 53651.98714 53653.98515 53669.94716 53678.93317 53685.89649 53691.88485 53694.88484 53696.65570 53701.64348 53703.85780 53705.84073 53708.83918 53710.83018 53713.60088 53723.55489 53725.79729 53727.78708 53728.78908 53730.56553 53730.76908 53748.73406 53749.72703 53753.71868 53758.70824 53942.95869 53950.94642 53954.92367 53956.92541 53958.90103 53960.92107 53970.88116 53976.88935 53979.86788 53979.87194 53984.84306 53988.84641 53993.83336 9.2 13.0 -4.3 -1.8 -3.5 15.5 7.5 -8.8 -26.8 -1.9 17.0 8.3 -12.8 -19.5 -5.1 0.6 -0.2 -13.6 -0.3 -6.1 8.6 4.5 -6.3 -25.1 -10.8 -17.0 -18.0 -9.7 -10.7 28.8 27.3 16.4 1.5 3.4 -6.3 17.5 9.3 22.5 18.4 17.9 9.2 5.7 -10.9 -14.4 3.2 19.4 8.3 11.9 11.7 10.8 10.3 10.5 10.8 12.3 10.2 10.3 10.7 11.2 11.1 10.6 13.9 11.8 11.7 11.1 11.9 12.9 13.9 13.2 11.8 13.6 11.4 11.6 12.7 12.1 11.3 11.0 10.7 12.7 12.8 11.2 10.3 11.4 11.4 11.7 12.1 12.9 10.9 12.9 11.1 10.8 10.1 11.3 -- 61 -- Table 18 -- Continued JD-2400000 Velocity (m s−1) Uncertainty (m s−1) 53997.82573 54001.81135 54018.98649 54021.74583 54035.71684 54044.92010 54051.88912 54055.66817 54059.65106 54065.64421 54067.62898 54071.84088 54071.84376 54073.83664 54075.60874 54133.68540 54141.66291 54147.65026 54153.61823 54167.59481 54169.57969 54174.58417 54321.93181 54347.86995 54379.00296 54394.73290 54397.72206 54399.71250 54403.94387 54409.90594 54419.90334 54419.90600 54425.65734 54427.64289 54427.64594 2.4 17.8 10.1 14.2 2.0 23.4 -6.1 19.9 -4.1 3.9 5.9 6.1 -17.6 6.6 -5.5 6.4 -2.2 6.0 6.0 -4.0 -14.2 7.7 -4.2 -9.4 -8.7 -11.1 -16.9 -10.4 -17.0 -18.6 -8.0 10.2 -5.1 -13.2 -23.0 10.6 11.2 10.9 10.7 11.7 12.0 12.4 11.6 11.2 11.4 12.2 11.9 11.9 13.0 13.5 12.2 10.8 10.5 11.3 13.4 12.0 11.4 10.7 10.9 11.1 11.9 11.3 10.2 10.0 12.4 13.7 15.3 10.6 12.3 11.2 -- 62 -- Table 19. 2.7m Radial Velocities for HD 20367 JD-2400000 Velocity (m s−1) Uncertainty (m s−1) 53632.00290 53635.96984 53691.80381 53808.62780 53967.90828 53968.90725 54018.98018 54158.59352 54189.59664 54189.61001 54190.61247 54191.60150 54192.60131 54345.84686 54377.85210 54402.79239 54460.80432 54496.58943 54555.60469 2.7 -3.7 -6.8 -7.9 11.5 11.4 -2.3 -5.9 -2.8 -6.9 -13.3 -1.9 4.3 0.1 -14.5 -3.6 -5.7 23.7 21.5 7.4 8.2 7.9 6.2 8.2 10.8 6.3 6.7 7.4 7.7 7.9 6.1 5.5 8.3 6.6 5.9 10.3 7.6 6.7 -- 63 -- Table 20. HET Radial Velocities for HD 23596 JD-2400000 Velocity (m s−1) Uncertainty (m s−1) 53581.96329 53592.92364 53593.94151 53594.92991 53605.89024 53607.89908 53608.88232 53609.89877 53627.84647 53629.84901 53636.81718 53668.74440 53677.71324 53677.71785 53680.71210 53682.70333 53691.89176 53694.67217 53696.66233 53701.64835 53703.87816 53708.84886 53710.83928 53712.85135 53712.85591 53734.78649 53741.76529 53748.74395 53800.58922 53956.94364 53958.92009 53960.92699 53969.92325 53973.90239 53976.89466 54057.66123 54059.65849 54064.88448 54066.88228 54068.65155 54071.64397 54073.86583 54084.59010 54092.57210 54094.56514 54130.71379 99.1 102.1 109.0 103.8 93.1 94.0 96.9 105.4 86.5 86.3 80.3 76.8 83.5 89.4 82.4 87.3 66.0 71.3 81.6 77.7 68.2 75.0 64.3 52.1 60.7 59.8 76.2 78.2 49.1 -5.3 -6.5 -10.2 -13.5 -9.6 -14.1 -45.2 -42.6 -53.3 -46.0 -53.9 -52.7 -12.4 -42.6 -48.2 -49.1 -58.2 8.3 8.2 8.1 8.0 9.1 8.1 8.8 8.7 8.5 8.6 10.1 10.0 9.6 9.8 8.9 9.5 12.4 11.8 11.1 11.7 12.6 11.6 12.1 13.9 13.1 10.9 10.7 11.5 10.4 8.6 9.3 7.4 13.9 11.6 9.0 9.8 9.6 9.3 9.8 11.4 10.7 11.1 11.3 11.7 12.3 8.8 -- 64 -- Table 20 -- Continued JD-2400000 Velocity (m s−1) Uncertainty (m s−1) 54136.69786 54147.66241 54156.63332 54159.61613 54328.92108 54370.81150 54392.99377 54394.74004 54400.73405 54403.72902 54411.92379 54419.91222 54419.91687 54419.92185 54425.66281 54427.65183 54427.65648 -52.2 -49.4 -65.8 -60.2 -98.8 -122.1 -128.6 -131.4 -124.0 -116.3 -111.3 -129.5 -126.1 -123.5 -114.9 -118.7 -120.0 11.0 10.1 10.7 11.3 9.0 7.9 7.9 9.0 8.9 11.0 9.6 10.3 11.0 10.3 9.1 11.3 9.9 Table 21. 2.7m Radial Velocities for HD 23596 JD-2400000 Velocity (m s−1) Uncertainty (m s−1) 53636.88979 53692.83650 53787.66076 53808.63976 54020.93260 54158.60541 73.6 48.7 11.9 13.0 -51.4 -95.8 7.1 7.6 6.0 7.1 8.2 6.1 -- 65 -- Table 22. HET Radial Velocities for HD 28185 JD-2400000 Velocity (m s−1) Uncertainty (m s−1) 53653.92112 53663.89377 53692.82400 53695.80664 53697.80395 53701.79069 53996.97502 53998.99146 54044.85384 54051.83487 54053.83800 54061.80571 54063.79720 54066.79838 54068.78387 54071.77898 54073.77510 54075.76724 54105.68544 54107.68662 54110.67889 54142.58869 54368.95856 54370.96073 54374.94882 54376.93838 54390.90631 54396.89578 54400.87895 54402.86922 54404.86864 54418.82544 54425.80631 54433.79452 -42.3 -38.0 24.8 42.7 34.2 43.3 -70.8 -64.8 -32.7 -17.9 -9.3 14.6 8.1 17.8 12.2 25.8 23.2 34.4 113.2 115.1 119.3 209.6 -53.6 -49.5 -50.1 -59.9 -43.2 -66.0 -53.6 -53.2 -44.3 -38.2 -38.5 -12.4 5.8 7.0 6.3 6.8 6.8 8.0 7.1 7.7 8.1 6.7 7.7 7.0 6.4 7.3 6.6 6.6 6.5 7.9 7.0 7.7 7.1 6.7 6.8 6.8 6.8 7.0 6.1 5.8 6.7 6.4 6.3 7.3 12.6 7.3 -- 66 -- Table 23. HET Radial Velocities for HD 38529 JD-2400000 Velocity (m s−1) Uncertainty (m s−1) 53341.50743 53355.84573 53357.85963 53358.72474 53359.72919 53360.84952 53365.81675 53367.81264 53369.70068 53371.68476 53377.78647 53379.67580 53389.75562 53390.76324 53391.75787 53392.75205 53395.73942 53414.69383 53416.68363 53708.89443 53709.88697 53711.76759 53712.87586 53724.84134 53730.71767 53731.70874 53733.70635 53735.71387 53739.69217 53742.68570 53751.77576 53752.76272 53753.77304 53754.76015 53755.75133 53757.63903 53758.75575 53764.74541 53989.99899 54020.92423 54021.92187 54022.92612 54028.90307 54031.92844 54035.00746 54035.88704 -186.9 -167.2 -181.3 -157.0 -144.0 -127.9 -82.4 -88.1 -158.4 -175.6 -92.2 -65.7 -101.2 -97.8 -88.0 -75.9 -57.5 -159.4 -136.9 2.7 -4.3 -45.3 -96.8 -4.7 -93.4 -77.7 -48.8 -20.3 -16.9 -107.9 24.2 22.8 11.6 -20.3 -61.2 -91.8 -87.3 6.2 21.1 81.0 95.7 101.0 28.2 55.3 105.3 111.6 3.2 5.3 5.4 5.1 5.9 5.4 4.6 5.3 5.4 5.3 6.2 4.6 4.9 4.7 5.1 5.1 4.8 7.0 5.4 4.8 4.5 4.2 5.6 4.5 4.9 4.7 4.7 5.0 4.5 4.4 4.8 4.7 5.2 4.8 4.5 4.3 5.5 4.4 4.5 4.0 4.3 7.9 4.4 3.4 4.1 4.8 -- 67 -- Table 23 -- Continued JD-2400000 Velocity (m s−1) Uncertainty (m s−1) 54037.87620 54039.86912 54040.97263 54043.85975 54048.89153 54051.84308 54052.83905 54053.84658 54054.82922 54056.92288 54060.91436 54061.91194 54062.80705 54063.80866 54071.89032 54072.77450 54073.89412 54075.75720 54081.86774 54100.83215 54105.80466 54109.80132 54110.69101 54128.72971 54132.72555 54133.71881 54163.63702 126.6 117.0 85.6 32.4 96.4 120.3 132.7 132.0 112.6 33.4 53.1 70.9 85.5 102.5 37.3 34.0 32.3 70.5 169.9 43.0 122.6 170.3 162.4 71.1 101.9 112.2 153.6 4.2 4.5 4.3 4.4 3.4 4.5 4.5 4.7 7.2 4.8 4.7 4.6 4.4 4.2 4.1 4.4 4.5 3.9 4.8 5.5 4.7 4.8 4.7 4.6 6.5 4.3 4.7 Table 24. 2.7m Radial Velocities for HD 38529 JD-2400000 Velocity (m s−1) Uncertainty (m s−1) 53633.96726 53636.91593 53691.91356 53746.78728 53809.64793 53984.94973 54020.90382 -44.5 -19.0 -24.9 -51.1 47.7 -1.4 93.2 5.7 4.9 5.8 5.7 5.5 5.1 6.4 -- 68 -- Table 25. HET Radial Velocities for HD 40979 JD-2400000 Velocity (m s−1) Uncertainty (m s−1) 53341.72130 53346.73432 53348.70934 53350.72298 53352.92178 53355.68899 53357.93997 53359.68105 53365.67253 53367.66116 53370.62842 53372.64584 53377.85739 53379.86066 53381.63007 53383.84999 53389.59831 53391.57735 53395.58085 53399.80045 53401.80944 53416.73452 53422.72494 53424.74945 53429.74027 53444.68952 53615.96771 53623.94032 53628.94576 53629.92910 53633.91493 53638.90545 53646.89729 53651.88585 53655.87437 53663.84616 53666.83032 53668.81351 53669.82586 53676.82778 53678.79984 53681.80384 53683.80383 53685.79896 53687.78847 53689.78041 -80.8 -99.4 -57.5 -81.2 -95.5 -48.4 -59.4 -88.5 -38.1 -10.1 11.0 55.1 20.0 48.2 69.2 77.1 67.5 84.0 117.5 104.5 131.8 97.1 105.6 71.9 89.5 91.8 -46.1 -20.7 -22.1 2.9 15.2 11.3 29.0 71.1 70.8 91.7 115.4 122.5 88.2 98.7 84.0 106.4 94.1 108.1 61.6 95.5 13.1 15.1 17.1 14.7 15.0 13.4 16.1 16.3 14.9 15.9 16.9 42.5 23.3 15.6 13.3 14.6 15.8 16.8 13.1 12.1 13.1 13.1 11.1 12.6 11.6 13.9 13.0 13.6 12.8 11.3 12.5 11.7 28.8 11.3 11.7 12.3 13.2 13.4 11.9 16.4 13.5 13.7 14.3 12.9 13.0 13.6 -- 69 -- Table 25 -- Continued JD-2400000 Velocity (m s−1) Uncertainty (m s−1) 53691.77056 53693.99549 53696.97730 53701.75161 53703.96738 53705.96824 53708.94157 53710.94020 53713.94615 53713.94937 53721.90230 53723.90854 53728.67332 53730.65950 53734.65127 53743.62390 53748.60983 53753.84141 53799.71352 53801.70827 53987.95684 54014.89116 54021.86435 54037.82927 54044.80399 54053.00792 54054.99336 54057.99292 54061.01007 54068.73628 54076.71914 54101.86096 54129.79577 54132.78782 54134.78060 54136.76258 54155.73508 54166.70337 54177.65752 54190.62821 54370.91031 54397.83722 54402.82966 54414.00568 54419.02519 98.2 59.6 80.3 26.3 60.3 31.7 10.9 1.8 35.2 30.9 -23.6 3.6 -12.5 -18.9 -9.3 -1.6 -11.3 -54.3 -93.5 -96.0 -7.1 -75.8 -72.7 -99.6 -137.4 -124.9 -103.5 -114.8 -100.4 -120.0 -122.6 -147.8 -108.9 -94.0 -90.7 -89.7 -29.7 -0.7 39.5 119.0 -128.8 -83.0 -77.0 -71.8 -36.0 12.4 12.4 11.9 12.3 13.4 14.7 14.1 12.3 14.1 12.6 15.9 12.7 14.8 14.1 13.7 17.0 12.4 12.7 13.7 11.9 14.3 12.4 13.0 14.2 14.1 11.7 13.0 11.9 11.8 13.4 14.2 15.7 13.2 14.4 12.7 12.9 13.4 12.3 12.8 11.9 13.0 11.5 11.4 13.1 12.0 -- 70 -- -- 71 -- Table 26. 2.7m Radial Velocities for HD 40979 JD-2400000 Velocity (m s−1) Uncertainty (m s−1) 53636.00527 53787.77595 53864.61096 54020.94453 78.9 -46.2 -23.4 -9.2 8.9 8.3 7.6 6.7 -- 72 -- Table 27. HET Radial Velocities for HD 72659 JD-2400000 Velocity (m s−1) Uncertainty (m s−1) 53342.98059 53346.89816 53348.89792 53351.88711 53355.85531 53357.86778 53359.86322 53366.92023 53370.82512 53375.89102 53377.89403 53379.88562 53383.88589 53389.78303 53391.78123 53395.75846 53399.83667 53401.82272 53408.71984 53416.69679 53422.69865 53424.76127 53429.75413 53439.64548 53446.69267 53447.69980 53448.70841 53708.90029 53710.89277 53713.96630 53723.86387 53728.92778 53734.90001 53742.80266 53746.79841 53751.79314 53753.79398 53755.85771 53764.75640 53773.72893 53780.78984 53802.64977 54050.97483 54053.02653 54056.95675 54061.02215 8.4 19.2 17.7 20.4 21.5 15.6 12.3 1.9 3.3 35.3 11.8 16.6 10.5 25.1 21.5 16.3 10.1 3.1 21.5 20.8 3.1 17.1 19.1 10.2 20.1 2.7 4.4 3.7 -4.4 2.2 2.1 -15.9 -15.7 -16.7 2.7 -10.5 -4.9 -8.2 -9.2 6.3 -1.5 6.8 -25.6 -18.5 -32.3 -29.7 11.7 10.5 11.3 9.6 9.2 9.7 11.8 12.2 10.6 13.6 10.2 10.7 9.5 12.8 10.4 10.9 10.3 9.3 9.7 9.5 8.3 9.4 9.3 8.7 8.3 8.3 8.0 9.0 8.8 10.0 8.8 9.2 11.2 10.6 9.6 10.3 10.4 10.6 9.7 11.0 9.4 8.3 9.5 9.5 9.9 9.6 -- 73 -- Table 27 -- Continued JD-2400000 Velocity (m s−1) Uncertainty (m s−1) 54064.00548 54127.75916 54158.75638 54161.66311 54167.72203 54420.02055 54431.99247 -22.3 -29.3 -31.1 -21.2 -25.1 -50.6 -40.2 10.3 8.4 8.1 9.8 8.6 11.1 9.4 -- 74 -- Table 28. HET Radial Velocities for HD 74156 JD-2400000 Velocity (m s−1) Uncertainty (m s−1) 53342.23249 53347.00129 53355.83378 53357.84523 53359.85043 53360.97444 53364.97638 53365.82597 53367.82107 53383.92077 53390.75248 53448.73888 53451.73096 53476.64913 53480.63719 53481.63192 53482.63133 53664.99582 53675.97107 53676.98443 53682.95189 53687.93153 53689.92922 53691.91679 53697.91227 53703.88616 53708.88180 53710.87934 53718.01348 53724.82993 53728.82485 53731.96775 53733.80186 53734.80949 53736.94379 53741.78326 53742.78352 53743.79031 53748.77268 53751.76772 53753.76343 53754.74931 53756.74864 53764.73571 53832.67552 53833.69639 -119.8 -102.6 -93.1 -90.2 -79.3 -97.9 -99.2 -121.7 -127.5 -128.5 -107.2 -19.3 -31.0 -151.4 -167.2 -127.9 -96.2 75.5 61.4 53.1 -46.4 -54.2 -1.3 27.3 50.3 71.2 81.0 81.4 62.8 70.8 47.9 16.0 -33.8 -58.8 -138.8 3.4 10.0 33.9 67.4 84.5 99.6 90.1 92.3 90.9 50.9 52.0 7.9 7.5 7.9 8.0 10.8 9.0 10.4 7.8 8.8 8.5 8.0 6.7 7.2 7.7 22.9 7.4 7.5 7.0 7.8 8.9 8.3 7.3 7.5 7.9 7.7 9.3 8.6 7.7 10.1 8.2 7.9 9.3 9.2 8.6 9.2 8.5 8.1 8.3 8.7 8.8 8.3 7.9 8.3 9.3 7.2 8.1 -- 75 -- Table 28 -- Continued JD-2400000 Velocity (m s−1) Uncertainty (m s−1) 53834.67429 53835.66675 53838.66377 53841.64454 53845.63199 53846.65288 54029.98957 54035.99026 54038.97806 54039.97014 54040.95851 54043.96476 54044.95388 54050.95326 54051.94621 54052.93930 54073.88022 54079.86417 54087.84388 54090.83839 54106.78411 54109.78878 54110.79989 54129.87053 54130.74036 54133.84698 54134.72603 54135.86669 54136.84113 54148.67767 54156.66058 54159.77984 54166.76336 54167.75867 54211.63030 54231.60002 33.9 12.8 -82.9 -93.6 17.1 34.0 56.8 47.8 37.1 30.0 18.0 -43.3 -92.8 -35.4 -16.1 1.4 56.5 63.8 53.6 28.0 9.3 20.6 26.4 51.6 48.8 36.1 46.7 43.7 43.9 -128.1 -1.7 24.0 45.8 40.7 7.8 58.8 7.5 7.4 7.3 11.0 7.7 7.7 7.6 8.2 7.5 7.6 7.7 6.8 6.7 7.7 8.4 7.0 6.8 7.2 6.8 7.0 8.6 7.1 8.3 8.8 8.6 8.9 8.9 8.5 9.7 10.6 9.3 8.8 7.9 8.1 7.9 6.4 -- 76 -- Table 29. HET Radial Velocities for HD 80606 JD-2400000 Velocity (m s−1) Uncertainty (m s−1) 53346.88103 53358.02089 53359.82400 53361.02985 53365.03079 53373.98282 53377.80112 53379.75230 53389.74170 53391.74400 53395.72763 53399.72518 53401.72497 53414.67819 53421.85529 53423.86650 53424.85231 53432.87120 53433.60628 53446.79322 54161.85400 54166.83797 54186.76189 8.1 -16.0 -26.7 -28.9 -46.8 -63.3 -62.8 -62.1 -95.6 -97.6 -104.1 -115.4 -123.7 -186.7 314.1 374.2 302.6 132.0 119.8 55.8 -62.4 -80.2 -134.1 8.6 8.1 8.8 7.7 8.0 9.8 9.6 9.3 7.8 8.0 8.6 9.4 9.1 9.1 8.1 7.9 7.3 7.6 7.6 8.2 8.9 7.3 7.4 -- 77 -- Table 30. HET Radial Velocities for HD 89744 JD-2400000 Velocity (m s−1) Uncertainty (m s−1) 53709.89685 53723.85367 53727.84573 53734.81973 53736.82101 53738.03441 53738.81040 53742.79299 53746.81778 53751.78379 53753.78381 53755.76218 53757.77181 53797.64834 53809.62700 53837.77709 53866.70329 53868.68562 53875.66956 53883.65837 53890.63954 53893.63138 54050.96453 54052.96762 54056.94785 54063.93165 54073.91476 54122.01243 54129.74491 54160.66031 54163.66643 54165.88148 54421.94999 93.4 57.9 70.0 64.1 60.8 76.6 76.7 66.5 51.1 53.7 45.4 47.2 35.1 -165.1 -353.0 1.2 97.6 100.3 102.4 114.7 107.9 125.0 -137.3 -160.3 -197.9 -327.2 -438.1 102.1 94.8 98.0 89.0 98.8 75.7 11.8 8.7 9.0 9.1 9.9 9.8 9.6 9.0 8.7 10.7 8.7 9.2 9.1 10.0 8.7 7.9 7.5 13.0 11.6 6.5 7.1 7.2 7.5 6.9 7.6 7.9 9.1 9.5 10.1 9.6 10.0 15.3 8.1 Table 31. 2.7m Radial Velocities for HD 89744 JD-2400000 Velocity (m s−1) Uncertainty (m s−1) 53690.03080 53805.87856 53806.73923 53807.83562 53809.79691 53840.78664 53864.76543 53911.61226 54068.94214 262.2 -129.3 -130.0 -188.4 -198.1 168.0 203.8 252.5 -240.7 7.7 7.1 8.9 10.8 7.8 9.8 8.9 9.3 8.7 -- 78 -- Table 32. HET Radial Velocities for 47 UMa JD-2400000 Velocity (m s−1) Uncertainty (m s−1) 53313.33003 53314.99108 53317.99083 53334.95254 53335.94709 53338.93789 53340.91724 53346.92207 53348.90940 53350.91700 53357.88009 53359.87543 53365.86490 53367.86390 53371.85734 53373.85855 53377.83382 53379.85077 53389.79762 53391.79286 53395.77820 53400.99471 53408.76968 53414.72833 53416.71040 53421.94116 53423.70482 53429.91553 53432.90803 53433.90682 53437.66207 53439.65955 53440.90030 53476.80401 53479.77845 53481.76620 53486.77731 53488.76788 53512.69186 53526.63730 53539.63923 53708.92005 53709.92253 53710.91317 53711.93277 24.4 31.8 30.8 26.0 26.8 20.1 24.8 17.7 24.0 14.7 4.3 7.9 -11.9 1.0 -1.0 1.0 3.4 -4.9 4.0 -6.9 -4.5 -3.3 -1.7 -10.6 -9.0 -7.1 -13.1 -7.8 -5.9 -7.8 -6.9 -7.0 -13.2 -20.0 -16.3 -16.9 -20.7 -20.1 -12.1 -15.7 -14.5 -27.1 -29.2 -29.6 -26.4 4.7 4.6 4.4 4.3 4.4 3.1 4.4 4.3 4.7 8.2 5.0 5.6 4.1 4.3 5.1 7.9 4.4 4.5 4.2 5.0 4.4 4.7 4.5 4.3 4.7 3.8 3.5 3.2 3.3 3.2 2.8 3.2 3.7 3.3 3.2 3.3 3.4 3.2 3.3 2.6 3.6 4.8 4.7 4.2 3.6 -- 79 -- Table 32 -- Continued JD-2400000 Velocity (m s−1) Uncertainty (m s−1) 53721.88029 53723.87033 53725.86147 53734.87812 53736.84137 53738.82751 53742.82008 53743.82024 53744.82292 53746.80758 53751.79987 53757.03749 53771.76109 53775.74040 53777.96664 53779.96405 53786.70738 53795.91970 53797.66773 53894.65514 53901.64093 54050.99305 54052.99278 54054.97972 54056.96223 54121.81209 54129.77513 54157.70092 54160.68446 54165.91204 54419.99530 54431.94398 -19.6 -29.2 -21.3 -39.1 -30.0 -24.4 -31.4 -36.6 -26.8 -23.2 -26.4 -29.7 -18.3 -16.1 -16.3 -21.1 -16.1 -17.0 -5.6 18.2 21.6 53.7 65.9 53.9 47.0 47.7 51.6 70.7 63.1 64.2 17.4 9.9 4.9 4.8 4.8 4.9 4.7 4.5 4.7 5.4 4.4 4.5 5.0 4.2 4.8 4.1 4.6 4.3 5.0 3.0 5.4 3.6 3.8 4.0 4.0 4.2 4.2 4.8 4.9 4.7 4.8 4.3 4.3 4.3 -- 80 -- Table 33. 2.7m Radial Velocities for 47 UMa JD-2400000 Velocity (m s−1) Uncertainty (m s−1) 51010.62898 51212.97474 51240.81250 51274.78993 51326.70558 51504.95996 51530.01978 51555.94972 51655.74023 51686.75156 51750.60418 51861.01895 51917.93086 51987.85527 52004.83235 52039.77936 52116.60554 52249.00010 52303.89238 52305.84757 52327.86285 52353.85949 52661.95399 53017.93695 53069.76686 53692.03243 53748.89147 53787.91198 53805.88756 53809.80777 53840.77154 53841.70627 53841.72168 53861.74397 53909.61977 54280.64401 54280.64893 54551.92162 54569.78354 54569.78813 54569.79271 54605.77977 54632.71638 52.2 1.7 -1.7 -8.8 -24.5 -39.7 -27.0 -15.5 -0.7 -8.8 13.1 53.7 48.2 53.7 60.5 54.5 55.7 21.6 -8.3 -1.5 21.2 2.6 -30.8 75.2 54.9 -44.4 -40.4 -28.0 -28.3 -25.0 -19.2 -19.3 0.2 -19.8 3.6 38.2 31.4 -36.6 -39.5 -37.8 -40.3 -52.7 -43.6 6.2 4.5 4.6 4.6 4.9 5.0 5.1 4.8 4.7 6.0 5.1 4.9 4.6 6.2 4.8 5.9 6.6 5.4 5.1 4.7 13.1 6.1 4.7 6.4 4.2 6.5 4.6 4.7 4.5 4.7 5.1 5.8 6.5 5.4 5.7 6.7 5.6 5.2 5.8 5.9 6.1 6.5 6.9 -- 81 -- Table 34. HET Radial Velocities for HD 106252 JD-2400000 Velocity (m s−1) Uncertainty (m s−1) 53351.00010 53392.87552 53396.02801 53399.86570 53422.81038 53423.95949 53424.95405 53429.93417 53436.93141 53439.93691 53440.93473 53448.72404 53449.71116 53450.72302 53451.72210 53452.88054 53454.86997 53455.86893 53457.70319 53480.62813 53487.78674 53498.76669 53543.64221 53736.92178 53736.92786 53743.93096 53744.90904 53745.89603 53753.89366 53755.88492 53758.87460 53765.02009 53779.82267 53796.76996 53866.75798 53868.73636 53877.71039 53891.67012 54090.96158 54110.90994 54122.04519 54161.76867 54191.69138 -86.7 -19.4 -18.2 -0.6 35.5 38.5 30.2 43.4 41.8 44.2 46.8 54.7 60.8 50.1 49.4 66.6 56.3 64.1 60.4 72.9 69.1 82.5 92.8 -13.3 -9.9 1.6 -12.4 -11.5 -17.2 -24.8 -32.4 -25.5 -24.6 -34.1 -41.3 -39.5 -35.1 -40.6 -100.9 -106.5 -118.5 -124.4 -124.3 10.2 11.1 10.0 11.5 8.9 9.9 8.4 8.3 8.3 9.3 10.0 8.7 9.4 8.7 8.3 8.6 8.8 8.8 8.6 10.0 9.0 9.9 8.7 10.6 9.9 11.4 9.7 9.8 9.3 9.0 9.9 11.8 10.3 8.4 10.2 9.5 8.4 8.4 8.1 8.8 9.6 9.8 8.6 -- 82 -- Table 35. 2.7m Radial Velocities for HD 106252 JD-2400000 Velocity (m s−1) Uncertainty (m s−1) 52116.61921 52307.00335 52328.89167 52357.80730 52743.85538 53465.74189 53504.65833 53564.63133 53566.62685 53808.85919 53842.77638 53861.78041 46.4 16.9 -32.1 -40.7 -129.6 55.7 73.1 73.7 84.6 -23.7 -24.6 -35.8 11.5 9.8 11.8 9.6 9.1 10.4 9.8 10.5 10.8 10.3 10.9 10.2 -- 83 -- Table 36. HET Radial Velocities for HD 108874 JD-2400000 Velocity (m s−1) Uncertainty (m s−1) 53370.93479 53377.90734 53392.86111 53399.84978 53423.00572 53424.00026 53429.76827 53446.93347 53449.92196 53451.70786 53452.91476 53454.90261 53455.91231 53457.89845 53460.88699 53708.98585 53723.95560 53730.92665 53751.89671 53753.88047 53755.87205 53773.81280 53806.96496 53847.61993 53866.78797 53868.77287 53880.76028 53895.71194 53897.70685 53912.66615 54080.96833 54084.96842 54109.89738 54127.85761 54142.02401 54158.76779 54160.75928 54162.75513 54180.92025 54190.88453 -4.1 3.4 -3.8 4.9 1.5 -1.7 9.2 3.5 -0.3 4.7 -0.6 -10.7 -2.2 -9.3 -5.9 -35.6 -17.0 -15.4 -3.0 -3.1 9.2 11.5 3.7 5.4 -3.3 -6.4 -3.9 -5.6 -3.1 -12.8 -24.1 -25.8 -1.9 3.1 17.2 28.1 23.2 19.0 23.0 29.0 7.6 8.0 8.3 7.1 6.0 6.4 5.6 6.4 5.6 5.5 5.0 6.1 5.6 5.8 6.0 7.5 7.7 7.5 7.1 7.8 6.3 6.7 6.8 6.1 6.2 5.8 6.5 5.3 5.7 5.3 7.0 7.1 7.0 6.8 6.6 7.7 6.7 7.0 5.8 6.2 -- 84 -- Table 37. HET Radial Velocities for HD 114783 JD-2400000 Velocity (m s−1) Uncertainty (m s−1) 53366.02839 53368.02525 53374.00882 53378.99750 53395.95201 53399.94906 53415.90535 53416.96080 53421.88527 53423.88025 53429.88531 53436.92048 53440.91863 53446.81643 53447.81352 53448.81555 53450.81211 53451.82012 53452.81032 53455.78717 53779.91198 54097.02928 54106.00717 54121.96750 54127.95124 54130.94440 54140.98702 54143.92794 54156.87064 54158.87301 54163.85683 54168.83901 54186.78882 54191.77547 -17.8 -25.3 -13.9 -16.3 -23.9 -25.0 -27.0 -27.8 -21.5 -29.5 -19.4 -22.8 -24.7 -20.4 -16.1 -13.6 -18.4 -22.9 -16.8 -20.4 6.9 27.2 32.2 24.7 28.9 35.9 26.0 21.4 35.6 34.8 38.2 42.2 32.9 36.7 6.0 5.5 6.5 5.3 5.4 5.2 6.2 5.7 5.5 5.9 5.7 5.6 6.9 5.2 5.5 5.6 5.5 5.2 5.5 5.3 5.7 5.7 6.0 5.3 5.3 5.3 5.7 6.0 5.8 5.8 5.8 5.6 5.0 5.2 -- 85 -- Table 38. HET Radial Velocities for HD 128311 JD-2400000 Velocity (m s−1) Uncertainty (m s−1) 53462.96527 53479.81567 53480.91902 53482.89103 53486.72671 53488.70965 53509.81664 53510.82254 53511.80800 53512.80394 53539.74332 53541.75663 53542.74601 53544.73511 53550.71539 53554.69916 53565.69081 53570.66963 53776.92892 53788.90110 53816.97394 53824.80309 53837.91730 53842.75039 53844.91973 53846.91450 53888.79633 53895.75921 53897.76919 53899.76175 53911.71053 53926.66439 53933.66149 54108.00385 54110.00828 54111.00546 54130.96180 54133.93944 54135.94222 54138.92468 54139.92656 54141.92573 54156.87834 54157.88752 54158.88100 54160.87801 -137.0 -147.0 -136.9 -144.3 -159.9 -168.8 -160.8 -172.4 -176.2 -159.5 -177.2 -152.9 -159.5 -163.9 -162.3 -165.3 -159.0 -124.8 -55.4 -65.9 -81.9 -88.4 -115.2 -53.5 -87.7 -118.5 -92.8 -149.3 -106.0 -76.0 -80.0 -131.2 -56.4 1.0 9.5 -1.6 11.2 9.9 34.9 14.5 -4.1 21.5 49.1 62.4 47.6 21.5 4.1 4.4 5.7 5.7 5.5 6.4 6.0 5.6 6.0 5.6 4.7 5.7 5.3 4.8 5.8 5.4 6.1 6.3 6.3 7.7 6.2 6.2 5.9 6.4 5.7 5.0 5.1 5.0 4.8 4.6 5.4 6.4 4.9 4.4 3.7 3.9 3.8 4.5 4.2 4.0 4.0 3.7 4.5 5.0 4.5 4.5 -- 86 -- Table 38 -- Continued JD-2400000 Velocity (m s−1) Uncertainty (m s−1) 54161.86725 54163.86833 54164.85995 54168.84642 54173.87121 54174.99073 54177.00300 54186.80132 54190.80340 54191.87047 54211.91614 54214.71537 54216.72250 54217.87848 54221.70097 54222.71695 54223.88092 54231.85098 54232.85136 54249.79485 54250.79070 54251.63068 54253.78617 54254.79359 54255.78044 54257.77841 54265.75347 54267.74859 54276.71906 54278.71354 54279.72454 54318.61493 19.6 41.8 54.8 70.2 31.0 50.0 62.6 42.8 69.2 70.0 59.3 69.1 41.2 31.4 59.8 58.2 64.8 64.0 53.4 41.0 39.8 19.8 46.3 39.2 33.1 50.0 22.7 31.4 0.8 3.1 11.9 -83.4 4.0 3.9 6.2 6.2 3.2 5.2 3.5 3.5 3.6 3.7 4.0 4.1 3.5 3.7 3.5 3.3 3.6 3.4 3.2 3.5 3.2 3.3 3.1 3.2 3.5 3.2 3.2 3.5 3.6 3.2 2.7 5.3 -- 87 -- Table 39. HET Radial Velocities for HD 130322 JD-2400000 Velocity (m s−1) Uncertainty (m s−1) 53471.80558 53481.88526 53486.85864 53488.75815 53509.79117 53512.78123 53527.74971 53542.69985 53543.70614 53550.70420 53837.89677 53842.88880 53868.80896 53882.78043 53897.72683 53900.72079 53936.63557 54122.01834 54128.00335 54135.98084 54139.97029 54140.96840 54144.96962 54157.01611 54158.92425 54163.92465 54168.90656 54173.98269 54176.87914 54191.92631 -99.8 -106.9 105.1 72.2 101.3 -65.6 27.0 55.4 -4.9 105.5 -12.6 27.6 -78.8 83.0 -44.2 -85.4 110.4 -12.7 47.0 -113.2 89.9 98.1 -99.4 -112.4 -40.9 26.3 -71.6 69.7 -90.6 20.5 7.2 6.6 6.4 5.9 6.1 5.1 6.2 5.6 6.1 6.1 5.9 6.3 5.7 6.0 6.1 6.0 6.6 6.8 6.7 6.5 7.2 6.1 6.6 6.8 6.8 6.7 6.5 7.4 5.7 6.2 Table 40. 2.7m Radial Velocities for HD 130322 JD-2400000 Velocity (m s−1) Uncertainty (m s−1) 53585.64900 53843.89253 53863.78301 53910.78043 54251.84318 83.7 -18.0 75.5 -68.5 -72.8 7.5 7.5 8.6 8.1 9.4 -- 88 -- Table 41. HET Radial Velocities for HD 136118 JD-2400000 Velocity (m s−1) Uncertainty (m s−1) 53471.83434 53480.89983 53482.88067 53486.86878 53487.88314 53509.80853 53527.76309 53544.72664 53575.62972 53755.05078 53757.04207 53758.03539 53766.02629 53767.02021 53769.01091 53778.99248 53779.98920 53787.98168 53803.01595 53805.98952 53808.90657 53809.90878 53815.88767 53816.93136 53818.87257 53820.89349 53832.83912 53835.85292 53836.85837 53840.89490 53842.89836 53844.90864 53866.77360 53867.75385 53868.75220 53877.74161 53880.80832 53883.77783 53888.69980 53890.67902 53891.68247 53892.68903 53893.76811 53895.74381 53897.74893 53898.67744 37.0 11.8 8.1 12.1 17.0 15.5 5.7 4.5 1.1 -101.1 -101.7 -88.4 -101.5 -87.2 -95.3 -97.5 -103.4 -92.7 -97.6 -85.6 -92.2 -93.7 -71.7 -81.2 -73.7 -80.5 -83.6 -80.7 -73.5 -81.0 -75.4 -68.2 -63.9 -71.4 -62.9 -68.8 -87.9 -76.4 -75.5 -52.4 -58.2 -68.8 -76.9 -47.5 -69.6 -73.1 7.8 11.6 9.0 14.7 9.0 12.8 9.4 9.3 9.4 11.1 10.1 15.3 9.7 16.8 8.8 14.3 15.6 8.9 13.2 13.7 7.6 10.3 8.2 6.9 9.2 8.8 9.2 9.6 13.4 8.6 14.4 7.9 8.0 7.7 11.9 7.3 8.9 7.5 8.0 8.2 7.8 7.9 7.5 15.6 7.5 8.7 -- 89 -- Table 41 -- Continued JD-2400000 Velocity (m s−1) Uncertainty (m s−1) 53901.74015 53905.73368 53911.73001 53917.68891 53937.64758 53939.63048 54129.03715 54131.02378 54144.99774 54164.01886 54176.99293 54180.88946 54186.88695 54190.87336 54211.81162 54221.78932 54253.69912 54282.63079 -66.1 -58.2 -42.9 -35.2 -71.9 -64.2 172.5 179.1 185.3 249.5 279.5 287.4 309.8 306.2 313.7 331.8 346.9 329.9 8.3 9.0 8.7 13.9 12.8 13.2 8.6 9.8 13.1 12.5 7.4 8.3 7.9 15.1 8.3 8.6 6.9 8.9 Table 42. 2.7m Radial Velocities for HD 136118 JD-2400000 Velocity (m s−1) Uncertainty (m s−1) 53585.66699 53805.93741 53863.76953 53911.75165 13.5 -1.7 -12.0 0.2 11.4 9.1 13.8 10.0 -- 90 -- Table 43. HET Radial Velocities for HD 178911B JD-2400000 Velocity (m s−1) Uncertainty (m s−1) 53653.66528 53801.00302 53803.02412 53837.93570 53846.89693 53866.83649 53868.81760 53883.79325 53954.82320 53955.82606 53956.82027 53958.81704 53960.80356 53965.80165 53966.78694 53971.77135 53976.76788 53979.76859 53988.72794 53988.73584 53993.71885 54014.66291 54016.65503 54035.60988 54039.58877 54055.55127 54063.54556 54165.01752 54167.00332 54190.93933 54251.77452 54323.81680 54332.79368 54335.79347 54338.78881 54340.77677 54344.77039 54365.69757 54396.61972 54400.59558 -57.4 -217.2 -268.6 257.0 323.3 -19.7 -86.7 -343.1 -349.3 -336.2 -326.8 -291.4 -245.8 -102.8 -69.2 72.4 181.7 242.2 328.9 330.3 306.6 -175.4 -236.1 -147.1 -29.0 302.5 327.1 -354.1 -362.1 182.1 -106.8 -86.4 173.3 228.2 275.7 301.3 324.4 22.9 -44.4 75.6 6.2 6.1 6.2 6.7 5.6 5.6 6.2 4.5 5.7 5.4 5.8 5.5 5.5 5.4 5.3 5.6 5.7 6.0 6.0 5.8 7.7 5.5 5.4 5.1 5.2 5.6 5.7 6.4 6.4 6.4 5.5 5.5 7.9 6.0 6.3 6.1 6.3 6.3 5.5 5.5 -- 91 -- Table 44. HET Radial Velocities for HD 190228 JD-2400000 Velocity (m s−1) Uncertainty (m s−1) 53581.89473 53589.86259 53605.81074 53606.81111 53607.80141 53609.80786 53628.75754 53635.71968 53653.67654 53655.68368 53686.58579 53844.93852 53867.88496 53877.84874 53883.82777 53888.81145 53897.79760 53935.90971 53956.84836 53966.81728 53976.79348 53996.75947 53998.73191 53998.73697 54008.70583 54013.70146 54019.66975 54032.62480 54217.92567 54265.79081 54284.96126 54326.62326 54328.62216 54331.61953 54336.59244 54344.79398 54352.77160 54368.71625 54370.72492 54377.68917 54401.62808 54428.57314 -75.5 -92.8 -83.7 -84.1 -84.9 -87.7 -85.0 -76.7 -81.1 -76.3 -61.1 -20.2 -34.8 -19.3 -12.2 -8.6 1.5 6.8 -9.1 -0.8 -2.1 5.3 7.8 -1.5 10.4 1.4 7.1 16.7 49.1 59.8 62.5 57.0 60.8 68.4 67.8 60.7 68.1 75.5 81.4 76.2 74.4 78.5 8.2 6.8 9.0 8.8 8.4 8.1 7.8 8.4 7.5 8.3 9.0 9.3 9.3 9.1 8.2 9.4 8.9 8.5 7.5 8.3 8.4 9.0 9.4 9.0 7.8 8.1 9.0 9.9 8.9 8.6 9.1 8.5 7.0 8.0 9.0 8.7 9.1 8.9 8.4 9.0 8.3 8.3 -- 92 -- Table 45. 2.7m Radial Velocities for HD 190228 JD-2400000 Velocity (m s−1) Uncertainty (m s−1) 53584.82785 53585.80785 53635.70876 53636.74761 53691.63145 53862.94967 53927.85541 54403.65001 -47.3 -27.8 -42.5 -41.8 -16.2 14.5 43.2 118.0 6.2 8.9 6.9 5.5 7.9 6.4 6.8 7.1
1708.06069
2
1708
2017-10-26T01:55:13
Likely Transiting Exocomets Detected by Kepler
[ "astro-ph.EP" ]
We present the first good evidence for exocomet transits of a host star in continuum light in data from the Kepler mission. The Kepler star in question, KIC 3542116, is of spectral type F2V and is quite bright at K_p = 10. The transits have a distinct asymmetric shape with a steeper ingress and slower egress that can be ascribed to objects with a trailing dust tail passing over the stellar disk. There are three deeper transits with depths of ~0.1% that last for about a day, and three that are several times more shallow and of shorter duration. The transits were found via an exhaustive visual search of the entire Kepler photometric data set, which we describe in some detail. We review the methods we use to validate the Kepler data showing the comet transits, and rule out instrumental artefacts as sources of the signals. We fit the transits with a simple dust-tail model, and find that a transverse comet speed of ~35-50 km/s and a minimum amount of dust present in the tail of ~10^16 g are required to explain the larger transits. For a dust replenishment time of ~10 days, and a comet lifetime of only ~300 days, this implies a total cometary mass of > 3 x 10^17 g, or about the mass of Halley's comet. We also discuss the number of comets and orbital geometry that would be necessary to explain the six transits detected over the four years of Kepler prime-field observations. Finally, we also report the discovery of a single comet-shaped transit in KIC 11084727 with very similar transit and host-star properties.
astro-ph.EP
astro-ph
Preprint typeset using LATEX style emulateapj v. 12/16/11 LIKELY TRANSITING EXOCOMETS DETECTED BY KEPLER S. Rappaport1, A. Vanderburg2,3,4, T. Jacobs5, D. LaCourse6, J. Jenkins7, A. Kraus8, A. Rizzuto8, D. W. Latham2, A. Bieryla2, M. Lazarevic9, A. Schmitt10 ABSTRACT We present the first good evidence for exocomet transits of a host star in continuum light in data from the Kepler mission. The Kepler star in question, KIC 3542116, is of spectral type F2V and is quite bright at Kp = 10. The transits have a distinct asymmetric shape with a steeper ingress and slower egress that can be ascribed to objects with a trailing dust tail passing over the stellar disk. There are three deeper transits with depths of (cid:39) 0.1% that last for about a day, and three that are several times more shallow and of shorter duration. The transits were found via an exhaustive visual search of the entire Kepler photometric data set, which we describe in some detail. We review the methods we use to validate the Kepler data showing the comet transits, and rule out instrumental artefacts as sources of the signals. We fit the transits with a simple dust-tail model, and find that a transverse comet speed of ∼35-50 km s−1 and a minimum amount of dust present in the tail of ∼ 1016 g are required to explain the larger transits. For a dust replenishment time of ∼10 days, and a comet lifetime of only ∼300 days, this implies a total cometary mass of (cid:38) 3 × 1017 g, or about the mass of Halley's comet. We also discuss the number of comets and orbital geometry that would be necessary to explain the six transits detected over the four years of Kepler prime-field observations. Finally, we also report the discovery of a single comet-shaped transit in KIC 11084727 with very similar transit and host-star properties. Subject headings: comets-general - minor planets, asteroids - (stars:) planetary systems - stars: individual (KIC 3542116, KIC 11084727) 1. INTRODUCTION Advances in both space-based missions and ground- based observational techniques over the past dozen years have led to a huge expansion in the number of confirmed exoplanet detections. Currently, there are over 3500 ex- oplanets confirmed to orbit a variety of host star spec- tral types. Growing catalogs of short-period transiting exoplanets derived from data returned by the CoRoT (Baglin et al. 2006) and Kepler (Borucki et al. 2010) spacecraft are complementing a census of longer pe- riod objects being compiled from radial velocity and mi- crolensing campaigns (Mayor & Queloz 1995; Marcy et al. 1997; Bond et al. 2004). Despite these successes, rela- tively little is known about the populations of extrasolar minor bodies within these systems (e.g., planetesimals, asteroids and comets). While planet formation theories generally predict that such minor bodies are a ubiquitous 1 Department of Physics, and Kavli Institute for Astrophysics and Space Research, Massachusetts Institute of Technology, Cambridge, MA 02139, USA, [email protected] 2 Harvard-Smithsonian Center for Astrophysics, Garden Street, Cambridge, MA 02138 USA; [email protected] 60 avander- 3 Department of Astronomy, The University of Texas at Austin, 2515 Speedway, Stop C1400, Austin, TX 78712 4 NASA Sagan Fellow 5 12812 SE 69th Place Bellevue, WA 98006; tomja- [email protected], USA 6 7507 52nd Place NE Marysville, WA 98270; [email protected], USA 7 NASA Ames Research Center, Moffett Field, CA 94035, USA 8 Department of Astronomy, University of Texas, Austin, 78712-1205, USA 9 Department of Physics, Northeastern University, 100 Forsyth St, Boston, MA 02115 10 Citizen Scientist, [email protected] , USA byproduct of protoplanetary disk evolution and should be found on scales loosely analogous to those observed in the solar system, their low masses and small radii present extreme challenges to detection via solid-body transits and radial velocity techniques. Even in the most favourable cases, the detection of extrasolar minor bodies in either radial velocity variations or solid-body transits would require sensitivity orders of magnitude higher than the current state of the art. Presently, the smallest solid-body which has been de- tected in transit is Kepler-37b, a 0.27 R⊕ object on a 13- day period around a solar-like main sequence star (Bar- clay et al. 2013). The smallest object detected via its host star's reflex motion is the lunar-mass PSR B1257+12 d, detected via exquisitely sensitive pulsar-timing observa- tions (Wolszczan 1994). In some cases, it is possible to detect even smaller sized objects in white-light tran- sit observations because these objects are surrounded by optically thick material (e.g., dust) which increases the transit depths. Examples of such smaller objects include the so-called "disintegrating planets" (KIC 12557548b, aka 'KIC 1255b' or KOI 3794, Rappaport et al. 2012; KOI 2700b, Rappaport et al. 2014a; K2-22b, Sanchis- Ojeda et al. 2015), which have been detected in transit. It is believed that these are rocky bodies of lunar size or smaller (Perez-Becker & Chiang 2013) in short-period orbits (9-22 hrs) that produce transits only by virtue of the dusty effluents that they emit (van Lieshout & Rappaport 2017). A perhaps similar scenario has been detected for the white dwarf WD 1145+017 (Vanderburg et al. 2015). This is an isolated white dwarf that is being orbited by debris with periods of ∼4.5-5 hours, which ap- parently emit dusty effluents that can block up to 60% of the star's light (Gansicke et al. 2016; Rappaport et 2 Rappaport et al. al. 2016; Gary et al. 2017). It is currently unknown how small the involved bodies are, but estimates range from the mass of Ceres on down. There are other avenues to studying extrasolar minor planets that lie outside of the traditional exoplanet de- tection methods. Radio observations, in particular de- tections of circumstellar CO emission around stars such as HD 181327 (Marino et al. 2016), Eta Corvi (Marino et al. 2017), and Fomalhaut (Matr`a et al. 2017) have been attributed to the presence of substantial popula- tions of minor bodies at large orbital separations. An- other sensitive method for detecting and understanding populations of extrasolar minor planets is through time- series spectroscopy, rather than time-series photometry (e.g., as performed by Kepler). Growing evidence for the existence of large populations of extrasolar minor bodies orbiting other stars has come from the detection of anomalous absorption features in the spectra of at least 16 A-type stars where Falling Evaporating Bodies ('FEBs') are proposed to randomly cross the observing line of sight (e.g., Ferlet et al. 1987; Beust et al. 1990; Welsh & Montgomery 2015). FEBs can be classified as planetesimals or exocomets which have been perturbed into eccentric orbits resulting in a star-grazing periapsis and significant sublimation of volatiles within (cid:46) 0.5 AU of the stellar photosphere. This phenomenon results in variable, often red-shifted, absorption-line features typi- cally superposed on the CaII H & K photospheric lines. Such features have been demonstrated to manifest them- selves on short time scales (hours to days) for a num- ber of known FEB systems associated with young A-B type stars including Beta Pictoris (Smith & Terrile 1984), 49 Ceti (Zuckerman & Song 2012), HD 42111 (Welsh & Montgomery 2013), HD 172555 (Kiefer et al. 2014a) and Phi Leonis (Eiroa et al. 2016). Beta Pictoris itself rep- resents an important benchmark system as it is young (∼23 Myr), hosts a massive directly imaged exoplanet, and has also been the target of an extensive 8-year long spectroscopic survey. The latter has revealed that the transient absorption features are bimodal in depth and may arise from two distinct populations of exocomets (Kiefer et al. 2014b). In this work we describe the first good evidence for ex- ocomet transits of a host star in continuum light. The object in question is KIC 3542116, a young, magnitude 10, spectral type F2V star observed from 2009 to 2013 by the Kepler mission. The paper is organised as fol- lows: In Section 2 we define the search methods and the analysis tools used in the identification of exocomet host- star candidates. In Section 3 we present what appear to be comet transits of KIC 3542116 with their distinctly asymmetric profiles. Section 4 discusses our data vali- dation methods and quality assessment of the archival photometry. Section 5 describes the supplemental infor- mation we have gathered regarding the host star KIC 3542116 including its photometric properties, UKIRT image, spectrum, Keck high-resolution imaging, and a study of the 100-year photometric history based on the Harvard Plate Stack collection. Section 6 describes the model fits for the six significant transit events. In Sec- tion 7 we interpret the model results under a variety of a priori assumptions and list several possible scenarios to explain the observations. In Section 8 we present ev- idence for a single similar comet-shaped transit in KIC 11084727, a near twin to KIC 3542116. A number of discussion items are presented in Section 9. Section 10 offers a summary of our work and draws some overall conclusions. 2. VISUAL SEARCH OF THE KEPLER DATA SET Much of the Kepler data base has been thoroughly and exhaustively searched for periodically occurring ex- oplanet transits (e.g., Borucki et al. 2010; Batalha et al. 2013) and binary eclipses (Prsa et al. 2011; Slawson et al. 2011; Matijevic et al. 2012) yielding some 3000 viable planet candidates and a comparable number of eclips- ing binaries. The types of algorithms employed include the Kepler team's Transiting Planet Search (TPS, Jenk- ins 2002), Box Least Squares technique (BLS, Kov´acs et al. 2002) and Fast Fourier Transforms ('FFTs'; see, e.g., Sanchis-Ojeda et al. 2014). There have also been a num- ber of searches carried out for single (i.e., 'orphaned') exoplanet transits (Wang et al. 2015; Foreman-Mackey et al. 2016; Uehara et al. 2016, Schmitt et al. 2017). Ad- ditionally, searches for astrophysical transit signals that are only quasi-periodic have been carried out in an au- tomated way (see, e.g., Carter & Agol 2013). In an effort to further explore the larger Kepler data set for isolated transits or aperiodic phenomena, one of us (TJ) undertook a detailed visual search of the complete Q1-Q17 Kepler lightcurve archive spanning 201,250 tar- get stars for Data Release 25 (Thompson et al. 2016a) produced by the final Kepler Science Operations Cen- ter 9.3 pipeline (Jenkins 2017). The survey was con- ducted using the LcTools11 software system (Kipping et al. 2015), a publicly available Windows-based set of applications designed for processing lightcurves in a fast and efficient manner. Two primary components from the system were utilised; LcGenerator for building lightcurve files in bulk and LcViewer for visually in- specting plots of the lightcurve files for signals of inter- est. In this survey, lightcurve files were built by LcGen- erator in batches of 10,000 files. To build a lightcurve file for a given star, LcGenerator (1) downloaded all available long-cadence time series files from MAST12 for Quarters 1-1713, (2) extracted the time stamps and PDCSAP flux values from the files excluding data points having a non-zero SAP−QUALITY value, (3) nor- malised the flux values to a mean value of 1.0, and (4) wrote the combined results to a text file. To expedite the survey, LcViewer was run concur- rently with LcGenerator. As one batch was building, another batch was being inspected. To inspect a batch in LcViewer, a 'Work Group' set of text files was first produced. Once a Work Group was established, each file from that set could be opened and displayed sequen- tially at the click of a button-the process was nearly instantaneous. If the host star had associated Kepler Objects of Interest (KOIs), as obtained from the NASA Exoplanet Archive14 (Akeson et al. 2013), the KOI sig- nals were automatically displayed in colour (highlighted) 11 https://sites.google.com/a/lctools.net/lctools/ 12 https://archive.stsci.edu/kepler/ 13 The Kepler data are downlinked and processed in approxi- mately 90-day "quarters". 14 https://exoplanetarchive.ipac.caltech.edu/ Transiting Exocomets 3 Fig. 1.- Kepler PDCSAP photometric lightcurve of KIC 3542116 spanning four years of the main Kepler mission. The data train has been harmonically filtered in preparation for searching for planets in the Transiting Planet Search module of the Kepler science pipeline (Jenkins et al. 2010). The three deepest transits are marked as "D992", "D1176", and "D1268", and have depths ranging from 0.12% to 0.15%. The three more shallow and narrow transits ("D140", "D742", and "D793") are partially obscured by some residual ∼20-day spot modulations that leak into the photometric aperture from KIC 3542117, and the ∼1-day spot rotation period of KIC 3542116. on the viewing screen with annotations when hovering the cursor over a transit signal for easy identification. Signals were highlighted for confirmed planets, planet candidates, and false positives. Any remaining signals were then examined in more detail as possible signals of interest. LcViewer allowed for rapid scrolling through each lightcurve presentation of the entire 17 Kepler quar- ters with excellent temporal and flux resolution. 3. DISCOVERY OF EXOCOMET TRANSITS IN KIC 3542116 After KIC 3542116 was initially identified as an object of interest, we performed a more thorough inspection of the four-year Kepler lightcurve. KIC 3542116 was ob- served during the entire prime Kepler mission with high photometric precision of about 35 ppm per 30 minute exposure thanks to its bright Kepler-band magnitude of Kp = 10. The full Kepler lightcurve is shown in Figure 1. The visual survey of the 201,250 unique Kepler tar- get lightcurves was conducted over the course of 5 months, beginning in January 2017. Approximately 2000 lightcurves were studied during each active day of the survey, requiring some 5 hours of visual study, allotting about 10 seconds to each target star that showed noth- ing interesting or unusual. Much more time was spent on the small percentage of stars that revealed one or more potentially interesting features. If an interesting photo- metric feature was noticed, that object was flagged for further study, vetting, and discussion. During the course of this comprehensive review, KIC 3542116 was identified as a target of interest due to three anomalous, asymmetric transit-like features occurring in Quarters 10, 12 and 13. These transits were not difficult to spot, with (cid:38) 0.1% depths and ∼1-day durations. The Kepler lightcurve of KIC 3542116 and the transits are de- scribed in detail in Section 3. We also detected a similar- looking single asymmetric transit in the lightcurve of an- other target: KIC 11084727. We discuss this object in more detail in Section 8. In addition to these two stars showing asymmetric transits in their lightcurve, we also identified other ob- jects of interest such as single exoplanet transits and mu- tual lensing events in binaries. These objects will be dis- cussed in detail in a future paper. Initially our interest was drawn to the three transit events described in Section 2. These events are high- signal-to-noise, with depths about 20 times greater than the typical scatter of the Kepler data points. Upon closer inspection, we identified three additional shallower tran- sits with depths about half that of the three deep tran- sits we initially identified. These shallower transits have similar asymmetric profiles to the deep ones, but shorter durations. We label these six dips in the full lightcurve plot shown in Figure 1. We label the dips by the date on which they took place (in the Kepler Julian Date refer- ence system BJD - 2454833). The deep dips are labeled D992, D1176, and D1268 and the shallow dips are labeled D140, D742, and D793. In order to assess the harmonic content in the flux times series of KIC 3542116, we take the Fourier trans- form of the PDCSAP time series (similar to that shown in Fig. 1). The FFT in Fig. 2 shows two close periods at 1.092 d and 1.160 d, which are likely due to the same underlying rotation period of KIC 3542116; the two pe- riods are most probably due to differential rotation of spots at different stellar latitudes (see, e.g., Reinhold et al. 2013). The (cid:39) 1 day signals have a semi-amplitude of about 175 ppm. The array of periodicities near 23 days is due to photometric leakage of a spot rotation period in 4 Rappaport et al. TABLE 1 Parameter KIC 3542116 KIC 11084727 RA (J2000) Dec (J2000) Ka p Bb ga V b ra ia za J c H c Kc W1d W2d W3d W4d T e eff (K) log gb (cgs) M f (M(cid:12)) Rf (R(cid:12)) [m/H]e RVe (km s−1) Distanceg (pc) Distanceh (pc) v sin ie (km s−1) α (mas yr−1) µh,b (mas yr−1) µh,b δ 19:22:52.94 38:41:41.5 19:28:41.19 48:41:15.1 9.98 10.49 10.38 10.03 9.99 9.53 ... 9.25 9.10 9.07 9.06 9.06 8.97 8.30 6918 ± 122 4.22 ± 0.12 1.47 ± 0.10 1.56 ± 0.15 0.04 ± 0.11 −21.1 ± 0.7 260+30−15 235 − 335 57.3 ± 0.3 +7.6 ± 1.1 −3.1 ± 1.1 9.99 10.45 10.13 10.04 9.94 9.95 10.00 9.23 9.07 9.06 8.98 9.00 9.03 8.49 6790 ± 120 4.18 ± 0.19 1.45 ± 0.12 1.55 ± 0.15 −0.06 ± 0.11 +1.5 ± 0.5 250+30−15 225 − 255 32 ± 0.9 +2.5 −22.9 Note. - (a) MAST; http://archive. (b) stsci.edu/k2/data_search/search.php. VizieR http://vizier.u-strasbg.fr/; UCAC4 (Zacharias et al. 2013). (c) 2MASS catalog (Skrut- skie et al. 2006). (d) WISE catalog (Wright et al. 2010; Cutri et al. 2013). (e) TRES spectrum; see Sect. 5.2. (f) Yonsei–Yale tracks (Yi et al. 2001). (g) Based on photometric parallax only. (h) The Gaia Mission; Prusti et al. (2016). KIC 3542117 (see also McQuillan et al. 2014), a neigh- bouring star some 10(cid:48)(cid:48) to the north. The 22-23 day signal has a similar semi-amplitude of about 150 ppm. In order to obtain a clearer view of the six transit events, we attempted to separate the transits from the two rotational signals present in the lightcurve. We were easily able to filter the 22 day period signal by fitting a basis spline to the lightcurve while iteratively exclud- ing outliers and dividing away the best-fit spline. For a more detailed description and illustration of this pro- cess, see Figure 3 from Vanderburg & Johnson (2014). It proved more difficult to filter the second rotational sig- nal from the lightcurve because it has a period of about a day, which is similar to the duration of the larger tran- sits we detect around KIC 3542116. This coincidence of timescales makes it particularly tricky to filter or re- move the stellar variability while preserving the tran- sits and not modifying their shapes. We attempted to separate the stellar variability from the transits using both Fourier filtering methods and fitting and remov- ing splines to the data and filtering, but found the re- sults unsatisfactory. We achieved better results filtering the data using Gaussian Process (GP) regression (Ras- mussen & Williams 2006). In brief, Gaussian process regression involves modelling the covariance properties of a dataset. The learned covariance properties can then be used to predict (either interpolating or extrapolating) Fig. 2.- Fast Fourier transform of the KIC 3542116 PDCSAP photometric time series. The peaks near 1 day periods, and their harmonics, are due to starspot rotation in KIC 3542116, while the messy peak structure near 23-day periods is leakage from the corresponding spot rotation period in the neighbouring star KIC 3542117 (see also McQuillan et al. 2014). how a dataset might behave in the absence of data. We took a snippet of the lightcurve around each tran- sit with a duration of about 15 to 20 days and removed both 3 − σ outliers from the lightcurve and data points taken during and around transit. We trained a Gaus- sian process with a quasi-periodic kernel function (see Equation 4 from Haywood et al. 2014) on the lightcurve, optimising the parameters describing the kernel function to best match the lightcurve's covariance properties. We then used our optimised kernel function to predict the behaviour of the stellar activity during the transits, and divided the Kepler lightcurve by the GP prediction to obtain a filtered lightcurve.15 We show the flattened Kepler lightcurve of KIC 3542116 around the three deep transits in Figure 3. All three transit profiles have remarkably similar widths, shapes, and depths. In particular, all the transits have steeper ingresses with positive curvature, followed by longer egresses with negative curvature. The transits are typically 0.12-0.15% deep and last for about a day. The three shallower transits we have identified are shown in Fig. 4. Although these dips have lower signal- to-noise, they all appear to have shapes consistent with the asymmetry that is more clearly evident in the deeper, higher signal-to-noise events. We have ignored all dip- like features whose depth was less than ∼450 ppm be- cause of the possibility of having substantial distortions from the spot modulations. We tentatively interpret the transits shown in Figs. 3 and 4 as being due to the passage of comet tails across the disk of the host star, KIC 3542116, as viewed from the direction of the Earth. In this work we henceforth refer to these as 'comet transits' and endeavour to demonstrate that they are indeed consistent with the hypothesis of 15 We note that in many cases, for example when dealing with stellar variability in the presence of periodic transits, it is preferable to fit a model to the signal along with a Gaussian process to ab- sorb the stellar variability (see, for example, Grunblatt et al. 2016, 2017). In our case, however, since we do not a priori know which models appropriately describe the transits around KIC 3542116, it is best to attempt to separate the stellar variability from the tran- sits without making assumptions about the shape of the transits. Transiting Exocomets 5 Fig. 3.- Kepler SAP photometry covering 3 days around each of the three larger comet transits. The data have been cleaned via a Gaussian processes algorithm so as to remove most of the 20-day and 1-day spot modulations, as well as other red noise (see text). The red curves are model fits which will be discussed in Section 5. transiting comet dust tails. being read out by the fourth CCD channel (van Cleve & Caldwell 2016). 4. ASSESSMENT AND CHECKS ON THE TRANSIT DATA We performed a set of validation checks on these transit-like events to establish their astrophysical nature and their likely source. These tests included assessing the difference images, analyzing potential video crosstalk (van Cleve & Caldwell 2016), and inspecting the data quality flags associated with these events. To determine the location of the source of the tran- sit signatures, we inspected the pixels downlinked with KIC 3542116 for the quarters containing the three deep events, namely quarters 10, 12, and 13. Since this star is saturated and 'bleeding' due to its bright Kepler band magnitude Kp = 9.9816, the standard difference image centroiding approach as per Bryson et al. (2013) is prob- lematic: small changes in flux can affect the nature of the bleed of the saturated charge and induce light centroid shifts, especially along columns. Indeed, a shift in the flux weighted centroids in the column direction does oc- cur during the Q12 transit, but the direction of the shift is away from KIC 3542116 and toward KIC 3542117, the dim Kp (cid:39) 15 M-dwarf discussed in Section 3 located ∼9.8(cid:48)(cid:48) away from KIC 3542116. This shift is incompati- ble with the source being KIC 3542117 as the direction is consistent with KIC 3542116 being the source. Fig- ure 5 shows the direct images of KIC 3542116 and the mean difference image between out-of-transit data and in-transit data, along with the locations of KIC 3542116 and KIC 3542117. Inspection of the pixel time series over the data segments containing the transits reveals that the transit signatures are occurring in the pixels in the core of KIC 3542116 and at the ends of the columns where saturation and 'bleed' are happening. While the loca- tion of the source of the dips cannot be determined with great accuracy due to the saturation and bleeding, the fact that the transit signatures are not apparent in the saturated pixels but are visible in the pixels just above and below the saturated pixels is strong evidence that the source of the transits is in fact co-located on the sky with KIC 3542116. As a further check on the astrophysical nature of these events, we also checked against video crosstalk. The Ke- pler CCD readout electronics do "talk" to one another so that dim images (and sometimes negative images) of stars read out on the adjacent three CCD readout chan- nels are electronically superimposed on the image data 16 Stars observed by Kepler saturate at a magnitude of ∼11.5. We looked for stars located on the Kepler detectors which might be the source of any video cross-talk sig- nals by inspecting the full frame images (FFIs) for the quarters during which the three most prominent transits occurred. There is a fairly bright, possibly saturated star near the edge of the optimal aperture on output 3 on the CCD on which 3542116 is imaged (it's on output 1 in all cases), but the crosstalk coefficient is even smaller than for the other two outputs, −0.00001, so that a 50% deep eclipse on this other star would be attenuated to a value of 5 ppm when its video ghost image is added to the di- rect image of KIC 3542116, assuming they are the same brightness. Furthermore, since the coefficient is negative, there would need to be a brightening event on the star on output 3 to cause a transit-like dip on output 1. Fortunately, the largest crosstalk coefficient to the CCD output that 3542116 finds itself on is +0.00029, so given that the signal we are looking at is ∼0.1%, a con- taminating star would need to be at least 10× brighter than 3542116 to cause a problem. If there were, it would be highly saturated and bleeding, which would make it difficult to square with the pixel-level analysis indicat- ing that the source is associated with the pixels under 3542116, as the extent of the bleeding would be signifi- cantly larger than for KIC 3542116. We also inspected the quality flags associated with the flux and pixel time series for KIC 3542116 and find that the situation is nominal with flags for occasional events such as cosmic rays and reaction wheel desaturations, but no flags for rolling band noise during the transit events.17 Finally, we considered 'Sudden Pixel Sensitivity Dropouts' (SPSDs) in the data, which are due to ra- diation damage from cosmic ray hits on the CCD, as a possible explanation for the dips in flux that we observe. However, the shape and behaviour of such dips do not resemble what we see (Thompson et al. 2016b). In par- ticular, the SPSD events have drops that are essentially instantaneous, and therefore are much shorter than the ∼20 and 8 long-cadence points on the ingresses that we see in the deeper and more shallow dips, respectively. Moreover, the location of the SPSDs on the CCD chip would not plausibly align with the source location and its bleed tracks for each and every one of the dips. Thus, we also discarded this idea as well. 17 See Thompson et al. (2016b) for more information about anomalies flagged in the Kepler pixel and flux time series. 6 Rappaport et al. Fig. 4.- Kepler SAP photometry covering 3 days around each of the three smaller comet transits. Other specifications are the same as for Fig. 3. Note that the vertical (flux) scale has been expanded by a factor of 2 compared to that of Fig. 3. Parameter (R∗/day) t 1. Depth (ppm) 2a. v(a) 2b. vt (km s−1) 3. λ(b) (R∗) 4. b(c) (R∗) 5. t(d) 0 Dip 140 491 ± 38 7.76 ± 0.31 89.8 ± 3.6 0.44 ± 0.04 0.66 ± 0.05 139.98 ± 0.02 Dip 742 524 ± 58 6.55 ± 0.73 75.8 ± 8.5 0.53 ± 0.09 0.47 ± 0.18 742.45 ± 0.02 TABLE 2 Dip 793 679 ± 125 7.42 ± 0.42 85.9 ± 4.9 0.85 ± 0.16 0.63 ± 0.14 792.78 ± 0.02 Dip 992 1200 ± 100 3.04 ± 0.16 35.2 ± 1.8 0.59 ± 0.10 0.27 ± 0.13 991.95 ± 0.02 Dip 1176 1500 ± 130 4.34 ± 0.39 50.2 ± 4.5 0.76 ± 0.11 0.44 ± 0.17 1175.62 ± 0.02 Dip 1268 1900 ± 150 3.70 ± 0.20 42.8 ± 2.3 0.72 ± 0.08 0.27 ± 0.14 1268.10 ± 0.02 Note. - a. Transverse comet speed during the transit; b. Exponential tail length from Eqn. (1); c. Impact parameter; d. Time when the comet passes the center of the stellar disk. Fig. 5.- Direct images and difference images for KIC 3542116 during each of the three major transit features for each quarter in which they occur. Top panels show the mean calibrated pixel values in the aperture masks returned by Kepler for Q10, Q12 and Q13, from left to right. Bottom panels show the mean difference between the calibrated pixels in transit-feature-wide segments on either side of each transit, and the pixel values during each transit feature. The source of the transit features will exhibit positive values in these difference images. The colour-bars indicate the pixel fluxes in units of e−/s. Note that the pixels exhibiting the strongest positive deviations in the bottom panels occur primarily at the ends of the saturated and bleeding columns and are approximately clustered about the location of KIC 3542116, which is marked by a red star. KIC 3542117's location is marked by a red circle in each panel. These difference images indicate that the source of the transit features is co-located with KIC3542116 to within the resolution of the Kepler data. Transiting Exocomets 7 stellar template spectra from a library of synthetic spec- tra generated from Kurucz (1992) model atmospheres. These cross-correlations yielded an absolute radial veloc- ity of −21.1 km s−1. Cross correlating the two spectra against one another and averaging the results over many different echelle orders yielded a shift of only 400 m s−1 between the two spectra. This is consistent with the photon-limited uncertainties for an F star with a rota- tional broadening of 57 km s−1 (at this SNR the preci- sion would be at least an order of magnitude better for a slowly rotating solar-type star). We estimated stellar parameters using the Stellar Parameter Classification code ('SPC', Buchhave 2012). SPC cross correlates a library of synthetic template spec- tra with varying temperature, metallicity, surface grav- ity, and line broadening against the observed spectrum and interpolates the parameters from the best-matched template peaks to estimate the actual stellar parameters. SPC was designed to measure stellar parameters of slowly rotating stars close in effective temperature to the Sun, and has been extensively tested and used for stars cooler than the Sun. For rapidly rotating stars hotter than the Sun (such as KIC 3542216), SPC has not been tested as fully and may have systematic errors, especially in the surface gravity and metallicity. An SPC analysis of the TRES spectra of KIC 3542116 yields an effective temperature of 6900 ± 120 K and a projected rotational velocity of 57 km s−1. This makes it fairly unusual for stars observed by Kepler, which mostly observed sun-like dwarfs and smaller stars, which are more amenable for searches for small Earth-like planets. The properties of KIC 3542116 measured with, or in- ferred from, the TRES spectra are summarised in Table 1. 5.3. High-Resolution Imaging We observed KIC 3542116 with the Near Infrared Camera 2 (NIRC-2) instrument behind the Natural Guide Star (NGS) adaptive optics system on the Keck II telescope on the night of 28 June 2017. We obtained standard adaptive optics ('AO') images, both with and without a coronagraph in place. We also recorded inter- ferograms produced by placing a sparsely sampled nine- hole non-redundant aperture mask ('NRM') in the pupil plane to resample the full telescope aperture into an in- terferometric array (Tuthill et al. 2006; 2010). This pro- cess makes it possible to detect companions closer to the target star than the traditional diffraction limit. We ob- tained four 20-s exposures in imaging mode in K(cid:48) band, as well as four exposures of the same duration with the coronagraph. Six additional 20-s exposures were taken with the NRM in place. The imaging observations were reduced following Kraus et al. (2016), and detections and detection limits were assessed using the methods they de- scribed. The summed set of the standard AO images is shown in the top panel of Fig. 7; it covers only the central 1(cid:48)(cid:48) × 1(cid:48)(cid:48) region of the field. The bottom panel in Fig. 7 shows the resultant image acquired with the coronagraphic disk in place, and it covers a wider 4(cid:48)(cid:48) × 4(cid:48)(cid:48) portion of the field. The images are colour coded so that roughly each con- trast change of 1 magnitude is represented by a change of one colour. From the ordinary AO image we can estimate that there is no neighbouring star of comparable K(cid:48) mag- Fig. 6.- UKIRT J-band image of KIC 3542116. The colours represent a logarithmic scaling. Note that the centre of the KIC 3542116 image is saturated. The grid lines are spaced at 4(cid:48)(cid:48) × 4(cid:48)(cid:48) to match the Kepler pixels (North is up and East is to the left). The neighbouring star, KIC 3542117, is 10(cid:48)(cid:48) to the north and has Kp = 14.9 (J = 12.4), compared to Kp = 9.98 (J = 9.3) for KIC 3542116. The faintest blue-coloured stellar images have J (cid:39) 16.0. We take all these evaluations as strong evidence that the dips we see are of astrophysical origin and that KIC 3542116 is indeed the source of them. However, we can- not categorically rule out the possibility that the dips are caused by some unknown peculiar type of stellar variabil- ity in KIC 3542116 itself. In spite of this caveat, we proceed under the assumption that the dips in flux are indeed due to the passage of objects in Keplerian orbit that are trailing tails of dusty effluents. 5. GROUND-BASED STUDIES OF KIC 3542116 The photometric properties of KIC 3542116 are sum- marised in Table 1. Fortunately, this is a relatively bright star that is amenable to follow-up ground-based study. 5.1. UKIRT Image The UKIRT image of KIC 3542116 is shown in Fig. 6. In addition to the bright target star KIC 3542116 at Kp = 9.98, the image shows a neighbouring star, KIC 3542117, with Kp = 14.9 some 10(cid:48)(cid:48) to the north. This star is the source of the 23-day modulations (see McQuil- lan et al. 2014) that leak into the flux data train of KIC 3542116, and may be a low-mass bound companion to this star (see Sect. 5.3). 5.2. TRES Classification Spectrum We observed KIC 3542116 with the Tillinghast Re- flector Echelle Spectrograph (TRES) on the 1.5m tele- scope at Fred L. Whipple Observatory (FLWO) on Mt. Hopkins, AZ. We obtained two high-resolution (λ/∆λ = 44,000) optical spectra of KIC 3542116 – the first on 9 June 2017 and the second on 14 June 2017. Expo- sure times of 300 s and 200 s yielded signal-to-noise ra- tios of 50 and 43 per resolution element at 520 nm. We cross-correlated the two spectra with a suite of synthetic 8 Rappaport et al. Fig. 8.- Contrast limits (in magnitudes) for possible neighbor- ing stars to the target star, KIC 3542116, as a function of radial distance. These were obtained with high resolution imaging using the Keck telescope (see Sect. 5.3). The red circles, green triangles, and blue squares are for the non-redundant mask technique, natu- ral guide star adaptive optics, and AO plus a coronagraph (see text for details). The yellow curve is a smooth fit to the best contrast constraints at any given radial distance. The working region within the field goes out to 4.5(cid:48)(cid:48). Neighbouring field stars farther from the target than ∼0.4(cid:48)(cid:48) are marked as "field interlopers excluded" be- cause they would have insufficient flux to produce a 0.15% dip in the optical flux of KIC 3542116 unless they are extremely blue. Physical binary companions with separations (cid:38) 0.04(cid:48)(cid:48) are labeled "binary companions excluded" because they also would have in- sufficient flux to produce a 0.15% dip in the optical flux of KIC 3542116. of < 1.5, 4.2, 5.1, and 4.8 K(cid:48) magnitudes can be rejected at the 99% confidence limit at distances of 0.01-0.02(cid:48)(cid:48), 0.02-0.04(cid:48)(cid:48), 0.04-0.08(cid:48)(cid:48), 0.08-0.16(cid:48)(cid:48), respectively. We summarise all of the constraints from the three different imaging modes in Fig. 8. We now consider the constraints we can place on pos- sible neighbouring stars in each of two different cate- gories: (i) random interloping field stars, and (ii) phys- ically bound companions. For unbound field stars, we see from Fig. 8 that for angular separations greater than (cid:39) 0.4(cid:48)(cid:48) there are no stars within 7 K(cid:48) magnitudes of the target star. The significance of this latter limit is that stars fainter than Kp = 17th magnitude18 could not pro- duce a dip as apparently deep as 0.0015 in the flux of KIC 3542116. It is possible, though most unlikely, that there could be a star accidentally aligned with the target to within (cid:46) 0.4(cid:48)(cid:48) that is the source of the dips. We estimate the probability of a nearby interloper star with the req- uisite magnitude of Kp (cid:46) 17 randomly lying within 0.4(cid:48)(cid:48) of KIC 3542116 as (cid:46) 0.1% (see, e.g., Fig. 9 of Rappaport et al. 2014b). Alternatively, the target star could have a physical bi- nary companion that is the host of the dips. In this case, the companion star would be at the same distance and coeval with the target star. Since the target star is not significantly evolved (see Table 1) any fainter com- 18 For neighbouring field stars cooler than Teff (cid:39) 7000 K the contrast limits in Kp are even more stringent than in K(cid:48). How- ever, we estimate that the contrast constraints obtained in K(cid:48) are still good to within (cid:39) 1 magnitude in Kp for Teff of the hypo- thetical interloper star up to (cid:39) 15, 000 K. For even higher Teff the constraints weaken further only very slowly. Fig. 7.- Images of KIC 3542116 from Keck-II/NIRC2 and nat- ural guide star adaptive optics, obtained without (top) and with (bottom) the 0.6(cid:48)(cid:48)-diameter coronagraphic spot. In both panels the pseudocolour is a logarithmic scaling such that each colour dif- ference represents a ∼1 magnitude change in brightness. The grid scale in the top (AO) image is 0.1(cid:48)(cid:48) × 0.1(cid:48)(cid:48), while the grid in the bottom panel is 1(cid:48)(cid:48) × 1(cid:48)(cid:48). The attenuation from the coronagraphic spot (∆K = 7.25 ± 0.10 mag; Kraus et al. 2016) allows for un- saturated observations with longer individual exposures and more Fowler samples, which reduces read-noise and allows the detection of fainter sources at wide separations. We see no neighbouring sources out to an azimuthally complete projected separation of ρ = 4.5(cid:48)(cid:48), with partial azimuthal coverage (from the corners of the detector) out to ρ = 7(cid:48)(cid:48). The wide-separation contrast limits are ∆K(cid:48) = 8.6 mag and ∆K(cid:48) = 10.8 mag, respectively. nitude within 0.05(cid:48)(cid:48) of the target star, and no star that is at most 4 magnitudes fainter within 0.15(cid:48)(cid:48). With the AO-plus-coronagraph image the sensitivities are compa- rable out to about 0.8(cid:48)(cid:48), beyond which the image goes 2 magnitudes deeper than the plain AO image. We can further constrain the magnitudes of any stars within ∼0.2(cid:48)(cid:48) of the target, using the 9-hole mask interfer- ogram. The analysis of those data shows that contrasts Transiting Exocomets 9 panion star would necessarily be redder in colour and lower in mass than the target star. For redder stars, the K(cid:48) band is obviously more sensitive than the Kp band. From Fig. 8 we see that for any binary (projected) sepa- ration (cid:38) 0.04(cid:48)(cid:48) all binary companions with ∆K(cid:48) (cid:38) 4 are ruled out. That already suggests that any companion- star mass satisfying this requirement must be (cid:46) 0.8 M(cid:12). However, for main-sequence stars of this mass, and lower, the value of ∆Kp to be expected in the Kepler band would be 2-3 magnitudes greater. Thus, it is safe to say that for angular separations (cid:38) 0.04(cid:48)(cid:48) there are no binary companion stars that could produce the observed dips. This translates to a binary orbital projected separation of (cid:46) 10 AU. If we combine this with the constraint on the change in RVs over a 5-day interval (Sect. 5.2), this sug- gests that any viable binary companion star that could produce the dips would likely have an orbital separation of ∼0.5-10 AU. Of course very faint binary companions that are not the source of the dips are allowed. They must, how- ever, still satisfy the constraints summarised in Fig. 8. In this regard, we note that the faint star, KIC 3542117 (see Fig. 6) has the colours (taken from the Sloan Digital Sky Survey images; Ahn et al. 2012) to be a (cid:39) 0.5 M(cid:12) (see also Dressing & Charbonneau 2013) companion star since it lies relatively close to the same photometric dis- tance as KIC 3542116, and because it has only a (cid:46) 15% chance of being found within 10(cid:48)(cid:48) of the target star by chance (see, e.g., Fig. 9 of Rappaport et al. 2014b). We note, however, that the proper motion of KIC 3542117 is not consistent with comovement, though the Deacon et al. (2016) seeing-limited astrometry might be limited by the brightness of KIC 3542116. Any possible association should become clear in the upcoming Gaia Data Release 2. 5.4. Historical Plate Stacks The available photometry of KIC 3542116 from the past century, taken from the "Digital Access Sky Cen- tury at Harvard" ('DASCH'; Grindlay et al. 2009) are shown in Fig. 9. The systematic drop in flux, by ∼10% across the 'Menzel gap', is likely due to a change in the plate emulsion response. No other obvious dimmings or outbursts of the star are observed. A Fourier transform of these data show no clear periodicities in the range of 1 − 100 days over the past 100 years down to a level of (cid:38) 2%. 6. MODEL FITS TO THE TRANSITS The repeatably asymmetric shape of the transits of KIC 3542116 is suggestive of an occulter with some sort of sustained or repeatable dust outflows causing the dim- ming. Here, we show that the observed asymmetry and the repeatable shape are consistent with, and what we might expect from, a large comet transiting the star with a dusty tail. We describe a simple model for the transit of a comet and show that the six transits detected in the lightcurve of KIC 3542116 are well fit by this model. Almost all cometary dust tails will lag behind the di- rection in which the comet is moving. This is as opposed to the ion tails of comets which are driven out nearly radially by the stellar wind of the host star (see, e.g., Fig. 9.- One hundred years of photometry on KIC 3542116 from the Harvard Plate Stack collection, "Digital Access Sky Century at Harvard" ('DASCH'; Grindlay et al. 2009). The systematic drop in flux, by ∼10% across the 'Menzel gap', is likely due to a change in the plate emulsion response. Reyes-Ruiz et al. 2010). When the dust is released from the immediate and gravitational environs of the comet, it finds itself in a reduced effective gravity due to the effects of radiation pressure on the dust grains. This, in turn, results in the dust moving too fast to remain in the same orbit as the parent comet. Somewhat paradoxi- cally, the higher speed causes the dust to go into a higher orbit which, in turn, causes its mean orbital speed to de- crease, and thus results in a trailing tail (in the sense of lagging in angular position). Leading dust tails are also possible, but usually in the context of the dust overflowing the Roche lobe of the par- ent body, and with little subsequent radiation pressure (see, e.g., Sanchis-Ojeda et al. 2016). However, Roche- lobe overflow typically requires orbital periods of (cid:46) 1 day. In this work we assume that the orbital periods of the putative comets are substantially longer than the transit durations (∼1 day), and that the transverse velocity, vt, during the transit is essentially a constant. For lack of more detailed information, we further assume that the dust tail is narrow compared to the size of the host star, and that its dust extinction profile is a simple exponential function of distance from the comet (see, e.g., Brogi et al. 2012; Sanchis-Ojeda et al. 2015). We take the comet location, projected on the plane of the sky, to be {xc, yc}, and an arbitrary location on the disk of the star to be {x, y}, where {0, 0} is the center of the stellar disk, and x is the direction of motion of the comet. We model the attenuation of starlight at {x, y}, as τ = C e−(xc−x)/λ for x < xc; y − b ≤ ∆b/2 τ = 0 otherwise (1) (2) where C is a normalisation constant equal to the optical depth of the dust tail just behind the comet, λ is the exponential scale length of the tail, b is the impact pa- rameter of the transit, and ∆b is the width of the dust tail as projected on the disk of the host star19. We as- sume that ∆b (cid:28) R∗. The comet position along the x 19 ∆b is an undetermined parameter that is essentially degener- ate with the normalisation constant C 10 Rappaport et al. direction is given as a function of time, t, by: xc = vt(t − t0) (3) where t0 is the time when the comet crosses the centre of the stellar disk. We also adopt a quadratic limb dark- ening law for the host star, with coefficients appropriate to a mid-F star (Claret & Bloemen 2011). We utilize a Markov Chain Monte Carlo ('MCMC') code (Ford 2005; Madhusudhan & Winn 2009; Rappa- port et al. 2017) to fit a comet-tail transit model to each of the six transits that we observe in KIC 3542116. There are six free parameters to be fit: t0, λ, C, vt, b, and DC, where the final term is the DC background flux level away from the transit. For each choice of parameters, we gen- erate a model lightcurve by integrating over the dust tail where it overlaps the stellar disk, and repeating this in increments of 6 minutes as the comet crosses the stellar disk. The model lightcurve is then convolved with the 30 minute integration time of the Kepler long-cadence sampling. Each model is then evaluated via the χ2 value of the fit to the data, and the code then decides, via the Metropolis-Hastings jump condition, whether to accept the new set of parameters or to try again. Each MCMC chain has 3×105 links, and we run a half- dozen chains. The results for the fitted parameters are given in Table 2 for each of the parameters and for each of the six transits. We show illustrative MCMC correlations among the three physically interesting parameters, λ, vt, and b, graphically for the fit to transit D1268 in Fig. 10. 7. INTERPRETING THE TRANSITS Now that we have shown that the shapes of the transits of KIC 3542116 are consistent with the shape caused by a dusty tailed comet, in this section, we will show that the parameters we derive from our fits correspond to plausible physical conditions. 7.1. Inferred Comet Orbital Velocities The first step in trying to understand what orbits the putative comets orbiting KIC 3542116 would be on, is to attempt to explain the observed transverse speeds of the bodies. This involves speeds of 35 − 50 km s−1 for the deeper transits and 75 − 90 km s−1 for the more narrow and shallow transits (see Table 2). To gain some insight into this problem, we carried out the following exercise. We chose random orbital periods from a distribution that is uniform in log Porb, and with a uniform distribution of eccentricities from e = 0 to 1. With respect to a fixed viewing direction, we also chose longitudes of periastron, ω, at random from 0 to 2π (ω is here defined as the angle from entering the plane of the sky to periastron). For simplicity, all orbits are taken to be in the same plane with an inclination angle of 90◦ with respect to the observer. For a Keplerian orbit there is a relatively simple rela- tion among the transverse speed during transit, vt, the orbital period, and the orbital eccentricity vt = (2πGM )1/3P −1/3 orb [1 − e sin ω](1 − e2)−1/2 (4) Also, at the time of transit, the separation between the host star and comet can be written analytically as d = a(1 − e2)[1 − e sin ω]−1 = (2π)−1/3(GM )2/3P 1/3 orb (cid:112) 1 − e2 v−1 t (5) After choosing 107 such random orbits we record the mean transverse speed of each body as it transits the host star. The results are shown in Fig. 11. To construct the left panel we simply add the value 1 to each pixel in the Porb− vt plane where a system appears. By contrast, in the right-hand panel, the weight given to each system is taken to be proportional to a/d, where a is the comet semi-major axis and d is the separation of comet and host star during the transit. The 1/d factor is supposed to represent the probability of a transit if the comets were not all in the same plane, but rather in a range of randomly tilted planes. The extra factor of a is simply to (i) render the weighting factor dimensionless, and (ii) restore some weight to the longer-period systems, so that the shorter period systems don't dominate the diagram. As discussed below in Sect. 7.3, comet dust tails are very unlikely to survive sublimation at distances from the host star of (cid:46) 0.1 AU (see Eqn. 5 and Fig. 12). There- fore, any systems with such close approaches during the transit are eliminated from the diagram. Equations (4) and (5) can be combined to yield analytic relations for the upper limit to vt and lower limit to Porb in Fig. 11 for a given minimum allowed star-comet separation dur- km s−1 and Porb > 3.5 d3/2 ing transit: vt < 160 d 0.1 days, where the subscript on d indicates that it is nor- malised to 0.1 AU. These boundaries are clearly evident in Fig. 11. −1/2 0.1 The white track in the Fig. 11 diagram is just the locus of points given by Porb ∝ v−3 t (6) that would hold for a simple circular orbit. Points to the right of the white track are eccentric orbits with the periastron passage on the side of the orbit nearest the ob- server while systems to the left are eccentric orbits with periastron on the far side of the orbit. In the right-hand panel of Fig. 11 the higher likelihoods to the right of the white band arise from the inverse weighting with sepa- ration between the host and comet during the transit. The higher speeds reflect closer distances, and therefore higher transit probabilities. As we can see from Fig. 11, transverse speeds of 35 − 50 km s−1 would correspond to circular orbit periods of ∼100 − 300 days, but periods as short as ∼6 days are plausible. Note, especially from the right-hand panel in Fig. 11, that arbitrarily long orbital periods are quite possible. Similarly, the more narrow and shallow dips imply transverse orbital speeds during transit ranging between ∼75 and 90 km s−1. From Fig. 11 we see that such orbits would correspond to circular orbit periods of ∼20-35 days. However, periods as short as ∼6 days or arbitrarily long are also quite acceptable. One possibility is that all three of the deeper transits arise from a single body in a periodic orbit (or nearly so). In that case, the orbital period would be 92 days, but would have to exhibit transit timing variations ('TTVs') of ∼1/3 day. Additionally, there would be the issue of why only three transits appear out of a possible ∼16 that Transiting Exocomets 11 Fig. 10.- Illustrative examples of the correlations among the MCMC fitted parameters for the comet speed, vt, impact parameter, b, and exponential tail length, λ (see Sect 6). This particular fit was for transit D1268. The colour scaling is logarithmic. Fig. 11.- Statistical assessment of different comet orbital periods and eccentricities. The orbital periods are chosen randomly from a uniform distribution per logarithmic interval, while the eccentricities and arguments of periastron were chosen randomly between 0 and 1, and 0 and 2π, respectively. For each orbit the transverse speed across the disk of the host star was recorded. In the left panel the weighting for each system is 1.0, while in the right panel, the weighting is proportional to a/d, where a is the comet semimajor axis, and d is the star-comet separation during the transit. All systems where the star-comet separation is < 0.1 AU during the transit are eliminated due to dust sublimation that would destroy the tail (this forms the lower and right boundaries of the diagram). The colour coding is logarithmic (white is highest and purple lowest) and reflects the chosen weighting. For other details see text. potentially could have been detected during the Kepler main mission. Presumably, such an explanation requires highly and remarkably variable dust emission from the body. In this regard, we note that the dust-tail optical depths of some Solar-system comets are highly variable (see, e.g., Montalto et al. 2008; Sekanina & Chodas 2012; Knight & Schleicher 2015), and that the transit depths in the 'disintegrating' planets KIC 1255b and K2-22b are also highly and erratically variable (Rappaport et al. 2012; Sanchis-Ojeda et al. 2015). If the three deeper transits are indeed due to a single body orbiting with TTVs of up to ∼1/3% of the orbital period, then only the discrete periods of 92/n, where n = 1, 2, 3, ... are allowed. Similarly, the three more shallow and narrow tran- sits could be due to another distinct body orbiting KIC 3542116. The maximum such period consistent with these three transits is ∼51 days (the time interval be- tween D742 and D793). If these three shallower transits are indeed due to a single body in a fixed orbit, then one must explain why only three of a possible 30 transits are detected. Again, this would require highly variable dust emission. Alternatively, all 6 of the transits could each be due to a separate body in the system. In that case, one needs to explain why all three of the deeper transits are so re- 12 Rappaport et al. markably similar in depth, shape, and duration. And, to a somewhat lesser degree, the same argument applies to the three more shallow transits. We therefore ten- tatively adopt the working hypothesis that there are at least two distinct orbiting minor bodies in KIC 3542116 with cometary tails that produce transits. Finally, in a related vein, we give some estimates of the transit probabilities for the comets we believe we have detected. Since the transit probability is R/d, we can rewrite Eqn. (5) as: (cid:18) (cid:19)(cid:18) Porb (cid:19)−1/3 (cid:39) 0.0076 R d vt 40 km/s 1 yr (1 − e2)−1/2 (7) where we have normalised the transverse speeds to 40 km s−1 which is roughly the mean value for the larger transits. Thus, the transit probability is not particularly large unless either the orbital period is quite short or, more likely, the orbital eccentricity is close to unity. For example, if e = 0.99, the transit probability rises to 5%. 7.2. Inferred Dust Mass Loss Rates It is straightforward to estimate a lower limit to the instantaneous mass contained in a comet tail that would be required to produce a ∼0.1% transit depth. Since the most effective visual extinction per unit mass occurs for ∼1 µm particles, consider a dust sheet of thickness, h, that is between the observer and the host star which we take to have radius, R∗. The minimum projected area of the dust cloud particles that is required to block 0.1% of the star's light is ∆Ad (cid:39) 0.001πR2∗. The corresponding mass in such a dust sheet is at least: ∆Md (cid:38) 0.001πR2∗hρd (cid:39) 1016 g (8) where ρd is the bulk density of the dust, and where we have taken ρd (cid:39) 3 g, and h = 1 µm. Without knowing the specific properties of the dust or the comet orbit, it is difficult to know the speed of the dusty effluents with respect to the comet. However, if we assume a minimal value for β, the ratio of radiation pres- sure to gravity, of ∼0.05, the relative dust speed could be ∼0.1 times the orbital speed of the comet (see, e.g., Rap- paport et al. 2014a), or some 5 km s−1. At this rate, the dust tail at ∼2 R∗ from the comet would be replenished every ∼5 days. This, in turn, corresponds to a minimum dust mass loss rate of Md (cid:38) 2.5 × 1010 g s−1. Finally in this regard, if we assume that the comet emits dust at this rate for even half of the interval be- tween the D992 and D1268 transits (276 days), during which time there were three of a possible four transits seen, then the minimum comet mass would be Mc (cid:38) 3 × 1017 g. This is just a little bit greater than the mass of Halley's comet. 7.3. Dust Sublimation In order for a dust-rich comet tail to exist it should not be so close to the host star that the dust grains leading to most of the opacity quickly sublimate (i.e., on less than a timescale of about a day). The equilibrium temperature of the dust, Tequil, depends mainly on the stellar flux at its location, its size, s, and the imaginary part of its index of refraction, k, at the wavelengths of interest. We have computed Tequil for three different characteristic grain Fig. 12.- Dust-grain equilibrium temperatures vs. distance from the host star KIC 3542116 for three different size particles, s = 0.1, 1 and 10 µm, and three different imaginary parts of their indices of refraction, k = 1, 0.1 and 0.01, as indicated in the legend (the first of the two numbers is s and the second is k). The bottom black curve is the idealized equilibrium temperature if the parti- cles absorb and emit as blackbodies, while the upper black curve is "Trayl", as defined in the text. sizes, s = 0.1, 1, and 10 µm as functions of their distance from KIC 3542116. In lieu of discussing any particular mineral composition for the grain size, we simply adopt three illustrative values for k equal to 1, 0.1 or 0.01, that are taken to be independent of wavelength, and are fairly representative of different refractory minerals (see, e.g., Beust et al. 1998; Fig. 13 of Croll et al. 2015; van Lieshout et al. 2014; Xu et al. 2017; and references therein). The Mie scattering cross sections were computed with the Boren & Huffman (1983) code. The results for Tequil are shown in Fig. 12 for distances ranging from 10 AU in to 0.01 AU from KIC 3542116. We also show a curve for Tequil,bb of the dust particles if they emitted and absorbed as blackbodies: (cid:114) (cid:18) d (cid:19)−1/2 Tequil,bb (cid:39) Teff (cid:39) 420 R 2d AU K (9) For small scaled particle sizes (and/or long wavelengths), X ≡ 2πs/λ, where λ is the wavelength of the light that is interacting with the grain, the normalised absorption cross sections scale simply as X. In that case, it has been shown (see, e.g., Rappaport et al. 2014a) that Tequil,rayl (cid:39) Teff (cid:39) 730 K (10) (cid:18) R (cid:19)−2/5 2d (cid:18) d (cid:19)−2/5 AU Here we retain the nomenclature of the "Rayleigh" tem- perature following Xu et al. (2017). This higher Tequil simply reflects the fact that the particles are better ab- sorbers of the starlight than emitters in the IR, at the same particle size, and therefore the equilibrium temper- atures that are attained are higher. As can be seen from Fig. 12, Tequil,bb adheres closely to the calculated values of Tequil for the larger particles, while the Tequil,rayl val- ues are higher and closer to the calculated values of Tequil for the smaller particles, The bottom line of these calculations is that for many common minerals (e.g., obsidian, magnetite, SiO, fay- alite, enstatite, forsterite, corundum, and SiC), which Transiting Exocomets 13 commence rapid sublimation at Tequil (cid:38) 1200−1700 K20, we can estimate that the dust tails begin to sublimate away at distances from KIC 3542116 of (cid:46) 0.1 − 0.3 AU (see also Beust et al. 1998). 8. ANOTHER POTENTIAL EXOCOMET CANDIDATE KIC 11084727 Of all the Kepler target stars that were visually ex- amined, the most compelling case for exocomet transits was KIC 3542116, discussed extensively above. However, there was one other target star, KIC 11084727, which ex- hibited a single transit event that was very similar to the three deeper transits found in KIC 3542116. The transit event in KIC 11084727 is shown in Fig. 13 along with a model fit. As is evident from a casual vi- sual inspection, and from the formal model fit (the red curve in Fig. 13), the transit properties are very similar in shape, depth, and duration to those listed in Table 2 for the deeper dips in KIC 3542116. The data validation process for this target was essen- tially the same as described in Sect. 4 for KIC 3542116. All indications are that this dip is astrophysical in origin and is associated with KIC 11084727. The target star KIC 11084727 is a near twin to KIC 3542116 as can be seen from a compilation of their photo- metric properties in Table 1, with nearly identical magni- tudes (at Kp = 9.99), similar Teff values, and comparable radii. The similarities between KIC 3542116 and KIC 11084727 are particularly striking given that the ma- jority of Kepler targets were cooler, Sun-like stars and suggest that comet transits may preferentially happen around stars of this spectral type. The fact that we have detected two individual stars with similar comet-like transit events also suggests that there may be more (perhaps shallower) comet-like tran- sits hidden in the Kepler dataset. 9. DISCUSSION In this section we discuss some possible follow-up ob- servations of KIC 3542116 and KIC 11084727 that may connect these systems to other exocomet systems found with ground-based observations. We also attempt to un- derstand the relative detection sensitivity of dusty tran- sits using photometry and spectral line changes. A num- ber of dynamical effects that might be responsible for driving comets into orbits close to the host star are briefly discussed. Finally, we compare our two systems with Boyajian's star (KIC 8462852). It might be profitable to carry out follow-up ground- based spectroscopic studies of KIC 3542116 to see if any of the same type of spectral line changes such as are found in β Pic, 49 Ceti, HD 42111, HD 172555, and φ Leo can be discerned in KIC 3542116. Of the 16 stars listed by Welsh & Montgomery (2015) as exhibiting spectral line changes that are likely due to exocomets, the magnitudes range from 3.6 to 7.2 with a mean of 5.6. Moreover, the spectral types of these stars range from B9 to F6, with 2 of the 23 being stars of the F spectral type. Thus, aside from the fact that KIC 3542216 and KIC 11084727 are more than an order of magnitude fainter than these other stars, it may still be possible, even if challenging, 20 An exception is magnetite which sublimates at a substantially lower temperature. Fig. 13.- Exocomet-like dip feature in KIC 11084727. The solid red curve is a model fit similar to those described in Sect. 6 and shown in Fig. 3. The shape, depth, and duration of this transit is quite similar to those of transits D992, D1176, and D1268 seen in KIC 3542116. to monitor the spectral line shapes of KIC 3542116 and KIC 11084727 for changes. We believe that connecting photometric transits to spectral transits in the same star could prove very rewarding. Neither KIC 3542116 nor KIC 11084727 shows any par- ticular evidence for being extremely young, e.g., via very rapid rotation or WISE excess flux. Nor is there any specific reason to believe that there is disk activity or populations of minor bodies at orbital separations much greater than these inferred for the comets in this work. Such debris might be expected to exhibit CO emission as is seen in HD 181327 (Marino et al. 2016), Eta Corvi (Marino et al. 2017), and Fomalhaut (Matr`a et al. 2017). Nevertheless, the stars reported on herein, are sufficiently unusual among the Kepler ensemble of 2×105 stars, that it could be worth the gamble to use ALMA to search for CO emission around KIC 3542116 and KIC 11084727. The observations of likely comets in two Kepler stars in this work raise some interesting questions by way of com- parison with the comets (FEBs) inferred from spectral- line changes in a substantial number of primarily bright A stars (e.g., Beust et al. 1990; Welsh & Montgomery 2015). In particular, why are we not detecting 'swarms' of comets as are suggested by the papers on FEBs? Of the 2 × 105 Kepler stars studied continuously for approximately four years, we have found only 6 comet- like dips in one star (KIC 3542116) and 1 dip in KIC 11084727. Presumably, Kepler is not as sensitive to small comets as the spectroscopic methods are. In the spectral approach, the comet is only blocking a small part of the light (a fraction of one spectral line), but which can be readily detectable in the line profile. By contrast, when looking at transits in Kepler data, we are collecting much of the bolometric flux from the star. Therefore, a much larger comet may be required to be detectable by Kepler than by a spectrograph on a large telescope. More quantitatively, we can write the total mass loss rate of a comet crossing the stellar disk of the host star as M (cid:39) Σjwjvj⊥f−1 j (11) where Σj is the observed mass column density of compo- 14 Rappaport et al. nent, j (e.g., dust, CaII, MgII, FeII), assumed uniform over a strip of width wj that extends across the disk of the stellar host; vj⊥ is the material outflow velocity, i.e., relative to the comet, and perpendicular to the line of sight; and fj is the mass fraction of component j in the comet effluents. We can write Eqn. (11) more specifi- cally as applied to the dust case (in regard to the Kepler dips): M (cid:39) 2ρdsdDdR∗v⊥f−1 (cid:20) Dd (cid:21)(cid:20) v⊥ d (cid:21)(cid:20) fd (cid:21)−1 0.001 5 km/s 0.1 (cid:39) 3 × 1011 g s−1 (12) and for the case of CaII ions as observed in the FEB sources: M (cid:39) 2µCaIINCaIIR∗v⊥f−1 CaII (cid:20) NCaII (cid:21)(cid:20) (cid:39) 1011 (cid:21)(cid:20) fCaII (cid:21)−1 v⊥ 1012/cm2 50 km/s 0.001 g s−1(13) where the subscripts "d" and "CaII" refer to dust and the CaII ion, respectively. The other symbols not yet defined are: sd the characteristic size of the dust grains (taken to be 1 µm); Dd the minimum depth of a comet dip that can be detected with Kepler; µCaII the atomic weight of a Ca atom; and NCaII is the average column number density of CaII as seen by a distant observer. In the second line of both equations, the values of some of the important parameters are normalised to sugges- tive illustrative values. The limiting column density was taken from Hobbs et al. (1985) for the case of the bright star β Pic at V (cid:39) 3.8; for stars that are up to a couple of magnitudes fainter, we assume that the sensitivity limit is only a few times higher, which would make the leading coefficient comparable to that for dust transits. The conclusion we draw from these expressions is that ground-based spectral observations of bright stars (mag- nitude 2-6) should be more sensitive, in terms of detect- ing a given M , than are the Kepler observations, but with substantial uncertainties in the choice of parameter val- ues. One caveat is in order, however, when interpreting equations (12) and (13). Presumably the dust will only survive rapid sublimation at distances beyond ∼0.1-0.3 AU (see Fig. 12). By contrast, the atoms (in particular CaII, MgII, and FeII) will mostly be present closer in where the dust, bearing many of the heavier elements, sublimates and the minerals become photo-dissociated and ionised. With regard to the number of comets that should be seen crossing the disk of the host star, this rate should depend heavily on whether the infalling comet orbits are distributed roughly isotropically (lower rate) or if the reservoir of bodies producing the dusty tails has orbits that are coplanar with the angular momentum vector of the system (higher rate if the observer lies in this plane). We do have some limited information on the viewing inclination angle with respect to the spin axis of KIC 3542116. From Table 1 we find v sin i (cid:39) 57.3 ± 0.3 km s−1, R∗ (cid:39) 1.56 ± 0.15 R(cid:12), and a rotation frequency of 0.888 ± 0.04 cycle d−1. When we use this informa- tion to compute the inclination angle, i, we find that 45◦ (cid:46) i (cid:46) 80◦ with 95% confidence. This is suggestive that we could be viewing the system from at least a par- tially favorable in-plane vantage point. It will be helpful to firm up these uncertainties in future work. In Sect. 7.1 we made some initial assessments of the kinds of orbits that were most likely responsible for the exocomet transits we report (see, in particular, the right panel of Fig. 11). There are basically two major dynam- ical mechanisms for generating potentially large numbers of transiting exocoments. These have been very well ex- plored in the context of the best-studied FEB system – β Pictoris. However, we should keep in mind that in β Pic, there is a high preponderance of red-shifted FEB events, which implies a particular orientation for the comet tra- jectories. With this caveat in mind, we note that these dynamical mechanisms involve secular perturbations by a distant planet. First, they may be generated via the Kozai-Lidov mechanism (see, e.g., Bailey et al. 1992). The second mechanism involves resonances, either secu- lar (Levison et al. 1994), or mean-motion (Beust & Mor- bidelli 1996; 2000). In the former case the exocomet or- bits should be roughly isotropically distributed thanks to a rotational invariance of the Kozai Hamiltonian. Con- versely, in the latter case the longitude of periastron of the perturbing planet controls the geometry of the infall. Also, in this case Beust & Valiron (2007) showed that the exocomets may have large inclination oscillations when reaching the FEB state even if they started out with only modest inclinations with respect to the orbit of the per- turbing planet. We note that in the solar system, most sun-grazers are thought to arise from the Kozai mech- anism (e.g., Bailey et al. 1992). In the case of β Pic, the mean-motion-resonance mechanism was favoured to match the abundant statistics of the FEB events in that system. By contrast, for KIC 3542116, with only a few transit events detected, all of the above mechanisms are worthy of consideration. Finally, the deep dips in the flux of KIC 8462852 (aka 'Boyajian's Star'; Boyajian et al. 2016) are worth try- ing to relate to what is observed in KIC 3542116. By contrast, the largest flux dips in the former star reach 22% which is more than two orders of magnitude greater than the transits we see in KIC 3542116. Furthermore, the dips in KIC 8462852 can last for between 5 and 50 days, depending on how the beginning and end points of the dip are defined. These are one to two orders of magnitude longer than for the transits in KIC 3542116. Finally, we note that none of the dips in KIC 8462852 has a particularly comet-shaped profile. There have been a number of speculations about the origin of the dips in KIC 8462852, including material resulting from collisions of large bodies and moving in quasi-regular orbits (Boy- ajian et al. 2016); swarms of very large comets (Boyajian et al. 2016); and even a ring of dusty debris in the outer solar system (Katz 2017). However, there is currently no compelling evidence for any of these scenarios. 10. SUMMARY AND CONCLUSIONS In this work we reported the discovery of six apparent transits in KIC 3542116 that have the appearance of a trailing dust tail crossing the disk of the host star. We have tentatively postulated that these are due to between 2 and 6 distinct comet-like bodies in the system. We also found a single similarly shaped transit in KIC 11084727. Both of these host stars are of F2V spectral types. We have carefully vetted the data from these target Transiting Exocomets 15 stars, including assessing the difference images in and out of transit, analyzing potential video crosstalk, and inspecting the data quality flags associated with the dip events. The vetting also included deep high-resolution imaging studies of our prime target, KIC 3542116. No companion stars were found within an outer working angle of ρ ∼ 5(cid:48)(cid:48), though the nearby star KIC 3542117 (ρ = 10(cid:48)(cid:48)) might plausibly be a bound companion. The spot rotation period in KIC 3542116 is about as long as the durations of the deeper transits we see, and therefore it is difficult to imagine that they are caused by highly variable spots (which tend to produce dips at a fraction of the rotation period). Nonetheless, we cannot categorically rule out the possibility that the transit-like events are caused by some previously unknown type of stellar variability. With this caveat in mind, we proceeded to study these systems under the assumption that the dips in flux are in- deed due to dusty-tailed objects transiting the host stars. We then fit these transits with a model dust tail that is assumed to have an exponentially decaying extinction profile. The model profiles fit the transits remarkably well. The inferred speeds of the underlying dust-emitting body during the transits are in the range of 35 − 50 km s−1 for the deeper transits in KIC 3542116 and for the single transit in KIC 11084727. For the more shallow and narrow transits in KIC 3542116, the inferred speeds are 75 − 90 km s−1. From these speeds we can surmise that the corresponding orbital periods are (cid:38) 90 days (and most probably, much longer) for the deeper transits, and (cid:38) 50 days for the shorter events. Solar system sun-grazing comets typically have ex- tremely long orbital periods (e.g., ∼2300 years for the members of the populous Kreutz group). Halley's comet, which has an apohelion distance of ∼125 R(cid:12), has a pe- riod of (cid:39) 75 years, while the shortest known period for a bright comet is Comet Encke at 3.3 years. The over- all shortest period is Comet 311/Pan-STARRS with a period of 3.2 years. Thus, if either the three deeper or the three more shallow transits in KIC 3542116 are from single orbiting bodies, then the maximum associ- ated periods of 51-92 days would be considerably shorter than for Solar-system comets. The periods of the comets producing the FEB events are largely unknown. How- ever, the characteristic infall velocities associated with the CaII FEB's (∼50 km s−1; e.g., Beust et al. 1990) are compatible with what we find for KIC 3542116 and KIC 11084727. The fact that we find comet-like transits in two Ke- pler target stars holds out the promise that such events are not particularly rare. This is especially true when we note that the survey was made visually and without the aid of a computer search. In turn, the fact that the search was carried out visually raises the issue of its com- pleteness. In this regard, we believe that there was no particular obstacle to finding asymmetric transits with depths of (cid:38) 0.1% and lasting for (cid:38) 1/3 day, even in the presence of significant star-spot activity. Furthermore, we likewise found that data breaks and artefacts would also not have impeded the search. We thus believe that we have found the majority of such comet-like transits in the Kepler data set, though we cannot preclude the possibility that there are many more such features with depths (cid:46) 0.1%. We reiterate that there are striking similarities between KIC 3542116 and KIC 11084727 in terms of both the stellar properties and the comet-like transit events. This is also noteworthy because the majority of Kepler tar- gets were cooler, Sun-like stars. This might suggest that comet transits may preferentially happen around stars of this spectral type, and it would be instructive to try to understand why this might be. One encouraging note in regard to finding more such comet-like transits in other stars is that dips in flux at the (cid:38) 0.1% level and lasting for hours to days should not be particularly challenging for the photometry in the upcoming TESS mission (Ricker et al. 2015). Further- more, the host stars are likely to be bright, plausibly even brighter than the 10th magnitude stars reported on here. Note added in manuscript: After this manuscript was submitted we became aware of a remarkably pre- scient paper by Lecavelier des Etangs, Vidal-Madjar, & Ferlet (1999) which predicted the photometric profiles of exocomet dust-tail transits of their host star (see also Lecavelier des Etangs 1999). The calculated profiles look rather remarkably like the ones we have found and re- ported on here. Therefore, it appears that this current work could help to confirm these earlier predictions, and similarly the predictions may help strengthen the case that we have indeed observed exocomet transits. We are especially grateful to the referee, Herv´e Beust, for a substantial number of very instructive suggestions leading to the improvement of the manuscript. We thank Tabetha Boyajian, Bradley Schaefer, and Ben- jamin Montet for discussions of the long-term photomet- ric variations in KIC 3542116. We also appreciate helpful discussions of comet properties with Nalin Samarasinha and Bruce Gary. A. V. was supported by the NSF Grad- uate Research Fellowship, grant No. DGE 1144152, and also acknowledges partial support from NASA's TESS mission under a subaward from the Massachusetts Insti- tute of Technology to the Smithsonian Astrophysical Ob- servatory, Sponsor Contract Number 5710003554. This work was performed in part under contract with the California Institute of Technology (Caltech)/Jet Propul- sion Laboratory (JPL) funded by NASA through the Sagan Fellowship Program executed by the NASA Ex- oplanet Science Institute. This research has made use of NASA's Astrophysics Data System and the NASA Exo- planet Archive, which is operated by the California In- stitute of Technology, under contract with the National Aeronautics and Space Administration under the Exo- planet Exploration Program. This paper includes data collected by the Kepler mission. Funding for the Kepler mission is provided by the NASA Science Mission direc- torate. Some of the data presented in this paper were obtained from the Mikulski Archive for Space Telescopes (MAST). STScI is operated by the Association of Uni- versities for Research in Astronomy, Inc., under NASA contract NAS5–26555. Support for MAST for non–HST data is provided by the NASA Office of Space Science via grant NNX13AC07G and by other grants and contracts. Facilities: Kepler/K2, FLWO:1.5 m (TRES), Keck Ob- servatory 16 Rappaport et al. REFERENCES Ahn C. P., et al., 2012, ApJS, 203, 21 Akeson, R. L., Chen, X., Ciardi, D., et al. 2013, PASP, 125, 989 Baglin, A., Auvergne, M., Boisnard, L. et al. 2006, 36 COSPAR Scientific Assembly, p. 3749. Bailey, E., Chambers, J.E., & Hahn, G. 1992, A&A 257, 315 Barclay, T., Rowe, J. F., Lissauer, J. J., et al. 2013, Nature, 494, 452 Kurucz R. L., 1992, in Barbuy B., Renzini A., eds, Proc. IAU Symp. 149, The Stellar Populations of Galaxies. Kluwer, Dordrecht, p. 225 Lecavelier des Etangs, A. Vidal-Madjar, A., & Ferlet, R. 1999, A&A, 343, 916 Lecavelier des Etangs, A. 1999, A&AS, 140, 15 Levison, H. F., Duncan, M. J., & Wetherill, G. W. 1994, Nature, Batalha, N.M., Rowe, J. F., Bryson, S. T., et al. 2013, ApJS, 204, 372, 441 24 Beust, H., Vidal-Madjar, A., Ferlet, R., & Lagrange-Henri, A. M. 1990, A&A, 236, 202 Beust, H. & Morbidelli, A. 1996, Icarus, 120, 358 Beust, H., Lagrange, A.-M., Crawford, I. A., Goudard, C., Spyromilio, J., & Vidal-Madjar, A. 1998, A&A, 338, 1015 Beust, H. & Morbidelli, A. 2000, Icarus, 143, 170 Beust, H. & Valiron, P. 2007, A&A, 466, 201 Bohren, C. F., & Huffman, D. R. 1983, Absorption and Scattering of Light by Small Particles (New York: Wiley) Bond, I.A., Udalski, A., Jaroszy´nski, M., et al. 2004, ApJ, 606, 155 Borucki, W. J., Koch, D., Basri, G., et al., 2010, Sci., 327, 977 Boyajian, T. S., LaCourse, D. M., Rappaport, S. A., et al. 2016, MNRAS, 457, 3988 Brogi, M., Keller, C. U., de Juan Ovelar, M., et al 2012, A&A, 545, L5 Bryson, S. T., Jenkins, J. M., Gilliland, R. L., et al. 2013, PASP, 125, 889 Buchhave L. A., et al., 2012, Nature, 486, 375 Carter, J. A., & Agol, E. 2013, ApJ, 765, 132 Claret, A., & Bloemen, S. 2011, A&A, 529, A75 Croll, B., Rappaport, S., DeVore, J., et al. 2014, ApJ, 786, 100 Cutri, R.M., Wright, E.L., Conrow, T., et al. 2013, wise.rept, 1C. Deacon, N. R., Kraus, A. L., Mann, A. W., et al. 2016, MNRAS, 455, 4212 Dressing, C. D. & Charbonneau, D. 2013, ApJ, 767, 95 Eiroa, C., Rebollido, I., Montesinos, B., et al. 2016, A&A, 594, 1 Ferlet, R., Hobbs, L. M., & Vidal-Madjar, A. 1987, A&A, 185, 267 Ford, E.B. 2005, AJ, 129 1706 Foreman-Mackey, D., Morton, T. D., Hogg, D. W., Agol, E., Scholkopf, B. 2016, AJ, 152, 206 Gansicke, B.T., Aungwerojwit, A., Marsh, T.R., et al. 2016, ApJ, 818, 7 Gary, B., Rappaport, S., Kaye, T. G., Alonso, R., & Hambsch, F.-J. 2017, MNRAS, 465, 3267 Madhusudhan, N., & Winn, J.N. 2009, ApJ, 693, 784 Marcy, G. W., Butler, R. P., Williams, E., Bildsten, L., Graham, J. R., Ghez, A. M., & Jernigan, J. G. 1997, ApJ, 481, 926 Marino, S., Matr`a, L., Stark, C., et al. 2016, MNRAS, 460, 2933 Marino, S., Wyatt, M. C., Pani´c, O., et al. 2017, MNRAS, 465, 2595 Matijevic, G., Prsa, A., Orosz, J. A., Welsh, W. F., Bloemen, S., Barclay, T., 2012, ApJ, 143, 123 Matr`a, L., MacGregor, M. A.,Kalas, P., et al. 2017, ApJ, 842, 9 Mayor, M., & Queloz, D. 1995, Nature, 378, 355 McQuillan, A., Mazeh, T., & Aigrain, S. 2014, ApJS, 211, 24 Montalto, M., Riffeser, A., Hopp, U., Wilke, S., & Carraro, G. 2008, A&A, 479, L45 Perez-Becker, D., & Chiang, E. 2013, ApJ, MNRAS, 433, 2294 Prsa, A., Batahla, N., Slawson, R. W., et al. 2011, AJ, 141, 83 Gaia Collaboration: Prusti, T., de Bruijne, J. H. J., Brown, A. G. A., et al. 2016, A&A, 595, 1 Rappaport, S., Levine, A., Chiang, E., et al 2012,ApJ,752, 1 Rappaport, S., Barclay, T., DeVore, J., et al. 2014a, ApJ, 784, 40 Rappaport, S., Swift, J., Levine, A., et al. 2014b, ApJ, 788, 114 Rappaport, S., Gary, B.L., Kaye, T., Vanderburg, A., Croll, B., Benni, P., & Foote, J. 2016, MNRAS, 458, 3904 Rappaport, S., Vanderburg, A., Nelson, L., et al. 2017. MNRAS, in press Rasmussen, C., & Williams, C. 2006, Gaussian Processes for Machine Learning, Adaptive Computation and Machine Learning (Cambridge, MA, USA: MIT Press), 248 Reinhold, T., Reiners, A., & Basri, G. 2013, A&A, 560, 4 Ricker, G. R., Winn, J. N., Vanderspek, R. et al. 2015, JATIS, 1, 14003 Sanchis-Ojeda, R., Rappaport, S., Pall'e, E., et al. 2015, ApJ, 812, 112 Schmitt, J. R., Jenkins, J. M., & Fischer, D. A. 2017, AJ, 153, 180 Sekanina, Z. & Chodas, P. W. 2012, ApJ, 757, 127 Skrutskie, M.F., Cutri, R.M., Stiening, R., et al. 2006, AJ, 131, 1163. Grindlay, J., Tang, S., Simcoe, R., Laycock, S., Los, E., Mink, D., Doane, A., & Champine, G. 2009, ASPC, 410, 101 Grunblatt, S. K., Huber, D., Gaidos, E. J., et al. 2016, AJ, 152, Slawson, R.W., Prsa, A., Welsh, W.F., et al. 2011, AJ, 142, 160 Smith, B. A., & Terrile, R. J. 1984, Sci., 226, 1421 Thompson, S. E., Caldwell, D. A.,et al. 2016a, Kepler Data 185 Release 25 Notes (KSCI-19065-002) Grunblatt, S. K., Huber, D., Gaidos, E., et al. 2017, ArXiv Thompson, S.E., Fraquelli, D., van Cleve, J.& Caldwell, D. e-prints, arXiv:1706.05865 Haywood, R. D., et al. 2014, MNRAS, 443, 2517 Hobbs, L. M., Vidal-Madjar, A., Ferlet, R., Albert, C.E., Gry, C. 1985, ApJ, 293, L29 Jenkins, J. M. 2002, ApJ, 575, 493 Jenkins, J. M., Chandrasekaran, H., McCauliff, S. D., et al. 2010, Proc. SPIE, Vol. 7740, 77400D Jenkins, J. M. 2017, "Kepler Data Processing Handbook" (KSCI-19081-002), https://archive.stsci.edu/kepler/ manuals/KSCI-19081-002-KDPH.pdf Katz, J. 2017, arXiv: 1705:08377 Kiefer, F., Lecavelier des Etangs, A., Augereau, J.-C., Vidal-Madjar, A., Lagrange, A.-M., & Beust, H. 2014a, A&A, 561, 10 Kiefer, F., Lecavelier des Etangs, A., Boissier, J., Vidal-Madjar, A., Beust, H., Lagrange, A.-M., H´ebrard, G., & Ferlet, R.2014b, Nature, 514, 462 2016b, "Kepler Archive Manual", KDMC-10008-006, http://archive.stsci.edu/kepler/documents.html Tuthill, P., Lloyd, J. Ireland, M., et al. 2006, SPIE, 6272, 103 Tuthill, P., Lacour, S., Amico, P., et al. 2010, SPIE, 7735, 1O Uehara, S., Kawahara, H., Masuda, K., Yamada, S., Aizawa, M. 2016, ApJ, 822, 2 van Cleve, J. & Caldwell, D. A. 2016, "Kepler Instrument Handbook", KDMC-19033-002, http://archive.stsci.edu/kepler/documents.html Vanderburg A., & Johnson J. A., 2014, PASP, 126, 948 Vanderburg, A., Johnson, J. A., Rappaport, S. 2015, Nature, 526, 546 van Lieshout, R., & Rappaport, S. 2017, to appear in Handbook of Exoplanets, H.J. Deeg & J.A. Belmonte, eds. (Springer) van Lieshout, R., Min, M., & Dominik, C. 2014, A&A, 572, 76 Wang, J., Fischer, D. A., Barclay, T., et al. 2015, ApJ, 815, 127 Welsh, B. Y., & Montgomery, S. L. 2015, Advances in Astronomy, Kipping, D. M., Schmitt, A. R., Huang, X., Torres, G., Nesvorn´y, Vol. 2015, Hindawi Publishing Corporation D.,Buchhave, L. A.,Hartman, J., & Bakos, G. ´A. 2015, ApJ, 813, 14 Knight, M. M., & Schleicher, D. G. 2015, ApJ, 149, 19 Kov´acs, G., Zucker, S., & Mazeh, T. 2002, A&A, 391, 369 Kraus, A.L., Ireland, M. J., Huber, D., Mann, A. W., Dupuy, T. J. 2016, AJ, 152, 8 Wolszczan, A. 1994, Science, 264, 538 Wright, E. L., Eisenhardt, P. R. M., Mainzer, A. K., et al. 2010, AJ, 140, 1868 Xu, S., Rappaport, S., van Lieshout, R., et al. 2017, submitted to MNRAS Yi, S., Demarque, P., Kim, Y.-C., Lee, Y.-W., Ree, C. H., Lejeune, T., Barnes, S. 2001, ApJS, 136, 417 Zacharias, N., Finch, C.T., Girard, T.M., Henden, A., Bartlett, Zuckerman, B., & Song, I. 2012, ApJ, 758, 77 J.L., Monet, D.G., & Zacharias, M.I. 2013, ApJS, 145, 44 Transiting Exocomets 17
1501.02219
2
1501
2015-03-31T22:50:33
The complete catalogue of light curves in equal-mass binary microlensing
[ "astro-ph.EP" ]
The light curves observed in microlensing events due to binary lenses span an extremely wide variety of forms, characterised by U-shaped caustic crossings and/or additional smoother peaks. However, all peaks of the binary-lens light curve can be traced back to features of caustics of the lens system. Moreover, all peaks can be categorised as one of only four types (cusp-grazing, cusp-crossing, fold-crossing or fold-grazing). This enables us to present the first complete map of the parameter space of the equal-mass case by identifying regions in which light curves feature the same number and nature of peaks. We find that the total number of morphologies that can be obtained is 73 out of 232 different regions. The partition of the parameter space so-obtained provides a new key to optimise modelling of observed events through a clever choice of initial conditions for fitting algorithms.
astro-ph.EP
astro-ph
Mon. Not. R. Astron. Soc. 000, 000 -- 000 (0000) Printed 8 October 2018 (MN LATEX style file v2.2) The complete catalogue of light curves in equal-mass binary microlensing Christine Liebig,1(cid:63) Giuseppe D'Ago,2,3† Valerio Bozza,2,3‡ Martin Dominik,1,4§ 1SUPA, School of Physics & Astronomy, North Haugh, University of St Andrews, KY16 9SS, Scotland, UK 2Dipartimento di Fisica "E. R. Caianiello", Universit`a di Salerno, Via Giovanni Paolo II 132, 84084 Fisciano (SA), Italy 3Istituto Nazionale di Fisica Nucleare, Sezione di Napoli, Italy 4Royal Society University Research Fellow Accepted, 30 March 2015. Received, 12 March 2015; in original form, 23 December 2014. ABSTRACT The light curves observed in microlensing events due to binary lenses span an extremely wide variety of forms, characterised by U-shaped caustic crossings and/or additional smoother peaks. However, all peaks of the binary-lens light curve can be traced back to features of caustics of the lens system. Moreover, all peaks can be categorised as one of only four types (cusp-grazing, cusp-crossing, fold-crossing or fold-grazing). This enables us to present the first complete map of the parameter space of the equal-mass case by identifying regions in which light curves feature the same number and nature of peaks. We find that the total number of morphologies that can be obtained is 73 out of 232 different regions. The partition of the parameter space so-obtained provides a new key to optimise modelling of observed events through a clever choice of initial conditions for fitting algorithms. Key words: Gravitational lensing: micro -- methods: numerical 1 INTRODUCTION Einstein (1936) showed that the light curve of a source microlensed by a single foreground compact object is given by an extremely simple symmetric bell-shape, described analytically by a very com- pact formula, now known as Paczy´nski curve (Paczynski 1986). It is somewhat frustrating that by adding just another lens the com- plexity of microlensing explodes so dramatically that after almost 30 years of active theoretical and observational research a complete classification of all possible light curve morphologies is still miss- ing even in the simplest static case! The lack of a complete knowl- edge of the light curve zoology represents a considerable handi- cap in the modelling of real microlensing events. In fact, in order to set initial conditions for fitting, one may follow two routes: ei- ther blindly set-up a dense grid in the parameter space or identify good initial seeds with light curve morphologies close to the one we wish to model. The first approach is more systematic but can be time consuming and redundant; furthermore, it does not guarantee the completeness of the exploration of all possible corners, which may remain hidden in the space between consecutive points in the grid. The second approach promises to be more efficient in terms of computing time but needs to be supported by a robust and rigorous theoretical framework in order to be safely pursued. (cid:63) [email protected][email protected][email protected] § [email protected] c(cid:13) 0000 RAS Historically, it is not uncommon for modellers to explore spe- cific morphological traits of light curves to narrow down the pa- rameter space to be searched, as has been done by authors such as Mao & Di Stefano (1995); Dominik & Hirshfeld (1996); Di Ste- fano & Perna (1997); Albrow et al. (1999a); Dominik (1999a); Han & Gaudi (2008), but literature that systematically covers the whole range of possible morphologies is more scarce. The modelling of observed multiple-lens microlensing light curves requires extensive computation of the magnification curves. Much effort has been in- vested into speeding up the modelling process, by improving the parametrisation (An et al. 2002; Cassan 2008; Bennett 2010; Ben- nett et al. 2012; Penny 2014), by employing neural networks to map light curve features to model light curves (Vermaak 2007). Of course this development happened alongside of substantial ad- vances in the code implementation of existing algorithms. Mao & Di Stefano (1995) discussed a new method for mod- elling binary microlensing events: the positions and amplitudes of binary light curve extrema are compared to those stored in a pre-compiled (unblended, point-source) light curve library to find promising candidate events, which in turn provide initial parameter sets for a more conventional fitting procedure. This approach works well for multi-peak events, where the source trajectory passes over or close to the binary caustics. Di Stefano & Perna (1997) developed the library approach fur- ther by describing any binary-lens light curve by the set of coeffi- cients of Chebyshev basis polynomials. They note that the Cheby- shev expansion will never exactly match the microlensing light curve, because there will be extra extrema and inflection points, but 2 C. Liebig et al. an arbitrarily precise agreement can be achieved (limited by com- putational power) by further expansion. In this way, a model search can be refined until the photometric precision of the data points is matched. They find model parameter solutions to smooth and caustic-crossing light curves by comparing the rough characteris- tics of the light curve (positions of extrema and inflection points and the magnification values at these points) with a pre-computed light curve library and then searching the nearby environment in the physical parameter space with an increased sampling density until they find a match (or multiple matches) that satisfies the desired precision. In principle, this method is quite good at finding degen- erate solutions and higher-order or even non-microlensing param- eters can easily be integrated, but again it remains unclear whether all relevant parameter-space regions have corresponding entries in the library. The optimistic assertion that the "morphological fea- tures change in a way that is gradual and consistent as the physical parameters are changed" (Di Stefano & Perna 1997) is most likely true for smooth light curves, but for caustic-crossing light curves, we know that very small changes in the source trajectory can have dramatic implications for the number of extrema and their relative positions. Night, Di Stefano & Schwamb (2008) make a broad distinc- tion between smooth light curves and caustic-crossing light curves, but the classification is not based on the light curve itself, but on the source trajectory and its closeness to the caustics, i.e. the known simulation parameters, not the observable data. They come to the conclusion that the ratio of smoothly-perturbed to caustic-crossing binary-lens light curves is rather low in survey detections, which can partly be explained by the fact that caustic-crossing peaks stand out unambiguously, whereas smooth perturbations often can have a range of competing explanations (such as binary sources, parallax effects, orbital motion). In Bozza et al. (2012), a detailed morphological assessment is used for the modelling of OGLE-2008-BLG-510 and furthermore the groundwork is laid for a real-time binary event modelling code (further based on Bozza (2001) and Bozza (2010)). The code re- lies on a wide choice of starting conditions ("seeds") from where a search for local χ2 minima is carried out. The choice of seeds is based on the morphology of the binary caustics, with the assump- tion that binary-lens light curves sampled from a given region of the parameter space lie on a smooth slope of the χ2 landscape as long as the morphology of the light curves does not change. The morphology is understood, in this case, as a given peak sequence of caustic crossings and grazings, with any newly created or destroyed peak leading to a change in morphology. Our intention, with the present work, is to take the move from the path traced by Mao & Di Stefano (1995) and Di Ste- fano & Perna (1997) and achieve the first complete classification of light curves in the binary microlensing problem. By studying peak-number plots, we can separate groupings of light curves in the binary-lens parameter space. We are not concerned with di- rectly establishing light curve models, but we want to ensure that we classify all possible light curves. We then want to improve our understanding of the relations between the parameter space and the light curves. The variety of microlensing light curves can seem overwhelm- ing, but the trained eye recognises familiar patterns and translates them back to the parameter space. In fact, the shape of a microlens- ing light curve does follow certain rules: not any arbitrary curve can be interpreted as a microlensing light curve. Specifically, the lim- ited topologies of the binary lens magnification maps allow only for a limited range of light curve morphologies. Consequently, the fundamental idea of this work is to iden- tify the building blocks of microlensing light curves and develop a classification scheme that can be directly applied to observed light curves and that allows for a significant narrowing of the mod- elling parameter space, while, unlike any other approach, guaran- teeing completeness. We want to gain a good understanding of the range of possible light curves and how the identified morpholog- ical classes relate to subspaces of the modelling parameter space. As a first step, we focus on an in-depth study of the equal-mass binary lens, while the framework developed applies to the general case. On reviewing the properties of this special case in Sec. 2, we use the opportunity to introduce a convenient notation for caustic elements. Sec. 3 introduces our morphology classification scheme, which is based on the four fundamental peak types that occur in mi- crolensing; we also discuss the practicalities, such as the light curve simulation, the peak counting and the identification of iso-maxima regions with light curve morphologies. In Sec. 4, we summarise and discuss the current results of this study, we leave some further con- siderations to Sec. 5, and stress its future potential in Sec. 6, while the bulk of the content is shown in tabular and graphical form in Table 3 and in Figures 11 to 18. 2 MICROLENSING OF EQUAL-MASS BINARY SYSTEMS 2.1 Parametrisation Gravitational microlensing is characterised by the angular Einstein radius (cid:114) θE = 4GM c2 DS -- DL DLDS , (1) where M is the total mass of the (foreground) lens object, while DL and DS denote the lens and source distances from the observer. In the course of a microlensing event, the separation be- tween each pair of images is of the order of θE, which is less than a milliarcsecond for typical observed events with the source in the bulge (DS ∼ 8 kpc), and the lens being a main sequence star half- way to the source (DL ∼ 4 kpc, M ∼ 0.3M(cid:12)) If one assumes uniform, rectilinear relative proper motion µ between the lens and source, the magnification due to a single point lens is described by only three parameters: u0, the closest angular impact of the source to the centre of mass expressed in units of θE, the Einstein radius crossing time tE ≡ θE/µ, and the time of closest approach t0 of the source to the centre of mass of the lens system, which is typically used to fix the epoch of observations, but is irrelevant for the light curve shape. Beyond the single lens parameters, we need the binary mass ratio q = m2/m1, where m1 is the primary mass of the binary lens system and m2 the secondary mass in units of the total mass of the system M, the angular sepa- ration of the binary components s in units of θE, the angular source star radius ρ still in units of θE, and the angle α, between the di- rection vector from the primary to the secondary and the direction of the source relative to the lens, see also Fig. 1. We assume uni- form, rectilinear relative proper motion between source and lens for the simulations and ignore higher-order effects. The observed light curve is the sum of the source flux FS, amplified by the microlens- ing effect A(t; u0, tE, t0, q, s, ρ, α), and the blend flux FB contributed by unresolved sources F(t) = FSA(t) + FB. (2) For the purposes of this morphological study, FS and FB have no c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 u0 > 0 1 α binary axis Lens 2 centre of mass Source Figure 1. Our definition of u0 and α. The impact parameter u0 is positive, when source and lens (centre of mass) pass each other on the right-hand side as projected on the plane of sky. α is the angle between the binary axis (pointing from primary to secondary mass) and the source trajectory. impact as they just represent a multiplicative and an additive factor respectively. Our parametrisation is equivalent to the convention detailed in Skowron et al. (2011, Appendix A), except that we regard the source rather than the lens system as moving, resulting in a differ- ence of αhere = αSkowron -- π. A change by π just means the source is travelling in the opposite direction on the same trajectory which does not affect the morphology of the light curve, in other words it is a time reversal of the light curve. More on the parameter space symmetries in Sec. 3.3. 0 2.2 Caustics In the theoretical treatment of multiple lens systems, caustics are singular lines where the flux of a point source is infinitely magni- fied. Schneider & Weiss (1986) have shown that there are exactly three distinct caustic topologies for the case of an equal-mass bi- nary lens. Erdl & Schneider (1993) confirmed this to be true for arbitrary mass ratios. They also noted the transition points in the binary lens separation where the caustic topology changes depend- ing on the lens mass ratio q (also cf. Dominik (1999b)). A caustic enters the close-separation topology domain when s < sc, and will show the wide-separation topology when sw < s, The three topologies (close, intermediate, wide) are shown in Figs. 2-4 for representative choices of the separation parameter. These figures also contain the labels of the notation to be intro- √ duced and discussed in Section 2.3. An isolated pair of lenses close 2/2 for q = 1) result in three caustics to each other (i.e. s < (Fig. 2): one diamond shaped at the centre of mass, and two small, triangular, secondary caustics set off from the binary axis. If the angular separation of the two lenses is of the order of one Einstein radius, there will be only one central, relatively large, six-cusped √ caustic, see Fig. 3. For the equal-mass binary lens "of the order of" means the exact range 2/2 < s < 2. If the two lenses are far from each other (s > 2), two diamond shaped caustics close to the true position of the lenses result. We recollect that caustic lines are al- ways concave in the coplanar binary lens case relevant for Galactic microlensing applications. Petters, Levine & Wambsganss (2001) go into more mathematical detail in describing caustics through singularity theory of differentiable maps. c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 (cid:32) 1 -- s4 c , 3 (cid:33)3 (cid:17)3 , m1m2 = 1 s8 c (cid:16) 3√ m1 + 3√ m2. s2 w = (3) (4) The catalogue of binary microlensing light curves 3 Figure 2. Caustic feature notation of the close-separation binary lens. Moving from point to extended sources, the singularities of the lens map are regularised by an integration over the finite angu- lar disk of the source. As Schneider & Weiss (1986, Fig. 9) have shown, an increase in angular source size leads to decreased peak magnification, a broadening of the peak width and a displacement of the peak, which means that the maximum will occur later when a larger source enters a caustic (and earlier at the exit). Magnification maps of extended sources feature closed high-magnification lines that can be easily recognised as originating from the smoothing- out of caustics. These high magnification lines asymptotically ap- proach the mathematical caustics as the source size shrinks to zero. As an aside, introducing a third lens can lead to exceedingly more complicated caustic structures (Rhie 2002). Danek & Hey- rovsk´y (2015b,a) have set out to explore the full range of triple- lens caustic topologies. To quote just one very specific example, the case of three masses positioned at the tips of an equilateral tri- angle with two equal masses at (1 -- µ)/2 and a third mass at µ, boasts 10 different caustic topologies. Many of those can be found in other triple-lens scenarios, but the list of ten is nowhere close to covering the whole range possible. 2.3 Notation for caustic elements All extrema of a binary-lens light curve can be traced back to fea- tures of the caustic of the lens system. We have developed a "short- hand" notation for these features, sketched in Figures 2, 3 and 4 and listed in Table 1. In this study, we use and depict this shorthand only for the equal-mass binary lens, but we point out its universal applicability to binary-lens caustics of any mass ratio. We denote folds of a caustic with a lower-case letter and cusps with an upper-case letter. Local maxima arise either when the source trajectory approaches or crosses a fold or a cusp. We discuss the peak types in detail in Sec. 3.1. We recall that the magnification of a point source diverges as LensbtBt2Bt1Bb1Bb2bb12topbottom2A1Abp1aatp1bs1ats1ats2abs2atp2abp2abpCCbstpCtsCsecondary causticsecondary causticprimary caustic 4 C. Liebig et al. Figure 3. Caustic feature notation of the intermediate-separation binary lens. Figure 4. Caustic feature notation of the wide-separation binary lens. fc,on/δy if one hits the cusp along its axis and as fc,off/δy2/3 if one hits it off-axis (Schneider, Ehlers & Falco 1992). For fold crossing, the magnification diverges as ff/δy1/2 when approaching the singu- lar line from the inside. Following these basic analytic formulae, all cusps are strongly magnifying compared to their immediate sur- roundings. The strength of magnification varies considerably be- tween one point on a fold line and another depending on the factor ff, which becomes weaker as we move off the binary lens axis. It is the cusps closest to the binary axis that are the strongest in comparison. In the equal-mass binary case, regardless of the spe- cific topology, the points of maximum magnification are the two "A"-cusps on the binary axis, followed by those parts of the "a"- folds closest to the axis. The four off-axis cusps ("B") in the intermediate case, cf. Fig. 3, can be traced across different separations. When the two lenses are moved closer together, the a-folds will eventually merge and split the single caustic line into three separate caustics. The newly created cusps are denoted by "C". A similar metamorphosis Notation Meaning a, b A, B, C, D 1 2 t b p s [ a. . . ; [ A. . . . . . a ]; . . . A ] [ . . . a . . . ] . . . A . . . fold cusp nearer binary mass 1 nearer binary mass 2 "above" binary axis (i.e. left of binary vector) "below" binary axis (i.e. right of binary vector) primary caustic (in close-separation case) secondary caustic (in close-separation case) caustic entry (via fold; via cusp) caustic exit (via fold; via cusp) fold grazing (always inside (or on) caustic for binary case) cusp grazing (always outside (or on) caustic for binary case) Table 1. Caustic feature notation, also illustrated in the sketches in Fig- ures 2, 3 and 4. takes place, when the two lenses are set further apart, except that in this case the "b"-folds will merge to form the new "D"-cusps. In the close topology, the closer the two lenses are posi- tioned, the further the two triangular, secondary caustics will move out from the axis and they will continually decrease in size and strength, whereas the central caustic only decreases in size but gains in strength, until the binary lens becomes indistinguishable from a single lens for s → 0. Conversely, in the wide topology, the two arrow-shaped caus- tics become more and more symmetric towards a diamond shape and decrease in size, until for s → ∞ the B-cusps point perpen- dicular to the axis and the D-cusps become more equal in strength to the A-cusps. Ultimately the two caustics shrink to two points, at which stage two independent single lenses will be observed rather than one binary system. All peaks arising from features closer to or facing the lens on the left side are furnished with an index "1", whereas those nearer the right side lens are indexed "2". We also want to distinguish the symmetric caustic features, which are mirrored across the binary axis. Quite arbitrarily, we denote them with "t" or top, if they are on the left-hand side of the binary axis (looking from primary to secondary) and "b" or bottom, if they lie on the right-hand side. Figures 2, 3 and 4 better illustrate the "logic" behind this choice. In the special case of an equal-mass binary under examination, we have a second symmetry axis through the centre of mass, i.e. through the midpoint between the two lenses and perpendicular to the binary axis. This does not affect the choice of notation. The chosen caustic feature notation scheme covers all scenarios with two point lenses, including mass ratios very different from unity. The notation scheme is summarised in Table 1. 3 CLASSIFICATION SCHEME AND METHODOLOGY Having revisited the basic structure of equal-mass binary lenses and having established an alphanumeric notation to identify ev- ery fold and cusp in each of the three topologies, we now move to the classification of microlensing light curves. First, we define a light curve morphology based solely on observable features of light curves (Section 3.1). By spanning the whole parameter space of binary microlensing, we simulate light curves (Section 3.2) and assign them to the corresponding morphology class. In this way, we can identify every region in the parameter space in which the same morphology arises as the result of the encounter of a determi- nate sequence of caustic features by the source along its trajectory (Section 3.3). c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 at2Bt1btBt2A2A1ab1ab2bbBb2Bb1at112topbottombBaaD2bb2t2Bb2b2t2t2A2bbBt1b1b1t1t1aaBb1D1A112bottomtop 3.1 The four peak types in microlensing Given that the most obvious characteristic of microlensing light curves is the sequence and shape of their local extrema, this se- quence provides a natural taxonomic key for our light curve clas- sification scheme. We propose that a class of light curves can be identified by the common sequence of peak types. We then recog- nise that any microlensing light-curve maximum is created by one of four basic mechanisms. We discuss the four peak types in detail below, but in short summary they are: (i) a cusp grazing (C), (ii) a caustic fold entry (F-) or exit (-F), (iii) a cusp entry (C-) or exit (-C), (iv) a fold grazing (-F-). Now, in detail: (i) the cusp grazing, C: The peak that arises when the source passes outside the caustic but close enough to one of the cusps to pass over the lobe of increased magnification, is a "cusp grazing". We unambiguously call a light curve "cusp-grazing", if the source trajectory is outside the caustic pre and post-peak and only a single peak results. The Paczy´nski curve can be understood as a grazing of the point caustic (or infinite-order cuspoid) of the single lens. The name Paczy´nski curve should be reserved for sin- gle lens light curves only, but in the limits where a binary lens re- sembles a single lens, when the source does not pass close to the caustics or when the caustics are very small relative to the solid angle of the source, a single-peaked light curve will result. We do not register any morphological difference to the cusp grazing in the narrow sense. (ii) the fold entry/exit, F-/-F: When the source enters on a caus- tic fold, this creates a very distinctly shaped curve (cf. Schneider, Ehlers & Falco (1992); Gaudi & Petters (2002)), with a steep, al- most vertical rise followed by a more parabolic fall, which does not descend as low as the caustic-exterior magnification. The morphol- ogy is mirrored in the fold exit. A pair of fold entry and exit peaks give rise to the familiar double caustic crossing signature. (iii) the cusp entry/exit, C-/-C: If the caustic is entered or exited along a cusp, the peak will have a more symmetric shape, because the lobe outside the caustic and the close proximity of the fold lines on the inside of the caustic attenuate the gradient of the passage on both sides. The fact that the magnification in the caustic interior is increased can help to distinguish it from a cusp-grazing1. (iv) the "interior fold approach" or fold grazing, -F-: This type of peak occurs inside the caustic, while the source trajectory passes close to a caustic fold. Due to the concavity of the caustic lines, the fold-grazing peak will only be observed if it is an interior approach. A special case is the peak that occurs when two or more caustic lines are close enough or strong enough to raise the magnification of an extended area between them, giving rise to a peak that cannot be directly attributed to one single fold. These "building blocks" of microlensing light curves can be sequenced, subject to a few rules: • a caustic entry must be followed by a caustic exit2 • a caustic exit cannot occur, if the caustic has not been entered before 1 Mao, Witt & An (2013) have recently shown that this is not necessarily the case for a multi-planar lens distribution. 2 For n > 3 lenses the number of entries and exits may be unequal as caustic lines can be intersecting and nesting. For n = 2, one caustic entry must be followed by one caustic exit, before another caustic entry can occur. c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 The catalogue of binary microlensing light curves 5 • a fold grazing can only take place inside a caustic • a cusp grazing can only take place outside a caustic • due to the concave curvature, a straight caustic-crossing tra- jectory must exit by a fold (or cusp) different from the one through which it entered (Cassan et al. 2010) All binary microlensing light curves (in the parameter space considered in this study) adhere to these rules, but just conforming to these rules does not guarantee a microlensing light curve since the possible caustic topologies are limited (Erdl & Schneider 1993). It is well known that similar light curve morphologies may arise in completely different situations, with source trajectories in- teracting with different cusps or folds in different topologies. Such disconnected regions can be identified by specifying the folds and cusps involved using the notation introduced in Sec. 2.3. Then the symbols identifying a sequence of peaks conforming to a specific morphology class (e.g. F-F C), can be replaced by the correspond- ing caustic elements involved (e.g. [at1bt]Bt2). Since the folds and cusp symbols already carry subscripts, in order to generate more reader-friendly sequences, we indicate the caustic entry and exits by square brackets and suppress the bar for the grazings. So a fold entry is "[a...", a cusp entry "[A...", with the exit being "...]". A fold grazing is "[...a...]" and a cusp grazing is given by "A". These notations detailing the caustic features involved in the light curve morphology sequence are also summarised in Table 1. We will use the synthetic notation (e.g. F-F C) for identifying a light curve mor- phology class irrespective of its possible interpretations in terms of source trajectories and caustics involved, and the detailed notation ([at1bt]Bt2 in this example) to identify the iso-maxima region(s) in the parameter space giving rise to that specific morphology. To see an example of a light curve classification "at work", consider the light curve in Fig. 11(h) where we see a (symmetric) cusp entry (C-) paired with an (asymmetric) fold exit (-F) and a post-caustic grazing of the cusp lobe (C), . C-F(cid:124)(cid:123)(cid:122)(cid:125) caustic traversal C(cid:124)(cid:123)(cid:122)(cid:125) cusp lobe The detailed sequence specifying the folds and cusps involved in this light curve is [A1at2]A2. Fig. 12(d) gives a nice example with a clear-cut fold entry (F-), followed by a second peak still inside the caustic, which can only be an inner fold approach (-F-), a fold exit (-F) and followed by a final cusp lobe grazing (C), so we classify it as F-F-F(cid:124)(cid:123)(cid:122)(cid:125) caustic traversal C(cid:124)(cid:123)(cid:122)(cid:125) cusp lobe . The detailed sequence specifying the folds and cusps involved in this light curve is [ab1at1at2]Bt2. Keeping the caustic geometry fixed and displacing the source trajectory, we can appreciate the changes in the light curve mor- phology, with peaks merging or disappearing while other peaks ap- pear or separate in two. These transition morphologies need some further attention in order to be assigned to specific classes without ambiguities. In this respect, consider the case of Fig. 5, representing the morphing from two fold crossings F-F to a cusp grazing C. When the extended source trajectory cuts a cusp nearly perpendicularly to its axis, the light curve features a transition morphology with a single peak preceded and followed by derivative discontinuities, typical of fold crossings (trajectories ST2 and ST3 in Fig. 5). Intro- ducing a new intermediate "cusp cutting" class would not be very useful, since the detection of the two discontinuities at the base of 6 C. Liebig et al. ST 1 ST 3 ST 2 ST 4 Figure 5. Comparison of fold-crossing and cusp-grazing source trajectories (ST 1 to 4) and resulting light curves. The angular source size is indicated by the white circles. From left to right, the light curve morphology evolves from a double fold crossing F-F for ST 1 to a cusp grazing C (ST 2, 3 and 4). Where exactly this transition occurs depends on the angular source size; with a smaller source, ST 2 would also lead to a double-peaked fold crossing. the peak could never be unambiguously assessed in real observa- tions. Only a very detailed analysis of the light curve would dis- tinguish a cusp-cutting from a cusp-grazing trajectory. Keeping in mind that the purpose of our study is to identify regions in the pa- rameter space that may give rise to independent seeds for model searches, we assign these cusp-cutting peaks to the broader cusp grazing class, extending its definition by including all trajectories for which the cusp cutting does not give rise to two fold-crossing peaks with a dip in between. In some sense, this statement is al- ready contained in the above definition, in which we required that the source is outside the caustic pre- and post-peak and only one peak occurs. This specific example should help avoiding any con- fusion. It follows that the peak classification does not just depend on the source trajectory relative to the lens positions, but equally on the angular source size relative to the caustic size. I.e. a given source trajectory (e.g. ST 2 in Fig. 5) can yield an F-F morphology for a smaller source and a C morphology for a larger source, whereas for a given source size ST 1 can result in an F-F pair, but ST 2 will only show a single peak and be classified as cusp-grazing C. Another situation almost complementary to the previous one occurs when a fold grazing morphs into two fold crossings as the source trajectory changes from fully internal to tangent and then ρ = 0.01 ρ = 0.03 ρ = 0.05 ρ = 0.07 Figure 6. Classification in the case of a beak-to-beak metamorphosis. The magnification curves result from the same source trajectory, but with dif- ferent source sizes (as indicated by the white circles). The smallest source produces an unambiguous fold-grazing, as the central peak occurs inside the caustic (C-¯F-C). Interestingly, the larger sources create a central pair of peaks instead, thus leading us to classify the light curves as C-C C-C. This might seem counterintuitive, before one considers the convoluted magnifi- cation pattern, where it becomes clear that a larger source shifts the posi- tion of the beak-to-beak fold merger -- thereby causing the caustic topology change to occur at a smaller separation compared to the smaller source. secant to the fold. Adopting the same convention as before, we ex- tend the "fold grazing" class to include the transition peak occur- ring when the source moves tangentially to the fold, as long as the peak remains single. Transition morphologies can be more complicated than these two cases illustrated above and may also involve changes in the caustic topologies. In Fig. 6, we have a fold-grazing source tra- jectory C-¯F-C, across a caustic that is close to the limits of the intermediate-to-wide transition, which morphs into a cusp exit/entry pair, C-C C-C, with an increased source size. In summary, all sorts of transitions can be safely treated by adopting the extended definitions of cusp grazing and fold grazing classes just described, including the transition peaks before they split into two. Now we are ready to apply our classification scheme to arbitrarily complicated light curves without facing any more am- biguities. 3.2 Light curve simulation and processing In order to achieve a complete classification of binary lens light curve morphologies, we process simulated light curves. We then consider light curves grouped in the parameter space by their num- ber of maxima. The parameter space we want to cover is the equal- mass lens (q = 1), the separation s across all topologies and the c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 source trajectory parameters 0 ≤ α < 2π and u0 as far as new mor- phologies can be expected to occur. We use an extended source with angular radius 0.002 θE. For each light curve we record the num- ber of peaks and visualise the results in peak-number plots (over α and u0). The resulting iso-maxima regions are examined with re- gard to the contained light curve morphologies. Broadly speaking, an iso-maxima region, covering a "bundle" of neighbouring source trajectories, corresponds to a specific sequence of caustic features. One step up in the classification hierarchy, different iso-maxima re- gions are collected in morphology classes (as introduced in Sec. 3). The fixed source size used in our investigation is small enough to probe the caustics of an equal-mass binary lens in detail but large enough to let cusp crossings occur in finite regions of the parameter space. Different choices will cause a slight shift of the boundaries of the iso-maxima regions (cf. Figs. 5 and 6). This point is further discussed in Sec. 5. In our examination of the equal-mass binary lens case, we simulate microlensing light curves for all (relevant) volumes of the (s, α, u0) parameter space. We simulate the light curves with inverse ray shooting, using a software library written in 2010 by Marnach3. Assuming static lenses, this means we can compute magnification maps for every (q, s) set, fold them with the source star profile with a radius ρ and then extract a large number of light curves differ- ing in α and u0 at virtually no computational cost. During the peak counting, numerical noise can create artificial peaks and troughs, especially for source trajectories that run at a small angle to fold lines. To avoid these, we require a minimal difference between the maximum and the minima on either side of 5% of the nearest local minimum value, before a trough-peak-trough occurrence is counted as a peak. Because of this threshold, sometimes true peaks will be disregarded in the maxima counting algorithm. But this is unlikely to make us miss a whole iso-maxima region, as generally the re- gion boundary (where the formerly disregarded peak becomes sig- nificant) will only be slightly shifted in the (u0, α) plane. 3.3 Iso-maxima regions Per examined separation, we plot the number of local maxima per light curve over α, u0 of its source trajectory, see Fig. 7. In the resulting plot, we can identify and isolate regions of a uniform peak number, which we call iso-maxima regions. Due to the inherent symmetries, we can restrict ourselves to the first quadrant, 0 < u0, 0 < α < π/2. Beyond the trivial pe- riodicity of α with period 2π, there are several symmetries in the two-dimensional (u0, α) space. Generally, for a binary lens, (u0, α) ⇔ ( -- u0, -- α) (5) is an exact degeneracy, which is caused by the intrinsic symmetry of the binary lens across the binary axis. Skowron et al. (2009, Ap- pendix A) argues (and this has been common practice for a while, see e.g. Albrow et al. (1999b)) that models for static binaries should be expressed in the range u ≥ 0 and 0 ≤ α < 2π, with the exception of cases that display parallax effect where the apparent source po- sition can appear on both sides of the lens. We generally subscribe to this view, nonetheless it is instructive to, at least once, visualise the "full" parameter space, see Sec. 4. Since we are interested in the morphology only, (u0, α) ⇔ ( -- u0, α + π), (6) 3 Published at https://github.com/smarnach/luckylensing. c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 The catalogue of binary microlensing light curves 7 gives the symmetry of a time reversal (where the sign of u0 has to change according to the convention, because the source now passes the lens on the other side). We can also combine the two, (u0, α) ⇔ (u0, π -- α). (7) For the special case of the equal-mass binary, we also have a perfect degeneracy u0 ⇔ -- u0, (8) i.e. the plot is axis-symmetric in u0. We note that whenever one moves from one iso-maxima re- gion to a neighbouring one, the morphology of the light curve peaks changes -- naturally, because the border will be overstepped when- ever a peak is created or destroyed. We record the caustic feature sequence for the light curves of each region, see Table 2, and re- alise that in a given quadrant, there are no two iso-maxima regions with the same number of peaks that contain the same sequence of caustic features. We then map the caustic features to the broader peak typology, thereby reducing the complexity of the light curve description and enabling us to collate different regions in more general morphology classes. 4 RESULTS Focussing on the equal-mass binary lens, we analysed peak-number plots spanning all three caustic topologies and the two transitioning cases: close (s = 0.5, 0.65), close-to-intermediate (s = 0.7), inter- mediate (s = 0.85, 1.0, 1.5), intermediate-to-wide (s = 2.05) and wide (s = 2.5). As discussed in Sec. 3.2, we were motivated to use an extended source with an angular radius, ρ = 0.002 (in units of θE) and work with a peak threshold of 5% above the nearest min- ima. Within the peak-number plots, we know the light curve com- position in each (substantial) iso-maxima region, i.e. we know which sequence of caustic features produces the observable peaks of all light curves in that region. We note that it is mostly a bijec- tive mapping, with only very few regions containing more than one kind of caustic feature sequences. In no case, do two unconnected regions share the same caustic feature sequence. The light curves (and iso-maxima regions) are collected in morphology classes, where each peak is morphologically classified as one of the following: cusp-grazing, cusp-crossing, fold-crossing or fold-grazing. A substantial subset of morphology classes can be found in all examined separation settings. Other classes only appear when a higher or lower separation leads to multi-caustic topologies, whereas the specific example of a double fold grazing is necessarily limited to the intermediate caustic cases. The extreme variety of binary microlensing phenomenology can be appreciated by summarising the results of our search in a few numbers. We have found 73 different light curve morpholo- gies according to our classification based on the sequence of peaks. These morphologies arise in 232 independent regions of the param- eter space. The simplest morphologies can be obtained in many dif- ferent ways. For example, the simplest caustic crossing light curve class, F-F, can be found in 7 disconnected volumes of the param- eter space. If we add shoulders to this caustic crossing, with the classical sequence C F-F C, we find 9 different volumes. We em- phasise the fact that thanks to our thorough investigation we are able to quantify the exact number of independent physical models that can qualitatively reproduce an observable light curve for the 8 C. Liebig et al. Figure 7. Plot of the number of maxima per light curve in the first quadrant of the (u0, α) parameter space for the equal-mass binary lens at separation s = 1.0 (intermediate caustic topology): white means the light curve has a single peak, dark blue means six peaks, higher values are numerical noise in this instance. Each labelled region represents a grouping of source trajectories and corresponding light curves that follow a specific caustic feature sequence, see Table 2. Rarely are two regions with the same number of peaks directly connected (cf. III b and III g above). first time! More complicated morphologies with multiple caustic crossings are rarer and appear in fewer regions of the parameter space. A microlensing light curve for an equal-mass binary lens can have up to 10 peaks, if the source moves near the vertical axis of a close configuration. 5 FURTHER CONSIDERATIONS Source size A hypothetical point source is often useful in theo- retical studies of the behaviour of gravitational lenses, but because we want to examine the range of real, observable light curve mor- phologies, we use an extended source size of 0.002 θE for our sim- ulations. The source size does influence the shape of a light curve, as discussed in Sec. 3. A pair of fold crossings can be merged into a single peak, a whole caustic can be crossed and appear as a sin- gle peak, but as long as the solid angle of the source area is small relative to the caustic extent, the absolute size will not change the number of distinct morphologies that can be studied. For the stud- ied mass ratio q = 1, we can afford to use a moderately large source that reduces the numerical noise in our samples. Meaningful stud- ies of planetary mass ratios q (cid:46) 10 -- 3, require a smaller source size. We point out that not all of the peak types of Sec. 3.1 can be sim- ulated with a point source: the cusp crossing can only occur, if the point source enters the caustic exactly over the infinitesimal cusp point. The probability for this occurrence is therefore zero. Two peaks will generally merge into one, if their angular sep- c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 The catalogue of binary microlensing light curves 9 Region label I a II a II b II c II d II e II f II g II h II i III a III b III c III d III e III f III g III h III i III j III k III l III m III n III o III p IV a IV b IV c IV d IV e IV f IV g IV h IV i IV j IV k IV l IV m IV n IV o V a V b V c V d V e VI a Caustic feature sequence Morphology class A1 or Bt1 [A1A2] [at1at2] Bt1Bt2 [ab1at2] [at1bt] [ab1bt] A1Bt1 [bbbt] [bbbt] [A1at2]A2 A1[at1at2] [at1btat2] A1[ab1at2] [at1bt]Bt2 [ab1at1at2] A1[at1Bt2] A1[at1bt] [ab1at2]Bt2 [ab1at1Bt2] [ab1bt]Bt2 [ab1at1bt] Bb1[ab1bt] [bbab1bt] [ab1at1]Bt1 Bb1A1B1 A1[at1at2]A2 [at1bt][btat2] A1[ab1at2]A2 A1[at1bt]Bt2 [ab1at1][at1at2] A1[at1at2]Bt2 [ab1at1at2]Bt2 [ab1at1][at1Bt2] [ab1at1bt]Bt2 Bb1[ab1at2]Bt2 [ab1at1][at1bt] [bbab1][ab1bt] Bb1[ab1at1bt] [bbab1at1bt] Bb1[ab1at1]Bt1 [ab1at1][at1at2]Bt2 [ab1at1][at1bt]Bt2 Bb1[ab1at1][at1bt] [bbab1][ab1at1bt] Bb1[ab1at2][at2bt] [bbab1][ab1at1][at1bt] ¯C C-C F-F C-C F-F F-F F-F C-C F-F F-F C-F ¯C ¯C F-F F-¯F-F ¯C F-F F-F ¯C F-¯F-F ¯C F-C ¯C F-F F-F ¯C F-¯F-C F-F ¯C F-¯F-F ¯C F-F F-¯F-F F-F ¯C ¯C ¯C ¯C ¯C F-F ¯C F-F F-F ¯C F-F ¯C ¯C F-F ¯C F-F F-F ¯C F-F ¯C F-¯F-F ¯C F-F F-C F-¯F-F ¯C ¯C F-F ¯C F-F F-F F-F F-F ¯C F-¯F-F F-¯F-¯F-F ¯C F-¯F-F F-F F-F ¯C F-F F-F ¯C ¯C F-F F-F F-F F-¯F-F ¯C F-F F-F F-F F-F F-F Table 2. Caustic feature sequences for the iso-maxima regions in Fig. 7 (q = 1.0, s = 1.0). Each sequence is unique to its region, while the morphology classes generally span several independent regions. aration is smaller than the angular source diameter (disregarding limb-darkening effects). In our simulations the source has a diame- ter of 4×10 -- 3 θE, i.e. peaks within 4×10 -- 4 tE of each other would be missed. We work with the assumption that a larger source size can only lead to a smaller number of identified morphologies. This has been visually demonstrated for q = 1.0, s = 1.0 in Fig. 8. Liebig (2014) also documents the entirety of morphological classes for a source radius of 10 -- 2 θE and they are a subset of the morphology presented here. c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 Figure 8. Comparison of peak-number plots resulting from different source sizes (from top to bottom: ρ = 0.005, 0.01, 0.02 θE), scale and ranges as in Fig. 7: x-axis: 0 ≤ u0 ≤ 1.0, y-axis: 0.0 ≤ α ≤ π/2. The change in iso-maxima regions is subtle, but noticeable. The smallest source not only leads to more iso-maxima regions, but also to more numerical artefacts. Also compare Fig. 7, where ρ = 0.002 θE. Error margin While we aim for completeness, due to the numer- ical nature of our study we have to ignore very small sub-regions of the studied parameter space and therefore might have missed out on a particular light curve morphology. Within this space we have examined all iso-maxima regions larger than 10 by 10 pixel, i.e. 102 × 1/(u0-sampling) × π /(α-sampling), meaning that within a given Einstein radius and with our sampling of 1600, the probabil- 2 10 C. Liebig et al. Figure 9. Extended plot of the iso-maxima regions for s = 1.0 to illus- trate existing symmetries and the seamless continuation of iso-maxima re- gions beyond the first quadrant. We have marked a caustic feature region: the green outline frames the area where the at1 fold gives rise to a light curve peak, more specifically the top and bottom regions contain the fold entry [at1 . . .] whereas the middle region contains the fold exit [. . . at1]. The green shade marks areas where the at1 fold is crossed twice (requir- ing [. . . at1] [at1 . . .] to be part of the light curve). Moving to a slightly smaller u0 from the shaded area, the light curves will display the fold graz- ing [. . . at1 . . .]. ity to observe that particular light curve morphology is smaller than (cid:46) 1/16000. Caustic feature regions It is highly instructive to consider the complete parameter space volume that corresponds to a peak cre- ated by a particular caustic feature. This caustic feature region will cover several iso-maxima regions, where the light curves show one or two peaks due to this particular caustic element, see an example in Fig. 9. If this information content could be properly harnessed, it would provide an immediate key for the mapping of the lens sys- tem to the light curve and from the light curve morphology to the caustic. For caustic-crossing point sources, this problem often reduces to registering the intersection points of the straight source trajec- tory with the caustic, which is a mathematically well-defined prob- lem. However, for non-caustic crossing peaks (i.e. cusp and fold approaches), caustic lines do not provide sufficient information. It would be necessary to study the magnification map around caustics in order to pin down the position of the maximum magnification along the source trajectory and then assign this position to a nearby caustic fold or cusp for classification. This approach appears too complex to implement in practice. 6 CONCLUSION AND FUTURE PROSPECTS We have compiled an unprecedented catalogue of microlensing light curve morphologies for the equal-mass binary lens. We re- alised that all peaks in microlensing light curves can be classified in just four categories: cusp-grazing, cusp-crossing, fold-crossing or fold-grazing. In order to achieve this complete classification, we have developed a general notation scheme for the features of binary-lens caustics. Our tool, plots of peak number over u0 and α, serves to provide insight into the the microlensing parameter space. Before this work, statements of the diversity of binary mi- crolensing light curves were only made on reasonable but vague arguments. With our detailed study we are able to assign numbers to all specific cases and open the way to more quantitative studies of binary microlensing. Apart from the pure taxonomical aspects, which are very in- teresting from the theoretical point of view, Table 3 stands out as a very powerful tool for modellers to relate an observed light curve to all possible regions of the parameter space in which this light curve can be found. This capability would help the construction of more fail-safe algorithms that will guarantee a full exploration of the mi- crolensing parameter space. In practice, seeds for fitting algorithms can be placed in the middle of each iso-maxima region so as to en- sure a full exploration of all possible cases. Among the currently running platforms for modelling using this principle for setting ini- tial conditions, we mention RTModel (Bozza 2010; Bozza et al. 2012). The inclusion of our catalogue in the template library con- sulted by RTModel would further diminish the chances of missing any particular region in the parameter space. Another interesting aspect that can be further investigated is the probability of the occurrence of a given morphology. Having traced the iso-maxima regions in the parameter space, it should not be difficult to translate the volumes of the iso-maxima regions in probabilities normalised by a physically motivated measure. In this way, we would be able to quantify how "rare" or common a mor- phology is. The current iso-maxima plots would already suffice for probabilities at fixed lens separations. However, for a more com- plete and useful study, one should move through different lens sep- arations with a much smaller sampling step, so as to characterise the shapes and the volumes of the regions in a more accurate way. Furthermore, the final result would depend on the assumed prior distribution function for the separations of binary systems. Sum- ming up, the study of the relative frequencies of the different mor- phology classes is certainly one of the most interesting directions opened up by our work, which deserves the greatest attention and an adequate space in dedicated future works. We have only very briefly mentioned the existence of caustic- feature regions as "meta regions" to the iso-maxima regions, i.e. the combination of all iso-maxima regions containing one specific, caustic-related peak. Unfortunately, we have not yet found a good way to extract and preserve the information about these meta re- gions, but in fact they can provide a more fundamental understand- ing of the parameter space, since iso-maxima regions are basically just "stacks" of caustic-feature regions. In contrast to iso-maxima regions, caustic-feature regions are smooth structures and, like the caustics they are derived from, they change continuously over the parameter space. If their boundaries could be analytically derived from the caustic lines, an elegant automatic classification could be achieved. Finally, we must remember that our work is limited to the equal-mass static lens case. Higher order effect such as parallax and orbital motion would dramatically increase the complexity of the classification, adding very few new morphologies (at least in reasonable physical cases) and would mainly distort existing iso- maxima regions. The only really relevant and humanly achievable upgrade of our catalogue should include a variable mass ratio. While previous literature has shown only up to three pairs of caustic crossings for a single microlensing caustic (e.g. Cassan c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 The catalogue of binary microlensing light curves 11 Cassan A., Horne K., Kains N., Tsapras Y., Browne P., 2010, As- tronomy and Astrophysics, 515, Danek K., Heyrovsk´y D., 2015a, arxiv:1501.06519 Danek K., Heyrovsk´y D., 2015b, arXiv:1501.02722 Di Stefano R., Perna R., 1997, Astrophysical Journal, 488, 55 Dominik M., 1999a, Astronomy and Astrophysics, 341, 943 Dominik M., 1999b, Astronomy and Astrophysics, 349, 108 Dominik M., Hirshfeld A. C., 1996, Astronomy and Astrophysics, 313, 841 Einstein A., 1936, Science, 84, 506 Erdl H., Schneider P., 1993, Astronomy and Astrophysics, 268, 453 Gaudi B. S., Petters A. O., 2002, Astrophysical Journal, 574, 970 Han C., Gaudi B. S., 2008, Astrophysical Journal, 689, 53 Liebig C., 2014, PhD thesis, University of St Andrews Mao S., Di Stefano R., 1995, Astrophysical Journal, 440, 22 Mao S., Witt H. J., An J. H., 2013, Monthly Notices of the Royal Astronomical Society Night C., Di Stefano R., Schwamb M., 2008, Astrophysical Jour- nal, 686, 785 Paczynski B., 1986, Astrophysical Journal, 304, 1 Penny M. T., 2014, Astrophysical Journal, 790, 142 Petters A. O., Levine H., Wambsganss J., 2001, Progress in math- ematical physics, Vol. 21, Singularity theory and gravitational lensing. Birkhauser Rhie S. H., 2002, arxiv:astro-ph/0202294v1 Schneider P., Ehlers J., Falco E. E., 1992, Gravitational Lenses. Springer Schneider P., WeissA., 1986, Astronomy and Astrophysics, 164, 237 Skowron J. et al., 2011, Astrophysical Journal, 738, 87 Skowron J., Wyrzykowski Ł., Mao S., Jaroszy´nski M., 2009, Monthly Notices of the Royal Astronomical Society, 393, 999 Vermaak P., 2007, PhD thesis, University of Cape Town X F-F F-F F-F F-F F-F q = 0.8 s = 0.73 u0 = 0.035 α = 1.5707 Figure 10. Five pairs of double caustic crossings in one ten-maxima light curve. This "quintuple F-F" morphology cannot be created with an equal- mass binary. The light curve is plotted as magnification on a logarithmic scale in the range from 1 to 50 (dotted line at magnification 10), the time axis covers four Einstein times. The small plot shows the corresponding caustic topology and source trajectory (4.2 by 4.2 Einstein angles). The caustic is computed with Caustic Finder by Schmidt, published 2008 at causticfinder.sourceforge.net. et al. 2010), it is possible to draw a microlensing trajectory experi- encing five pairs of caustic crossings for a binary lens with a mass ratio slightly smaller than one, see Fig. 10. This is a morphology that is not present in this equal-mass catalogue. Indeed, most of the new morphologies appearing in unequal-mass binary microlensing would come from the close topology. Our plots of the iso-peak re- gions in the (u0, α) space are done at fixed separation s. We can follow the evolution of iso-maxima regions with a variation of s in different plots and it is easy to refine the sampling in s in or- der to catch all possible regions arising only in limited ranges of s. The addition of a new parameter would make the search much more complicated, as we should trace the evolution of iso-maxima regions in a two-dimensional space with the danger that tiny inter- sections may escape a search with a too coarse grid. New tricks are needed in order to carry out this search efficiently and without omissions. The equal-mass catalogue represents a good basis for this exploration and an already powerful map for the comprehen- sion of microlensing of binary systems. ACKNOWLEDGEMENTS This publication was supported by NPRP grant NPRP-09-476-1-78 from the Qatar National Research Fund (a member of Qatar Foun- dation). REFERENCES Albrow M. D. et al., 1999a, Astrophysical Journal, 522, 1022 Albrow M. D. et al., 1999b, Astrophysical Journal, 522, 1011 An J. H. et al., 2002, Astrophysical Journal, 572, 521 Bennett D. P., 2010, Astrophysical Journal, 716, 1408 Bennett D. P. et al., 2012, Astrophysical Journal, 757, 119 Bozza V., 2001, Astronomy & Astrophysics, 374, 13 Bozza V., 2010, Monthly Notices of the Royal Astronomical So- ciety, 1271 Bozza V. et al., 2012, Monthly Notices of the Royal Astronomical Society, 424, 902 Cassan A., 2008, Astronomy and Astrophysics, 491, 587 c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 12 C. Liebig et al. , 2 t B ] 1 t b 1 b a [ , 1 t B e d i W - 2 t B 1 b B , 1 D 1 b B , 1 t B 1 A , 2 t B 1 t B ] 1 t b 1 b b [ , ] 1 t b 1 b a [ 1 D , 1 t B , 1 A ] 1 t b 1 b B [ ] t b b b [ , ] t b 1 b a [ , ] t b 1 t a [ , ] 2 t a 1 b a [ , ] 2 t a 1 t a [ ] t b 1 b B [ , ] 2 t a 1 A [ , ] 2 t B 1 t a [ 1 t B 1 A , 2 t B 1 t B ] 2 A 1 A [ ] 1 p t a 1 p b a [ , ] t b 1 s t a [ , ] 2 p t a 1 p b a [ , ] 2 s t a 1 s t a [ , ] 2 p t a 1 p t a [ 2 t B s t C , 2 t B 1 A , s t C 1 A , p t C 1 A , 1 t B 1 A , 2 t B 1 t B ] 2 p t a 1 A [ , ] 2 t B 1 s t a [ ] 2 A 1 A [ 1 t B , 1 A s t C , p t C , 1 t B , 1 A . s e i g o l o p o t e d i w d n a e t a i d e m r e t n i , e s o l c r o f s e s s a l c y g o l o h p r o m f o w e i v r e v O : 3 e l b a T e t a i d e m r e t n I e s o l C s s a l C y g o l o h p r o M 2 t B ] 1 t b 1 b a [ 1 A , 1 t B ] 1 t a 1 b a [ 1 b B , 2 t B ] 1 t b 1 b b [ 1 b B , 2 t B ] 1 t b 1 t a [ 1 A - - 2 t B 1 A 1 b B , 2 t B 1 t B 1 A , 2 t B 1 D 1 b B , 1 t B 1 A 1 b B 2 t B ] t b 1 b a [ 1 b B , 2 t B ] t b 1 b a [ 1 A , 1 t B ] 1 t a 1 b a [ 1 b B , 2 t B ] 2 t a 1 b a [ 1 b B , 2 t B ] 2 t a 1 t a [ 1 A , 2 t B ] t b 1 t a [ 1 A , 2 A ] 2 t a 1 b a [ 1 A , 2 A ] 2 t a 1 t a [ 1 A - 1 t B 1 A 1 b B ] 2 t a t b 1 A [ ] 1 t b 1 b b 1 b a [ , ] 1 t b 1 t a 1 b a [ ] 2 t a t b 1 b a [ , ] t b 2 t a 1 b a [ , ] t b 1 b a b b [ , ] t b 1 t a 1 b a [ , ] 2 t a 1 t a 1 b a [ , ] 2 t a t b 1 t a [ 2 t B ] 1 t b 1 A [ 2 t B ] t b 1 A [ , ] 2 t B 1 t a [ 1 A , 2 A ] 2 t a 1 A [ , 2 t B ] 2 p t a 1 p b a [ 1 b B , 2 t B ] 2 s t a 1 s t a [ 1 A , 2 A ] 2 p t a 1 p b a [ 1 A , 2 A ] 2 p t a 1 p t a [ 1 A p t C ] 2 p t a 1 p b a [ p b C , 1 t B ] 1 p t b 1 p b a [ 1 b B - ] 2 p t a 1 p t a 1 p b a [ , ] 2 s t a t b 1 s t a [ ] p t C 1 p b a [ p b C , 2 A ] 2 p t a 1 A [ 2 t B s t C 1 A , 1 t B 1 A 1 b B ] t b s t C [ 1 A ] 2 t a 2 t b [ ] 1 b b 1 b a [ , ] 1 t b 1 b b [ ] 1 b b 1 b a [ , ] 1 t b 1 t a [ ] 1 t a 1 b a [ , ] 2 t a 2 t b [ ] 1 t b 1 t a [ , ] t b 1 b a [ ] 1 b a b b [ , ] t b 1 t a [ ] 1 t a 1 b a [ , ] 2 t a 1 t a [ ] 1 t a 1 b a [ , ] 2 t a t b [ ] t b 1 t a [ ] 2 t a t b [ ] t b 1 b a [ , ] t b 2 t a [ ] 2 t a 1 b a [ 2 t B ] 1 t b 1 b b 1 b a [ , ] 1 t b 1 t a 1 b a [ 1 b B , ] 2 t a t b 1 b a [ 1 A , ] 2 t a t b 1 t a [ 1 A , ] t b 1 t a 1 b a [ 1 b B , 2 t B ] t b 1 t a 1 b a [ , 2 t B ] 2 t a 1 t a 1 b a [ ] t b 2 t a 1 b a [ 1 b B ] t b 1 s t a [ ] 1 p t a 1 p b a [ , ] 2 p t a 1 p t a [ ] 1 p t a 1 p b a [ , ] 2 s t a t b [ ] t b 1 s t a [ ] t b 2 s t a 1 s t a [ 1 A , 2 t B ] 2 p t a 1 p t a 1 p b a [ . e g a p t x e n n o d e u n i t n o C - 2 t B 1 D ] 1 t b 1 b a [ 1 A , 2 t B 1 D ] 1 t b 1 t a [ 1 A , 2 t B 1 D ] 1 b b 1 b a [ 1 A 2 t B ] 1 t b 1 b b 1 b a [ 1 A - ] 2 t a 2 t b [ 2 D ] 1 b b 1 b a [ 2 A ] 2 t a 2 D [ ] 1 D 1 A [ 2 A ] 2 t a 2 t b [ ] 1 t b 1 A [ ] 2 t a 2 t b [ ] 1 D 1 b a [ 1 A ] 2 t a 1 t b [ ] 1 b b 1 b a [ 1 A , 1 D ] 1 t b 1 b b [ ] 1 b b 1 b a [ , ] 2 t a 2 t b [ ] 1 t b 1 b a [ 1 A , ] 2 t a 2 t b [ ] 1 t b 1 t a [ 1 A , ] 1 t b 1 t a [ ] 1 t a 1 b a [ 1 b B , 2 t B ] 1 t b 1 b b [ ] 1 b b 1 b a [ - - - ] 2 t a 2 t b [ ] 1 t b 1 A [ 2 t B 1 D ] 1 b b 1 b a [ ] 2 A 2 D [ ] 1 D 1 A [ 2 t B 1 D ] 1 t b 1 A [ ] t b 2 t a 1 b a b b [ , ] t b 2 t a 1 t a 1 b a [ , ] t b 1 t a 1 b a b b [ ] 2 t a t b [ ] t b 1 A [ , ] 2 t B 1 t a [ ] 1 t a 1 b a [ - - , ] t b 2 t a [ ] 2 t a 1 b a [ 1 b B , ] t , ] t b 2 t a [ ] 2 t a 1 t a 1 b a [ - - - - - b 1 t a [ ] 1 t a 1 b a [ 1 b B , 2 t B ] 2 t a 1 t a [ ] 1 t a 1 b a [ ] 2 t a t b [ ] t b 1 b a [ 1 A , ] 2 t a t b [ ] t b 1 t a [ 1 A , ] t b 2 s t a [ ] 1 p t a 1 p b a [ 1 b B , ] t b 1 s t a [ ] 1 p t a 1 p b a [ 1 b B , 1 t B ] 2 p t a 1 p t a [ ] 1 p t a 1 p b a [ ] t b 2 s t a [ ] 2 s t a 1 s t a [ 1 A - - ] t b s t C [ ] 1 p t a 1 p b a [ 2 t B ] p t C 1 p b a [ p b C , ] t b 2 s t a [ s t C 1 A , ] t b 1 s t a [ s t C 1 A , 2 t B s t C ] 1 p t a 1 p b a [ , 2 t B p t C 1 t B s t C ] 1 p t a 1 p b a [ ] 1 p t a 1 p b a [ , ] t b 2 t a 1 t a [ ] 1 t a 1 b a [ ] 2 t a t b [ ] t b b b 1 b a [ , ] t , ] t b 1 t a 1 b a [ ] 1 b a b b [ b 2 t a 1 b a [ ] 1 b a b b [ 2 A ] 2 t a t b 1 t a [ 1 A ] t b 2 s t a 1 s t a [ ] 1 p t a 1 p b a [ 2 t B ] 2 p t a 1 p t a 1 p b a [ 1 b B - - - - - - , 2 t B p t C ] 1 p t a 1 p b a [ p b C , 2 t B p t C ] 2 p t a 1 p b a [ p b C ] t b 1 s t a [ s t C ] 1 p t a 1 p b a [ , 2 t B p t C 1 t B s t C ] 1 p t a 1 p b a [ 1 b B ] 1 p t a 1 p b a [ 1 b B , ] t b 2 s t a [ s t C ] 1 p t a 1 p b a [ - - - ] t b s t C [ ] 1 p t a 1 p b a [ 1 b B ] 1 t a 1 b a [ , ] 1 t b 1 b a [ 1 b B , ] 1 t b 1 t a [ 1 A , 2 t B ] 1 t b 1 t a [ 2 t B ] 1 b b 1 b a [ , ] 1 t b 1 b b [ 1 A , ] 1 t b 1 b b [ 1 b B , 2 t B ] t b 1 b a [ , 2 t B ] 2 t a 1 b a [ ] t , ] t b 1 t a [ 1 A , 2 t B ] t b 1 t a [ b b b [ 1 b B , 2 t B ] 2 t a 1 t a [ , 1 t B ] 1 t a 1 b a [ , ] 2 t a 1 b a [ 1 A , ] 2 t a 1 t a [ 1 A , ] t b 1 b a [ 1 b B , p t C ] 1 p t a 1 p b a [ , p t C ] 2 p t a 1 p b a [ , 2 t B ] 2 p t a 1 s t a [ , 2 t B ] 2 s t a 1 s t a [ , 1 t B , 2 t B ] 2 p t a 1 p b a [ , ] t b 1 s t a [ 1 A , 2 t B ] t b 1 s t a [ , ] 2 p t a 1 p b a [ 1 A , ] 2 p t a 1 p t a [ 1 A 2 t B ] 1 p t a 1 p b a [ ] 1 p t a 1 p b a [ ¯C F - F ¯C ¯C C C - F - C F - F ¯C F - ¯F - F C - F ¯C ¯C ¯C ¯C F - ¯F - C F - C ¯C ¯C F - F ¯C F - F F - F F - ¯F - F ¯C F - ¯F - ¯F - F F - F F - C ¯C ¯C F - F F - C F - F ¯C ¯C F - C ¯C C - F ¯C F - F F - F ¯C F - ¯F - F F - F ¯C F - ¯F - F ¯C ¯C ¯C F - F ¯C F - F ¯C F - F F - C F - F ¯C ¯C F - C C C - ¯C F - F F - C F - F C - F ¯C - C C C C - I I I I I I V I V c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 The catalogue of binary microlensing light curves 13 ] 2 t a 2 t b [ 1 D ] 1 t b 1 t a [ 1 A , ] 2 t a 2 t b [ 1 D ] 1 t b 1 b a [ 1 A , ] 2 t a 2 t b [ 1 D ] 1 b b 1 b a [ 1 A - - - - - - - - - - - - - - ] 2 t a 2 t b [ ] 1 t b 1 b b 1 b a [ 1 A 2 A ] 2 t a 2 t b [ 1 D ] 1 t b 1 A [ 2 A ] 2 t a 2 t b [ 2 D ] 1 D 1 A [ 2 A ] 2 t a 2 t b [ ] 1 D 1 b a [ 1 A 2 A ] 2 t a 2 t b [ 2 D ] 1 D 1 b a [ 1 A 2 A ] 2 t a 2 t b [ 2 D 1 D ] 1 t b 1 A [ , 2 A ] 2 t a 2 t b [ 1 D ] 1 t b 1 b a [ 1 A , 2 A ] 2 t a 2 t b [ 1 D ] 1 b b 1 b a [ 1 A 2 A ] 2 t a 2 t b [ 1 D ] 1 t b 1 t a [ 1 A . e g a p t x e n n o d e u n i t n o C - - - - - - ] 2 t a 2 t b [ ] 1 t b 1 b b [ ] 1 b b 1 b a [ 1 A - - - - - - - - - - - - - - - - - - - - - - - - - - - - - ] t b 2 s t a [ p t C ] 2 p t a 1 p b a [ 1 b B , ] t b 2 s t a [ s t C ] 1 p t a 1 p b a [ 1 b B , ] t b 2 s t a [ p t C ] 1 p t a 1 p b a [ 1 b B - - - ] t b 2 s t a [ ] 2 p t a 1 p t a 1 p b a [ 1 b B ] t b s t C [ ] p t C p b C [ ] s b C b b [ ] t b s t C [ p t C ] 1 p t a 1 p b a [ 1 b B ] t b s t C [ ] 1 p t a 1 p b a [ ] 1 s b a b b [ ] t b 2 s t a [ ] p t C 1 p b a [ ] 1 s b a b b [ 1 t B p t C ] 1 p t a 1 p b a [ p b C 1 b B , 2 t B p t C ] 1 p t a 1 p b a [ p b C 1 b B , 2 t B p t C ] 2 p t a 1 p b a [ p b C 1 b B 2 t B s t C ] p t C 1 p b a [ p b C 1 b B 2 t B s t C p t C ] 1 p t a 1 p b a [ 1 b B ] t b 1 s t a [ s t C p t C ] 1 p t a 1 p b a [ 1 b B , ] t b 2 s t a [ s t C p t C ] 1 p t a 1 p b a [ 1 b B - - - ] t b s t C [ p t C ] 1 p t a 1 p b a [ ] 1 s b a b b [ 2 t B s t C ] p t C 1 p b a [ p b C s b C 1 b B 2 t B s t C p t C ] 1 p t a 1 p b a [ p b C 1 b B ] t b s t C [ ] p t C p b C [ s b C ] 1 s b a b b [ ] t b s t C [ p t C ] 1 p t a 1 p b a [ p b C 1 b B ] t b 2 s t a [ ] 2 p t a 1 p t a [ ] 1 p t a 1 p b a [ 1 b B , ] t b 2 s t a [ p t C ] 2 p t a 1 p b a [ p b C ] 1 s b a b b [ , ] t ] t b 1 s t a [ p t C b 2 s t a [ p t C ] 1 p t a 1 p b a [ p b C ] 1 p t a 1 p b a [ p b C ] 1 s b a b b [ ] 1 s b a b b [ ] t b 1 s t a [ s t C ] p t C 1 p b a [ p b C ] 1 s b a b b [ , ] t b 2 s t a [ s t C ] p t C 1 p b a [ p b C ] 1 s b a b b [ ] t b 2 s t a [ s t C ] p t C s b C [ s b C ] 1 s b a b b [ ] t b s t C [ ] p t C 1 p b a [ p b C s b C ] 1 s b a b b [ ] t b 1 s t a [ s t C p t C ] 1 p t a 1 p b a [ ] 1 s b a b b [ , ] t b 2 s t a [ s t C p t C ] 1 p t a 1 p b a [ ] 1 s b a b b [ ] t b s t C [ p t C ] 1 p t a 1 p b a [ p b C ] 1 s b a b b [ , ] t b 1 s t a [ p t C ] 1 p t a 1 p b a [ ] 1 s b a b b [ , ] t ] t b 2 s t a [ p t C b 2 s t a [ p t C ] 1 p t a 1 p b a [ ] 1 s b a b b [ ] 2 p t a 1 p b a [ ] 1 s b a b b [ e d i W - - - ] 2 t a 2 t b [ 1 D ] 1 t b 1 A [ 2 A ] 2 t a 2 t b [ ] 1 b b 1 b a [ 1 A , 2 t B ] 1 t b 1 b b [ ] 1 b b 1 b a [ 1 A , 2 A ] 2 t a 2 t b [ ] 1 t b 1 b a [ 1 A , 2 A ] 2 t a 2 t b [ ] 1 t b 1 t a [ 1 A 2 A ] 2 t a t b [ ] t b 1 t a [ 1 A ] t b 2 t a [ ] 2 t a 1 b a [ ] 1 b a b b [ , ] t b 2 t a [ ] 2 t a 1 t a [ ] 1 t a 1 b a [ , ] t b 1 t a [ ] 1 t a 1 b a [ ] 1 b a b b [ , ] t b 2 s t a [ ] 2 s t a 1 s t a [ ] 1 p t a 1 p b a [ e t a i d e m r e t n I - - - . ) d e u n i t n o c ( s e s s a l c y g o l o h p r o M : 3 e l b a T e s o l C - , ] t ] t b 1 s t a [ ] 1 p t a 1 p b a [ ] 1 s b a b b [ b 2 s t a [ ] 1 p t a 1 p b a [ ] 1 s b a b b [ 2 t B ] 2 p t a 1 p t a [ ] 1 p t a 1 p b a [ 1 b B 2 t B s t C p t C ] 1 p t a 1 p b a [ 2 t B ] p t C 1 p b a [ p b C 1 b B s s a l C y g o l o h p r o M F - F ¯C F - C ¯C ¯C ¯C F - F ¯C C - F ¯C ¯C F - F F - F F - F ¯C F - F F - F ¯C F - F ¯C F - F ¯C F - F F - ¯F - F ¯C ¯C F - F ¯C F - C F - C C C C - F - F - C ¯C F - F ¯C F - C F - F F - F F - F C - F F - F ¯C F - F ¯C C C - ¯C F - F C - F ¯C ¯C ¯C F - F ¯C ¯C ¯C ¯C C - F ¯C ¯C ¯C ¯C ¯C F - F ¯C F - F ¯C F - F F - F F - F ¯C ¯C F - F ¯C F - C ¯C F - F F - F ¯C ¯C C - F ¯C ¯C ¯C ¯C ¯C ¯C F - F ¯C ¯C - F - C C C ¯C F - F F - C ¯C F - F ¯C ¯C ¯C F - F ¯C C - F ¯C ¯C F - F ¯C ¯C F - C ¯C F - F ¯C F - F ¯C F - F F - F F - F ¯C F - F ¯C F - F ¯C F - F F - F ¯C C C ¯C F - F - F - F ¯C C - F ¯C F - F F - C C - F ¯C ¯C F - F F - C ¯C F - F ¯C F - F F - F ¯C ¯C F - F F - F I V I I V I I I V c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 14 C. Liebig et al. e d i W e t a i d e m r e t n I e s o l C . ) d e u n i t n o c ( s e s s a l c y g o l o h p r o M : 3 e l b a T - - - - - - - - 2 A ] 2 t a 2 t b [ 2 D 1 D ] 1 b b 1 b a [ 1 A - - - - - - - - ] t b 1 s t a [ s t C p t C ] 1 p t a 1 p b a [ p b C s b C 1 b B , ] t b 2 s t a [ s t C p t C ] 1 p t a 1 p b a [ p b C s b C 1 b B ] t b 1 s t a [ s t C p t C ] 1 p t a 1 p b a [ p b C ] 1 s b a b b [ , ] t b 2 s t a [ s t C p t C ] 1 p t a 1 p b a [ p b C ] 1 s b a b b [ ] t b 2 s t a [ s t C ] p t C 1 p b a [ p b C s b C ] 1 s b a b b [ ] t b s t C [ p t C ] 1 p t a 1 p b a [ p b C s b C ] 1 s b a b b [ 1 t B s t C p t C ] 1 p t a 1 p b a [ p b C s b C 1 b B , 2 t B s t C p t C ] 2 p t a 1 p b a [ p b C s b C 1 b B ] t b 1 s t a [ s t C p t C ] 1 p t a 1 p b a [ p b C 1 b B , ] t b 2 s t a [ s t C p t C ] 1 p t a 1 p b a [ p b C 1 b B ] t b s t C [ p t C ] 1 p t a 1 p b a [ p b C s b C 1 b B , ] t ] t b 2 s t a [ s t b 1 s t a [ s t C p t C C p t C ] 1 p t a 1 p b a [ p b C ] 1 p t a 1 p b a [ p b C s b C s b C ] 1 s b a b b [ ] 1 s b a b b [ s s a l C y g o l o h p r o M ¯C ¯C ¯C F - F ¯C ¯C ¯C F - F ¯C ¯C F - F ¯C ¯C F - C ¯C F - F ¯C ¯C ¯C ¯C F - F ¯C ¯C F - F ¯C F - F ¯C C - F ¯C ¯C F - F F - C ¯C F - F ¯C ¯C F - F F - F ¯C ¯C F - F ¯C ¯C ¯C F - F ¯C ¯C F - F ¯C F - F F - F ¯C ¯C F - F ¯C ¯C F - F X I X c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 Figure 11. Morphology class sample light curves, cf. Table 3, all to same scale (magnification 1 to 50, with dotted line at 10). Scale and ranges as in Fig. 10. The catalogue of binary microlensing light curves 15 (b) II F-F s = 1.0 u0 = 0.55 α = 0.05 (d) II C-C s = 1.0 u0 = 0 α = 0 (f) ¯C F-F III s = 1.0 u0 = 0.5 α = 0.7854 (h) III C-F ¯C s = 1.0 u0 = 0.015 α = 0.05 (j) III C-¯F-F s = 1.5 u0 = 0.125 α = 0.21 (a) ¯C I s = 1.0 u0 = 1 α = 1.55 (c) ¯C ¯C II s = 1.0 u0 = 0.73 α = 0.05 (e) II C-F s = 2.05 u0 = 0.6 α = 1.015 (g) III F-¯F-F s = 1.0 u0 = 0.58 α = 0.05 (i) ¯C ¯C ¯C III s = 1.0 u0 = 0.34 α = 1.55 c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 16 C. Liebig et al. Figure 12. Morphology class sample light curves. (Continued.) (a) ¯C C-F III s = 0.5 u0 = 0.205 α = 1.45 (c) IV F-F F-F s = 1.0 u0 = 0.25 α = 1.2 (e) IV F-¯F-¯F-F s = 1.0 u0 = 0.135 α = 1.55 (g) IV F-F ¯C ¯C s = 2.05 u0 = 0.04 α = 0.2 (i) IV F-F C-F s = 0.5 u0 = 0.07 α = 1.53 (b) IV ¯C F-F ¯C s = 1.0 u0 = 0.34 α = 0.7854 (d) IV F-¯F-F ¯C s = 0.85 u0 = 0.16 α = 1.2 (f) IV C-F F-F s = 2.05 u0 = 0.06 α = 0.07 (h) IV C-C C-C s = 2.05 u0 = 0 α = 0 (j) IV C-F ¯C ¯C s = 2.5 u0 = 0.055 α = 0.05 c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 The catalogue of binary microlensing light curves 17 Figure 13. Morphology class sample light curves. (Continued.) (b) V ¯C F-F F-F s = 1.0 u0 = 0.2 α = 1.5 (d) V ¯C F-¯F-F ¯C s = 2.05 u0 = 0.055 α = 0.148 (f) V F-F ¯C F-F s = 2.05 u0 = 0.005 α = 0.15 (h) V C-C C-F ¯C s = 2.05 u0 = 0.001 α = 0.0028 (j) V ¯C F-C F-F s = 2.5 u0 = 0.04 α = 0.046 (a) IV ¯C F-C ¯C s = 0.5 u0 = 0.009 α = 1.505 (c) V F-F F-¯F-F s = 1.0 u0 = 0.15 α = 1.55 (e) V ¯C F-F ¯C ¯C s = 2.05 u0 = 0.045 α = 0.15 (g) V ¯C F-F C-F s = 0.7 u0 = 0.04 α = 1.52 (i) V C-F F-F ¯C s = 2.05 u0 = 0.043 α = 0.048 c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 18 C. Liebig et al. Figure 14. Morphology class sample light curves. (Continued.) (a) V C-F ¯C F-F s = 2.5 u0 = 0.031 α = 0.027 (c) V ¯C ¯C F-C ¯C s = 0.5 u0 = 0.005 α = 1.53 (e) VI ¯C F-F F-F ¯C s = 2.5 u0 = 0.042 α = 0.002 (g) VI ¯C F-¯F-F F-F s = 2.05 u0 = 0.04 α = 0.113 (i) VI F-C C-C C-F s = 0.5 u0 = 0.001 α = 1.571 (b) V F-F ¯C ¯C ¯C s = 0.5 u0 = 0.04 α = 1.53 (d) VI F-F F-F F-F s = 1.0 u0 = 0.165 α = 1.55 (f) VI ¯C F-F ¯C F-F s = 2.5 u0 = 0.025 α = 0.05 (h) VI C-F ¯C F-F ¯C s = 2.5 u0 = 0.025 α = 0.022 (j) VI ¯C F-F ¯C C-F s = 0.5 u0 = 0.027 α = 1.555 c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 The catalogue of binary microlensing light curves 19 Figure 15. Morphology class sample light curves. (Continued.) (a) VI F-F F-F C-F s = 0.7 u0 = 0.028 α = 1.535 (c) VI C-C ¯C F-F ¯C s = 2.5 u0 = 0.01 α = 0.01 (e) VI ¯C ¯C F-F ¯C ¯C s = 0.5 u0 = 0.03 α = 1.571 (g) VI ¯C F-F ¯C ¯C ¯C s = 0.5 u0 = 0.03 α = 1.535 (b) VI F-F F-C F-F s = 0.65 u0 = 0.01 α = 1.535 (d) VI ¯C F-C F-F ¯C s = 2.5 u0 = 0.025 α = 0.028 (f) VI ¯C ¯C F-C ¯C ¯C s = 0.5 u0 = 0.003 α = 1.55 (h) VII F-F F-F ¯C F-F s = 0.65 u0 = 0.019 α = 1.54 (i) VII ¯C F-F ¯C ¯C F-F (j) VII F-F F-F ¯C C-F s = 0.65 u0 = 0.055 α = 1.507 s = 0.65 u0 = 0.021 α = 1.55 c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 20 C. Liebig et al. Figure 16. Morphology class sample light curves. (Continued.) (a) VII ¯C ¯C ¯C F-C ¯C ¯C (b) VII ¯C ¯C F-F ¯C ¯C ¯C s = 0.5 u0 = 0.001 α = 1.555 s = 0.5 u0 = 0.01 α = 1.55 (c) VII F-F ¯C C-C C-F (d) VII ¯C ¯C F-F ¯C C-F s = 0.5 u0 = 0.002 α = 1.571 s = 0.5 u0 = 0.019 α = 1.56 (e) VII ¯C F-C ¯C F-F ¯C (f) VII C-F ¯C ¯C F-F ¯C s = 2.5 u0 = 0.015 α = 0.017 s = 2.5 u0 = 0.015 α = 0.012 (g) VII ¯C F-F ¯C F-F ¯C (h) VII ¯C F-F F-F F-F s = 2.5 u0 = 0.01 α = 0.042 s = 2.05 u0 = 0.0443 α = 0.13 (i) VIII F-F ¯C F-F ¯C F-F (j) VIII F-F ¯C C-C ¯C F-F s = 0.65 u0 = 0.001 α = 1.54 s = 0.5 u0 = 0.001 α = 1.565 c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 The catalogue of binary microlensing light curves 21 Figure 17. Morphology class sample light curves. (Continued.) (a) VIII F-F ¯C F-C ¯C F-F (b) VIII F-F ¯C ¯C F-C C-F s = 0.65 u0 = 0.002 α = 1.562 s = 0.65 u0 = 0.002 α = 1.568 (c) VIII F-F ¯C F-F ¯C C-F (d) VIII F-F F-F ¯C ¯C F-F s = 0.65 u0 = 0.01 α = 1.56 s = 0.65 u0 = 0.02 α = 1.545 (e) VIII ¯C ¯C ¯C F-F ¯C ¯C ¯C (f) VIII ¯C ¯C F-F ¯C ¯C F-F s = 0.5 u0 = 0.027 α = 1.571 s = 0.5 u0 = 0.015 α = 1.555 (g) VIII ¯C ¯C ¯C F-F ¯C C-F (h) VIII ¯C F-F ¯C ¯C F-F ¯C s = 0.5 u0 = 0.015 α = 1.561 s = 2.5 u0 = 0.02 α = 0.002 (i) IX F-F ¯C ¯C F-C ¯C F-F s = 0.5 u0 = 0.002 α = 1.56 (j) IX F-F ¯C ¯C F-F ¯C C-F s = 0.5 u0 = 0.01 α = 1.565 c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000 22 C. Liebig et al. Figure 18. Morphology class sample light curves. (Continued.) (a) IX ¯C ¯C ¯C F-F ¯C ¯C F-F s = 0.5 u0 = 0.008 α = 1.556 (b) IX F-F ¯C F-F ¯C ¯C F-F s = 0.65 u0 = 0.011 α = 1.555 (c) X F-F ¯C ¯C F-F ¯C ¯C F-F s = 0.5 u0 = 0.015 α = 1.57 c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
1901.01935
2
1901
2019-01-30T12:52:35
Two Jovian planets around the giant star HD202696. A growing population of packed massive planetary pairs around massive stars?
[ "astro-ph.EP" ]
We present evidence for a new two-planet system around the giant star HD202696 (= HIP105056, BD+26 4118). The discovery is based on public HIRES radial velocity measurements taken at Keck Observatory between July 2007 and September 2014. We estimate a stellar mass of 1.91$^{+0.09}_{-0.14}M_\odot$ for HD202696, which is located close to the base of the red giant branch. A two-planet self-consistent dynamical modeling MCMC scheme of the radial velocity data followed by a long-term stability test suggests planetary orbital periods of $P_{\rm b}$ = 517.8$_{-3.9}^{+8.9}$ days and $P_{\rm c}$ = 946.6$_{-20.9}^{+20.7}$ days, eccentricities of $e_{\rm b}$ = 0.011$_{-0.011}^{+0.078}$ and $e_{\rm c}$ = 0.028$_{-0.012}^{+0.065}$ , and minimum dynamical masses of $m_{\rm b}$ = 2.00$_{-0.10}^{+0.22}$\,$M_{\mathrm{Jup}}$ and $m_{\rm c}$ = 1.86$_{-0.23}^{+0.18}$,$M_{\mathrm{Jup}}$, respectively. Our stable MCMC samples are consistent with orbital configurations predominantly in a mean period ratio of 11:6 and its close-by high order mean-motion commensurabilities with low eccentricities. For the majority of the stable configurations we find an aligned or anti-aligned apsidal libration (i.e.\ $\Delta\omega$ librating around 0$^\circ$ or 180$^\circ$), suggesting that the HD202696 system is likely dominated by secular perturbations near the high-order 11:6 mean-motion resonance. The HD202696 system is yet another Jovian mass pair around an intermediate mass star with a period ratio below the 2:1 mean motion resonance. Therefore, the HD202696 system is an important discovery, which may shed light on the primordial disk-planet properties needed for giant planets to break the strong 2:1 mean motion resonance and settle in more compact orbits.
astro-ph.EP
astro-ph
Draft version January 31, 2019 Preprint typeset using LATEX style AASTeX6 v. 1.0 9 1 0 2 n a J 0 3 . ] P E h p - o r t s a [ 2 v 5 3 9 1 0 . 1 0 9 1 : v i X r a A GROWING POPULATION OF PACKED MASSIVE PLANETARY PAIRS AROUND MASSIVE STARS? TWO JOVIAN PLANETS AROUND THE GIANT STAR HD 202696. Trifon Trifonov1, Stephan Stock2, Thomas Henning1, Sabine Reffert2, Martin Kurster1, Man Hoi Lee3,4, Bertram Bitsch1, R. Paul Butler5, Steven S. Vogt6 1 Max-Planck-Institut fur Astronomie, Konigstuhl 17, D-69117 Heidelberg, Germany 2 Landessternwarte, Zentrum fur Astronomie der Universitat Heidelberg, Konigstuhl 12, D-69117 Heidelberg, Germany 3 Department of Earth Sciences, The University of Hong Kong, Pokfulam Road, Hong Kong 4 Department of Physics, The University of Hong Kong, Pokfulam Road, Hong Kong 5 Department of Terrestrial Magnetism, Carnegie Institution for Science, Washington, DC 20015, USA and 6 UCO/Lick Observatory, Department of Astronomy and Astrophysics, University of California at Santa Cruz, Santa Cruz, CA 95064, USA ABSTRACT We present evidence for a new two-planet system around the giant star HD 202696 (= HIP 105056, BD +26 4118). The discovery is based on public HIRES radial velocity (RV) measurements taken at Keck Observatory between 2007 July and 2014 September. We estimate a stellar mass of 1.91+0.09−0.14M(cid:12) for HD 202696, which is located close to the base of the red giant branch. A two-planet self-consistent dynamical modeling MCMC scheme of the RV data followed by a long-term stability test suggests planetary orbital periods of Pb = 517.8+8.9−3.9 and Pc = 946.6+20.7−20.9 days, eccentricities of eb = 0.011+0.078 −0.011 and ec = 0.028+0.065 −0.012 , and minimum dynamical masses of mb = 2.00+0.22−0.10 MJup and mc = 1.86+0.18−0.23 MJup, respectively. Our stable MCMC samples are consistent with orbital configurations predomi- nantly in a mean period ratio of 11:6 and its close-by high-order mean-motion commensurabilities with low eccentricities. For the majority of the stable configurations, we find an aligned or anti-aligned apsidal libration (i.e. ∆ω librating around 0◦ or 180◦), suggesting that the HD 202696 system is likely dominated by secular perturbations near the high-order 11:6 mean-motion resonance. The HD 202696 system is yet another Jovian-mass pair around an intermediate-mass star with a period ratio below the 2:1 mean-motion resonance. Therefore, the HD 202696 system is an important discovery, that may shed light on the primordial disk-planet properties needed for giant planets to break the strong 2:1 mean motion resonance and settle in more compact orbits. Keywords: Techniques: radial velocities − Planets and satellites: detection, dynamical evolution and stability − (Stars:) planetary systems 1. INTRODUCTION By 2018 November, 647 known multiple-planet sys- tems were reported in the literature,1 144 of which were discovered using high-precision radial velocity (RV) measurements. The RV technique is very successful in determining the orbital architectures of multiple extra- solar planetary systems. In some exceptional cases, N - body modeling of precise Doppler data in resonant or [email protected] 1 http://exoplanet.eu near-resonant multiple systems can reveal the system's dynamical properties and constrain the planets' true dynamical masses and mutual inclinations (e.g. Rivera et al. 2010; Tan et al. 2013; Trifonov et al. 2014, 2018; Nelson et al. 2016, and references therein). Therefore, multiple-planet systems discovered with the Doppler method are fundamentally important in order to under- stand planet formation and evolution in general. After publication of the Keck HIRES (Vogt et al. 1994) velocity archive by Butler et al. (2017), it be- came apparent that the early K giant star HD 202696 most likely hosts another RV multi-planet system. The HIRES data set comprises 42 precise RVs of HD 202696 2 taken between 2007 July and 2014 September, which we find to be consistent with at least two planets in the Jovian-mass regime with periods of ∼520 days for the inner planet and ∼950 days for the outer planet. The HD 202696 system is remarkable because of the small orbital separation between the two massive plan- ets forming an orbital period ratio close to the 9:5 mean-motion resonance (MMR). Notable examples of small-separation pairs around intermediate-mass stars are 24 Sex, η Ceti and HD 47366 (between 9:5 MMR and 2:1 MMR; Johnson et al. 2011; Trifonov et al. 2014; Sato et al. 2016); HD 200964 and HD 5319 (4:3 MMR; Johnson et al. 2011; Giguere et al. 2015); and HD 33844 (5:3 MMR; Wittenmyer et al. 2016). These systems may indicate that more massive stars tend to form more massive planetary systems, a fraction of which pass the strong 2:1 MMR during the inward planet migration phase and settle at high-order resonant or near-resonant configurations with period ratios < 2:1. The dynam- ical characterization of the relatively compact system around HD 202696 may help to enhance our understand- ing of the formation of massive multiple-planet systems around intermediate-mass stars. This paper is organized as follows. In Section 2 we present our estimates of the stellar properties of HD 202696. Section 3 presents our Keplerian and N - body dynamical RV modeling analysis, which reveals the planets HD 202696 b and c. In Section 4 we present the long-term stability analysis of the HD 202696 system and its dynamical properties. In Section 5 we provide a brief discussion of the HD 202696 system in the con- text of current knowledge of compact multiple-planet systems around giant stars. Finally, Section 6 provides a summary of our results. 2. STELLAR PARAMETER ESTIMATES OF HD 202696 The giant star HD 202696 (= HIP 105056, BD +26 4118) is a bright V = 8.2 mag star of spectral class K0III/IV with B− V =1.003 mag (ESA 1997). The sec- ond Gaia data release (Gaia DR2; Gaia Collaboration et al. 2016, 2018) lists a parallax of π = 5.28 ± 0.04 mas, an effective temperature Teff = 4951.4 K, a radius estimate of 6.01 R(cid:12), and a mean G = 7.96 mag. Bailer- Jones et al. (2018) carried out a Bayesian inference using the DR2 parallax information with a prior on distance and estimated a distance of 188.5+1.6−1.6 pc. For the dynamical analysis of the HD 202696 system, we need to incorporate the stellar parameters of the host star, most importantly its stellar mass. This is achieved by applying a slightly modified version of the methodol- ogy and Bayesian inference scheme used in Stock et al. (2018), which was particularly optimized for subgiant and giant stars such as HD 202696 and is capable of Table 1. Stellar parameters of HD202696 and their 1σ un- certainties. Parameter Spectral type V [mag] B − V [mag] E(B − V ) [mag] AV [mag] π [mas] RVabsolute [km s−1] Radius [R(cid:12)] Teff [K] Distance [pc] From Bayesian inferenceα Mass [M(cid:12)] Radius [R(cid:12)] Luminosity [L(cid:12)] Age [Gyr] Teff [K] log g [cm · s−2] Evolutionary stage Evolutionary phaseβ From Spectraγ Teff [K] log g [cm · s−2] [Fe/H] v · sin(i) [km s−1] RVabsolute [km s−1] HD202696 reference K0III-IV 8.23 ± 0.01 1.003 ± 0.005 0.08 ± 0.04 0.25 ± 0.13 5.28 ± 0.04 -- 34.45 ± 0.22 6.01+0.33−0.19 4951+78−129 188.5+1.6−1.6 [1] [1] [1] [2] [2] [3] [3] [3] [3] [4] This paper This paper This paper 1.91+0.09−0.14 6.43+0.41−0.39 23.4+2.4−2.0 1.32+0.35−0.25 5040+71−85 3.11+0.07−0.07 RGB (P= 0.996) This paper 8.30+0.13−0.11 This paper This paper This paper This paper 4988 ± 57 3.24 ± 0.15 0.02 ± 0.04 2.0 ± 0.8 -- 33.17 ± 0.05 This paper This paper This paper This paper This paper Note -- α Following Stock et al. (2018), β A value of 8.0 marks the RG base while 9.0 marks the RGB bump (Bressan et al. 2012), γ from HIRES spectra using CERES (Brahm et al. 2017a) and ZASP E (Brahm et al. 2017b), [1] ESA (1997), [2] Gontcharov & Mosenkov (2018), [3] Gaia Collaboration et al. (2016, 2018) , [4] Bailer-Jones et al. (2018) determining the post-main-sequence evolutionary stage. The method was shown to provide robust estimates of stellar parameters such as mass M , radius R, age τ , effective temperature Teff , surface gravity log g, and lu- minosity L in agreement with other methods, such as asteroseismology, spectroscopy, and interferometry. The method uses the trigonometric parallax, a metallicity es- timate, and photometry in at least two different bands as input parameters. We decided to use the Hipparcos V and B photometry as in Stock et al. (2018) instead of the available high-precision Gaia DR2 photometry be- cause it was shown that results based on this choice of photometry are in good agreement with literature val- ues. While Gaia DR2 photometry is more precise, it has 3 Figure 1. Bayesian stellar parameters estimated using the Gaia DR2 parallax, HipparcosB and V photometry, and the spectroscopic metallicity. yet to be tested whether comparisons with stellar evo- lutionary models do not suffer from systematics, which is especially important regarding the small photomet- ric uncertainties. Stock et al. (2018) adopted the stellar models based on the PAdova and TRieste Stellar Evo- lution Code (PARSEC; Bressan et al. 2012) and used bolometric corrections by Worthey & Lee (2011). The Bayesian inference is normally carried out in the plane of the astrometric-HR diagram (Arenou & Luri 1999) which has a color as the abscissa and the astrometry- based luminosity (ABL) in a specific photometric band λ as the ordinate. The ABL is a quantity that is lin- ear in the trigonometric parallax, allowing for unbiased comparisons of stellar positions to evolutionary models if the parallax error dominates. However, HD 202696 is affected by a significant amount of extinction, which in addition is very uncertain. Gontcharov & Mosenkov (2018) estimated a reddening of E(B − V ) = 0.08 mag and an extinction of AV = 0.25 mag in the V band for HD 202696. They also stated that their reddening estimates are better than 0.04 mag, a value that we conservatively adopt as the formal reddening error for HD 202696. We used the python SED fitting tool by Robitaille et al. (2007) together with the model atmo- spheres of Castelli & Kurucz (2004) to fit the available broadband photometry of HD 202696. We derived an extinction of AV = 0.144+0.280 −0.012 mag, consistent with the value and uncertainties by Gontcharov & Mosenkov (2018). However, due to the sparse sampling of the model atmospheres, the errors of the SED fitting are probably underestimated and should be regarded with caution. For the analysis of the stellar parameters in Figure 2. Padova evolutionary tracks of metallicity Z = 0.015875 ([Fe/H]≈ 0.02) with various initial stellar masses from 1.6 (red line) to 2.2 (pink solid line) M(cid:12) in steps of 0.1 M(cid:12) in the CMD. The position of HD 202696 is high- lighted as the black point. this paper, we used the extinction by Gontcharov & Mosenkov (2018). As the uncertainty of the extinction is the dominating factor in the determination of the stel- lar parameters of HD 202696, it is justified to use the absolute magnitude of the star instead of the ABL for the determination of stellar parameters and include the uncertainties of the extinction. We use the Bayesian distance estimate by Bailer-Jones et al. (2018) for the determination of the absolute magnitude from the dis- tance modulus. This decreases biases that arise when determining the absolute magnitude from the trigono- metric parallax, which is why the ABL should normally be the preferred quantity. We need to determine the spectroscopic metallicity of HD 202696 in order to derive the stellar mass. We de- rive spectroscopic stellar parameters such as Teff , surface gravity log g, metallicity [Fe/H], and v · sin(i) by using the available HIRES spectral observations of HD 202696. As the spectra obtained at Keck Observatory have se- vere iodine (I2) line contamination imposed in order to measure RVs using the I2 method (Marcy & Butler 1992; Valenti et al. 1995; Butler et al. 1996), we use the single I2-free stellar template spectrum2 required for the RV modeling process (see Butler et al. 1996) to determine our spectroscopic parameters. We reduce the I2-free spectrum, with a signal-to-noise ratio of approximately 2 Downloaded from http://nexsci.caltech.edu/archives/ koa/ 1.41.61.82.02.2Mass [M€]0.00.20.40.60.81.05.05.56.06.57.07.58.0Radius [R€]0.00.20.40.60.81.04800500052005400Temperature [K]0.00.20.40.60.81.08.839.299.76log(age[yr])0.00.20.40.60.81.02.83.03.2log(g[cm/s]) 0.00.20.40.60.81.0202530Luminosity [L€]0.00.20.40.60.81.00:60:70:80:91:01:1B¡V[mag]0:51:01:52:02:5V[mag]1:6M¯1:7M¯1:8M¯1:9M¯2:0M¯2:1M¯2:2M¯HD202696 4 120, using the CERES pipeline (Brahm et al. 2017a) and then analyze it by using ZASPE (Brahm et al. 2017b). Table 1 provides our spectral stellar parameter estimates of HD 202696. Using the spectroscopic metallicity estimate, the B and V photometry of the Hipparcos catalog and the Bayesian distance estimate of Bailer-Jones et al. (2018), together with their uncertainties, we apply the slightly modified Bayesian inference method of Stock et al. (2018) and determine that HD 202696 is, with a proba- bility of 99.6%, an early red giant branch star, right at the beginning of its ascent. We neglect the small proba- bility of the star belonging to the horizontal branch. We estimate a stellar mass of M = 1.91+0.09−0.14 M(cid:12), a lumi- nosity of L = 23.4+2.0−2.4 L(cid:12), and an age of 1.32+0.35−0.25 Gyr. The derived effective temperature of Teff = 5040+71−85 K is within 1 σ of our spectroscopic estimate (4988 ± 57 K) and the estimate by Gaia DR2 (4951+78−129 K). The same is true for the surface gravity of log g = 3.11+0.07−0.07 dex when compared to our spectroscopic estimate (3.24 ± 0.15 dex) and the radius R = 6.43+0.41−0.39 R(cid:12) when com- pared to the Gaia DR2 estimate (6.01+0.33−0.19 R(cid:12)). Given the Bayesian stellar radius and the v · sin(i) = 2.0 ± 0.8 km s−1 estimated from the spectra, we calculate a stel- lar rotational period of Prot = 162.7 ± 65.9 days, which is typical for an early evolved star such as HD 202696. The good agreement between the Bayesian stellar pa- rameters and the other estimates strengthens our confi- dence in the adopted extinction of the star, as the ex- tinction is significant and our results rather sensitive to the precise extinction value. Neglecting extinction, the resulting mass would have been ∼ 1.4 M(cid:12), considerably smaller than before, and the stellar parameters would not have been consistent with constraints provided by spectroscopy and by Gaia DR2 data. However, the fact that extinction and reddening are consistent with zero at the 2 σ level leads to a relatively large uncertainty of the stellar mass, especially toward lower masses. The posterior distribution for each of the stellar parame- ters determined from the Bayesian inference is shown in Fig. 1, while Fig. 2 shows the Padova evolutionary tracks in the color-magnitude diagram (CMD) for various ini- tial stellar masses and the CMD position of HD 202696. The Bayesian stellar parameters, which were determined from the maximum of the distribution, along with their 1σ error estimates are summarized in Table 1. 3. ANALYSIS OF RV DATA SET 3.1. Period search The HIRES RV time series of HD 202696 consists of 42 precise Doppler measurements taken between 2007 July and 2014 September. With a median RV precision of 1.37 m s−1 and an end-to-end amplitude of 127 m s−1, Figure 3. The top panel shows the available precise RV data of HD 202696 measured with HIRES between 2007 July and 2014 September. The solid curve is the best two-planet Ke- plerian model fitted to the data, while the shaded area is composed of 1000 randomly chosen confident fits from an MCMC test constructed to study the parameter distribu- tion. The bottom panel shows the best-fit residuals. After removing the two Keplerian components of the RV signal with semi-amplitudes of ∼ 32 and 28 m s−1 the RV scatter has an rms of 8.3 m s−1, which is mostly consistent with the expected short-period stellar jitter for this target. Figure 4. Two Keplerian signals with periods of 521.0 days and 956.1 days, respectively, shown in a phase-folded repre- sentation. these data clearly show systematic motions, resembling two Keplerians caused by two massive orbiting planets. Figure 3 shows the precise RV data of HD 202696 to- gether with our best two-planet Keplerian model and its uncertainties, while Fig. 4 shows the two fairly well- sampled Keplerian signals phase-folded at their best-fit periods (see Section 3.2 for a quantitative analysis of the data). Note that the data uncertainties displayed in Figures 3 and 4 already contain the best-fit RV jitter estimate of ∼8 m s−1, which was automatically added in quadrature to the error budget during the fitting. This 0100200300400500phase[d]−80−60−40−20020406080RV[m/s]02004006008001000phase[d]−80−60−40−20020406080 RV jitter estimate is typical for early K giants such as HD 202696 and is likely due to solar-like oscillations, which have periods of less than a day and are thus not re- solved in the long-term RV time series. The scaling rela- tions from Kjeldsen & Bedding (2011) predict 3.7 m s−1 RV jitter for HD 202969, while the typical RV jitter value for a star with a B − V color of 1.003 mag in the Lick sample is 10 -- 20 m s−1 (see Fig. 3 in Trifonov et al. 2014). Since the stars in the Lick sample are, on aver- age, slightly more evolved than HD 202696, it might be reasonable to expect HD 202696 to have slightly smaller RV jitter than most other stars in this sample. It is worth noting that with the release of the HIRES database, Butler et al. (2017) reported the presence of at least three periods in the Keck data of HD 202696. They identified two significant signals with periods of 522.3 ± 6.4 and 946.0 ± 19.0 days, classified as plane- tary "candidates", and an additional period of 214.2 ± 4.8 days which required further confirmation. As an independent test, we derive the general- ized Lomb-Scargle periodogram (GLS; Zechmeister & Kurster 2009) of the HIRES RV data, as well as that of the S- and H-index activity indicators (Fig. 5) also provided by Butler et al. (2017). For reference, in all panels of Fig. 5 the range of possible stellar rotation fre- quencies of HD 202696 is indicated by the red shaded area, with the red dashed line at the most likely value −0.00177 day−1, see Section 2). The (1/Prot. = 0.00615+0.00418 horizontal dotted, dot-dashed, and dashed lines corre- spond to false-alarm probability (FAP) levels of 10%, 1%, and 0.1%, respectively. We use an FAP <0.1% as our significance threshold. The top panel of Fig. 5 re- veals a dominant signal around 540 days (blue dashed line) in the HIRES RVs, which is significant. The sec- ond panel of Fig. 5 shows the periodogram of the RV data after the ∼ 540 days signal has been subtracted; it indicates a second significant signal with a period of ∼ 970 days (green dashed line). The middle panel of Fig. 5 shows that no period with an FAP < 0.1% can be de- tected after the two most significant signals have been consecutively subtracted from the data. The ∼214 d signal previously reported by Butler et al. (2017) is also visible in the residuals of the data, but with 1% < FAP < 10%, which we consider insignificant. Moreover, the 214 day signal falls within the red shaded area of possible stellar rotation frequencies and thus is not considered further in our analysis. No significant periodicities can be identified in the two panels of Fig. 5, which show the periodograms of the H- and the S-index activity indicators, respectively. Additionally, we inspected the V -band measurements of HD 202696 from the All-Sky Automated Survey (ASAS; Pojmanski et al. 2005). We find only 48 high- quality ASAS V -band measurements (graded A and B 5 p = 0.1% p = 1% p = 10% Figure 5. The GLS periodograms of the available HIRES RV time series reveal two significant signals around ∼540 and ∼970 days, marked with blue and green dashed lines, respec- tively. The possible stellar rotation frequencies are within the red shaded area, the red line denoting the most likely stellar rotation frequency (see Sec. 2). No evident periodic- ity with an FAP < 0.1% is detected after the prewhitening (i.e. consecutive subtraction of each signal from the data.) These two signals do not have counterparts in the S- and H- index activity indicators of the HIRES data, nor do they fall into the range of possible stellar rotational frequencies. The mean ASAS V -band photometry time series of HD 202696 is also clean of significant periodicities at these periods. with σV = 0.037 mag.), mostly taken between 2003 May and 2007 October, i.e. prior to the HIRES Doppler ob- servations of HD 202696. Thus, the sparse ASAS pho- tometry is only relevant to check for evident intrinsic photometric variability, which could be associated with the RV signals. No significant periodicity is observed in any of the five available ASAS apertures. The bottom panel of Fig. 5 shows the GLS periodogram of the aver- aged ASAS photometry measurements, indicating that HD 202696 is most likely a photometrically stable star. The two significant signals at ∼ 540 and ∼ 970 days are consistent with those reported by Butler et al. (2017) 2468101214HIRESRVs2468101214o-cofthe540dsignal2468101214Powero-cofthe540dand970dsignals2468101214HIRESH-index2468101214HIRESS-index0.0050.0100.0150.020frequency[1/d]2468101214ASASV-bandphot.50.067.0100.0200.0500.0period[d] 6 as planet candidates. We find that these signals are not consistent with the range of possible stellar rotation fre- quencies of HD 202696. Also, the S- and H-index activ- ity indicators of the HIRES data and the ASAS pho- tometry do not show any significant periodic structure or correlation that could be associated with the RV sig- nals. Thus, we consider the ∼ 540 and ∼ 970 day signals as robust planetary detections. 3.2. Orbital parameter estimates In order to determine the orbital parameters of the HD 202696 b and c planetary companions, we adopt a maximum-likelihood estimator (MLE) scheme cou- pled with a Nelder-Mead algorithm (also known as a downhill simplex algorithm; see Nelder & Mead 1965; Press et al. 1992). In our scheme, we minimize the negative logarithm of the model's likelihood function (− lnL), while we simultaneously optimize the plane- tary RV semi-amplitudes Kb,c, periods Pb,c, eccentrici- ties eb,c, arguments of periastron ωb,c, mean anomalies Mb,c and the RV zero-point offset. Prior information on the period, phase, and amplitude for each of the two- planet candidates is taken from the GLS periodogram test (see Fig. 5), which provides good starting parame- ters for the MLE algorithm. All parameters are derived for a uniform reference epoch of 2,450,000 (BJD). Fol- lowing Baluev (2009), we also include the RV jitter3 as an additional parameter in the modeling. We apply two models to the RV data: a standard un- perturbed two-planet Keplerian model and a more ac- curate self-consistent N -body dynamical model, which models the gravitational interactions between the plan- ets by integrating the equations of motion using the Gragg-Bulirsch-Stoer integration method (Press et al. 1992). We cross-check the obtained best-fit results for consistency with the Systemic Console 2.2 package (Meschiari et al. 2009) and find excellent agreement be- tween our MLE code4 and Systemic. For parameter distribution analysis and uncertainty estimates, we couple our MLE fitting algorithm with a Markov chain Monte Carlo (MCMC) sampling scheme using the emcee sampler (Foreman-Mackey et al. 2013). For all parameters, we adopt flat priors (i.e. equal prob- ability of occurrence) and we run emcee from the best fit obtained by the MLE. We select the 68.3% confidence levels of the posterior MCMC parameter distribution as 1σ parameter uncertainties. 3 The RV jitter is most likely due to a combination of astro- physical stellar noise and possible instrument and data reduction systematics. 4 the source code and GUI of our tools can be found in https://github.com/3fon3fonov/trifon (Trifonov et al. 2018 in preparation) Table 2. HD 202696 system. Best Keplerian and dynamical fits for the Two-planet Keplerian Model Parameter Semi-amplitude K [m s−1] Period P [days] Eccentricity e Arg. of periastron ω [deg] Mean anomaly M0 [deg] Mean longitude λ [deg] Semi-major axis a [au] Minimum mass m sin i [MJup] RV offset [m s−1] RV jitter [m s−1] − lnL HD 202696 b HD 202696 c 28.1+1.1−5.3 956.1+22.4−29.9 0.261+0.060 −0.202 129.1+142.7 −33.6 152.0+112.8 −74.3 281.1+50.5−58.9 2.357+0.037 −0.049 2.028+0.073 −0.354 31.7+4.8−2.1 521.0+6.2−7.3 0.056+0.066 −0.040 259.2+106.0 −89.2 69.7+85.8−115.8 328.8+50.4−44.9 1.573+0.012 −0.015 1.930+0.285 −0.132 -- 15.32+2.59−1.15 8.20+3.07−0.46 148.799 Two-planet Dynamical Model Parameter Semi-amplitude K [m s−1] Period P [days] Eccentricity e Arg. of periastron ω [deg] Mean anomaly M0 [deg] Mean longitude λ [deg] Semi-major axis a [au] Min. dyn. mass m [MJup] Inclination i [deg] Node ∆Ω [deg] RV offset [m s−1] RV jitter [m s−1] − lnL HD 202696 b HD 202696 c 26.6+1.1−3.5 958.0+2.2−26.4 0.265+0.005 −0.229 140.0+88.7−44.9 155.5+70.41 −52.2 259.5+1.0−68.5 2.360+0.003 −0.040 1.917+0.116 −0.212 32.5+3.1−1.6 528.2+1.8−12.5 0.026+0.100 −0.006 289.7+35.1−91.8 70.2+67.9−40.8 359.9+8.2−63.0 1.587+0.004 −0.025 1.991+0.174 −0.011 90.0 (fixed) 90.0 (fixed) 0.0 (fixed) -- 15.12+1.95−1.44 8.31+3.08−0.47 148.973 Note -- The parameters are valid for BJD = 2,450,000 and are derived using our stellar mass estimate of M = 1.91 M(cid:12). The uncertainties of ab,c, mb,c sin i (Keplerian model), and mb,c (dynamical model) do not take into account the small uncertainty of the stellar mass. The best-fit parameters of our two-planet Keplerian and dynamical models, together with their 1σ MCMC uncertainties, are tabulated in Table 2. We first fit the Keplerian model and as a next step, we adopt the best two-planet Keplerian parameters as an input for the more sophisticated dynamical model. For consistency with the unperturbed Keplerian frame and in order to work with minimum dynamical masses, we assume an edge-on and coplanar configuration for the HD 202696 system (i.e. ib,c = 90◦ and ∆Ω = 0◦). We find that the models are practically equivalent, suggesting a pair of planets with equal masses of the order of ∼2.0 MJup and a period ratio of Prat. ≈ 1.83, near the 11:6 MMR (for details, see Table 2). Both the Keplerian and dynamical fits reveal a mod- erate best-fit value of ec for the eccentricity of the outer planet but with a large uncertainty toward nearly cir- cular orbits. This indicates that ec is likely poorly con- strained, and perhaps this moderate eccentricity value is a result of overfitting. Indeed, a two-planet Keplerian model with circular orbits (i.e. eb,c, ωb,c = 0) has − lnL = 150.286, meaning that the difference between the cir- cular and nonciruclar Keplerian models is only ∆ lnL ≈ 1.5, which is insignificant5. Thus, we conclude that neither the Keplerian nor the dynamical model applied to the current RV data can place tight constraints on the orbital eccentricities. The inclusion of the planetary inclinations ib,c and ∆Ω (i.e. ∆i) in the dynamical modeling was also not justified. We find that all dynamical fits with coplanar inclinations in the range of 5◦ < ib,c < 90◦ are prac- tically indistinguishable from the coplanar and edge-on configuration (ib,c = 90◦ and ∆Ω = 0◦). Fits with ib,c < 5◦ are of poorer quality due to the increased dynamical masses of the companions and the strong gravitational interactions, which are not consistent with the data any- more. Therefore, any further attempt to constrain the planetary inclinations as coplanar or mutually inclined leads to inconclusive results. Overall, our orbital parameter estimates show that taking the planetary gravitational interactions into ac- count in the fitting does not yield a significant improve- ment over the Keplerian model. Yet we decide to rely on the dynamical modeling, since this scheme provides physical constraints on the derived orbital parameters and naturally penalizes highly unstable (unrealistic) or- bital solutions. 4. DYNAMICAL CHARACTERIZATION AND STABILITY 4.1. Numerical setup The long-term stability and dynamical properties of the HD 202696 system are tested using the SyMBA N - body symplectic integrator (Duncan et al. 1998), which was modified to work in Jacobi coordinates (e.g., Lee & Peale 2003). We integrate each MCMC sample for a maximum of 1 Myr with a time step of 4 days. In case of planet-planet close encounters, however, SyMBA au- 5 Assuming that the 2∆ ln L follows a χ2 distribution for nested models with k = 4 degrees of freedom (e.g. see Baluev 2009). Alternatively, the relative probability R = e∆ ln L between two competing models constructed from the same parameters (relevant for the MCMC test) requires ∆ ln L > 5 to claim significance (see Anglada-Escud´e et al. 2016) 7 tomatically reduces the time step to ensure an accurate simulation with high orbital resolution. SyMBA also checks for planet-planet or planet-star collisions or plan- etary ejections and interrupts the integration if they oc- cur. A planet is considered lost and the system unstable if, at any time: (i) the mutual planet-planet separation is below the sum of their physical radii (assuming Jupiter mean density), i.e. planets undergo collision; (ii) the star-planet separation exceeds two times the initial semi- major axis of the outermost planet (rmax > 2 × ac in.), which we define as planetary ejection; or (iii) the star- planet separation is below the physical stellar radius (R ≈ 0.03 au), which we consider as a collision with the star. All these events are associated with large plane- tary eccentricities leading to crossing orbits, close plan- etary encounters, rapid exchange of energy and angular momentum, and eventually instability. Therefore, these somewhat arbitrary stability criteria are efficient to de- tect unstable configurations and save CPU time. 4.2. Stable orbital space Figure 6 shows the posterior MCMC distribution of the fitted parameters with a dynamical modeling scheme whose orbital configuration is edge-on, coplanar, and stable for at least 1 Myr. The histogram panels on the top in Figure 6 provide a comparison between the prob- ability density distribution of the overall MCMC sam- ples (black) and the stable samples (red) for each fitted parameter. The two-dimensional parameter distribution panels represent all possible parameter correlations with respect to the best dynamical fit from Table 2, whose po- sition is marked with blue lines. The black 2D contours are constructed from the overall MCMC samples (gray) and indicate the 68.3%, 95.5%, and 99.7% confidence levels (i.e. 1σ, 2σ and 3σ). For clarity, in Fig. 6 the stable samples are overplotted in red. We find that ∼46% of the MCMC samples are stable. The histograms in Figure 6 clearly show that the distri- butions of all MCMC samples and their stable counter- parts are in mutual agreement, peaking near the position of the best dynamical fit. The main exception is the dis- tribution of ec, which is bimodal with a stronger peak suggesting nearly circular orbits and a smaller peak at somewhat moderate eccentricities from where the best dynamical fit originates. Here Pb and ωb also show bi- modal distributions, which are related to the bimodal distribution of ec. As we discussed in Section 3.2, how- ever, the best-fit eccentricities are statistically insignifi- cant, and a simpler two-planet model with circular orbits is equally good in terms of − lnL. This indicates that the best-fit solution is likely a global − lnL minimum, which lies in a smaller and isolated orbital parameter space with moderate ec. This minimum and its sur- roundings, however, are highly unstable and therefore 8 Figure 6. The MCMC distribution of orbital parameters consistent with the HIRES RV data of HD 202696 assuming an edge-on coplanar two-planet system fitted with a self-consistent dynamical model. The position of the best dynamical fit from Table 2 is marked with blue lines. The black contours on the 2D panels represent the 1σ, 2σ and 3σ confidence levels of the overall MCMC samples, while on the 1D histograms, the dashed lines mark the 16th and the 84th percentiles of the the overall MCMC samples. The samples that are stable for at least 1 Myr are red and represent ∼46% of all samples. The stable samples are mostly within the 1σ confidence region from the best fit. See text for details. unlikely to represent the HD 202696 system configura- tion. Thus, running an MCMC test reveals a larger, sta- ble, and statistically confident orbital parameter space toward eb,c → 0, where the stable fraction of the sample distribution is ∼75%. From the long-term dynamical analysis of the posterior MCMC samples, we can place a boundary at planetary eccentricities eb,c ≤ 0.1, while larger eccentricities are very unlikely. The overall poste- rior distribution of ωb,c and Mb,c are consistent with the best-fit estimate, but the stability test cannot constrain these parameters. Because of the low eccentricities of the stable solutions, these parameters are ambiguous and can be found anywhere between 0◦ and 360◦. How- ever, the orbital mean longitudes λb,c = ωb,c + Mb,c are 510520530540550Pb[day]0.080.160.240.320.40eb160240320400ωb[deg]080160240Mb[deg]20253035Kc[m/s]90095010001050Pc[day]0.080.160.240.320.40ec80160240320ωc[deg]2530354045Kb[m/s]80160240320Mc[deg]510520530540550Pb[day]0.080.160.240.320.40eb160240320400ωb[deg]080160240Mb[deg]20253035Kc[m/s]90095010001050Pc[day]0.080.160.240.320.40ec80160240320ωc[deg]80160240320Mc[deg] 9 Table 3. Mode of the dynamically stable MCMC distribu- tion for the HD 202696 system (peak of the red histograms from Fig. 6) Adopted Two-planet Coplanar Edge-on Configuration Based upon Stability Parameter Semi-amplitude K [m s−1] Period P [days] Eccentricity e Arg. of periastron ω [deg] Mean anomaly M0 [deg] Mean longitude λ [deg] Semi-major axis a [au] Dyn. mass m [MJup] Inclination i [deg] Node ∆Ω [deg] RV offset [m s−1] RV jitter [m s−1] HD 202696 b HD 202696 c 25.3+2.1−3.3 946.6+20.7−20.9 0.028+0.065 −0.012 136.8+143.0 −62.7 180.0+98.9−108.9 262.8+37.8−53.3 2.342+0.034 −0.035 1.864+0.177 −0.227 34.1+2.3−2.9 517.8+8.9−3.9 0.011+0.078 −0.011 201.6+97.8−98.5 79.2+198.6 −28.1 298.8+70.0−14.6 1.566+0.016 −0.007 1.996+0.220 −0.100 90.0 (fixed) 90.0 (fixed) 0.0 (fixed) -- 14.42+1.70−2.22 9.36+2.03−0.57 Note -- Same comments as in Table 2 ∆ lnL of the dynamical models in the range of i = 90◦ -- 15◦ shows that these models are statistically equivalent, but the fit quality rapidly deteriorates for i < 15◦ (not shown in Fig. 7). This is likely due to the significantly increased planetary masses, which may alter the system stability even at lower eccentricities. Inclination i = 20◦ seems to be a stability boundary of the coplanar case, leaving the companion masses in the planetary regime with a maximum of about three times the minimum planetary masses of the coplanar edge-on case. However, since orbital inclinations near i = 90◦ are statistically more likely, we limit our dynamical analysis of the HD 202696 system to the coplanar and edge-on configuration. 4.4. Dynamical properties of the stable MCMC samples Overall dynamical analysis of the stable samples re- veals that the mean period ratio over 1 Myr of inte- grations is Prat. = 1.83+0.04−0.05, which is consistent with the period ratio from which the system is initially inte- grated, Prat.in. = 1.82+0.03−0.03. This implies that the stable MCMC orbital configurations remain regular and well separated at any given time of the stability test. From these stable samples, ∼29% are found with an aligned apsidal libration (∆ω ∼ 0◦), ∼34% in an anti-aligned apsidal libration (∆ω ∼ 180◦; similar to the orbital evo- lution of the long-term stable configuration presented in Table 3), and ∼37% in apsidal circulation (∆ω in the range 0◦ -- 360◦), respectively. In fact, the latter case usually involves eb decreasing from near its maximum Figure 7. Coplanar dynamical fits with nearly circular orbits as a function of inclination. The stability boundary is at i = 20◦, which sets a planetary mass limit of approxi- mately three times the minimum planetary mass of the edge- on coplanar case (e.g. see Table 3). rather well constrained to ∼ ± 50◦ (see Table 2). Given the stability constraints from our MCMC test, the most realistic configuration of the HD 202696 sys- tem is represented by the parameter distribution of the stable fits. In Table 3 we present our final orbital esti- mates of the two-planet system assuming coplanar and edge-on (ib,c = 90◦, ∆Ω = 0◦) configurations. These parameters are adopted from the maximum value of the stable parameter probability density function, while the uncertainties are the 0.16 and 0.84 quantiles of the dis- tribution. Our new orbital estimate is consistent with circular orbits eb = 0.011+0.078 −0.012 but otherwise consistent with the best two-planet dynamical fit shown in Table 2. −0.011 and ec = 0.029+0.065 Integrating this configuration for 100 Myr reveals a stable planetary system with a mean period ratio Prat. ≈ 1.83 and orbital eccentricities oscillating in opposite phase in the range from 0 to 0.05 for eb and 0 to 0.04 for ec. This configuration librates mostly in an anti-aligned geometry with a secular apsidal angle ∆ω = ωb − ωc ∼ 180◦ exhibiting a large libration amplitude of ± 80◦. The long-term stability of the orbital estimate shown in Table 3 boosts the confidence that the most likely orbital configuration of the HD 202696 system has been identified. 4.3. Stability constraints on the planetary masses While the dynamical modeling of the HIRES Doppler measurements of HD 202696 was unable to qualitatively constrain the masses of HD 202696 b and c, we also use the long-term stability of fits with forced inclination as an efficient way to limit the range of the possible plan- etary masses further. Figure 7 shows the results from this test. We inspect the range of inclinations from i = 90◦ to 5◦. To ensure enhanced stability, we adopt the planetary eccentricity estimates from Table 3 which for each adopted i remained fixed in the fit. The small 15.030.045.060.075.090.0coplanarinclination[deg]0.00.10.20.30.40.50.6∆lnLstableunstablestabilityborderatib,c∼20deg. 10 ∆ω ∼ 0◦ ∆ω ∼ 180◦ Figure 8. Distribution of the mean period ratio Prat. and the amplitude of the apsidal argument ∆ω for all dynamical fits that are stable up to 1 Myr (see Fig. 6) and consistent with libration of ∆ω. The left panel shows those stable samples whose dynamical properties are consistent with aligned apsidal libration (∆ω ∼ 0◦), while the right panel shows the samples with anti- aligned apsidal libration (∆ω ∼ 180◦). Both geometries are consistent with Prat. ∼ 1.83 (near 11:6 MMR) and have similar ∆ω libration amplitudes. Sample concentrations near the 9:5 and 13:7 MMRs (slightly enhanced) are also within the 1 σ confidence levels of the stable posterior distribution. A small and insignificant fraction of the samples are found in the 2:1 MMR. value to near zero and ec increasing from near zero to near its maximum value at ∆ω ≈ 90◦, and vice versa at ∆ω ≈ 270◦, with episodes of rapid circulation when eb or ec is nearly zero (see right panel of Fig. 9). This type of evolution is consistent with a phase space dominated by large libration islands about ∆ω = 0◦ and 180◦ and having only a narrow region of circulation (see, e.g., Lee & Peale 2003). Figure 8 shows the distribution of the mean period ratio Prat. and the libration amplitudes of ∆ω for the stable dynamical samples with aligned apsidal libration ∆ω ∼ 0◦ (left panel) and with anti-aligned apsidal li- bration ∆ω ∼ 180◦ (right panel). From Fig. 8 it is clear that both alignment and anti-alignment have similar li- bration amplitudes of ∆ω peaking around 70◦ and with a sharp border at 90◦, beyond which the system is in- volved in apsidal circulation. An interesting pileup in the posterior distribution of Prat. can be seen around 1.80, 1.83, and 1.86, which correspond to the 9:5, 11:6, and 13:7 MMRs. We inspected these possible high-order MMRs for librating resonance angles,6 but we did not detect any. Despite the apsidal libration of the ∆ω in ∼60% of the configurations, which is suggestive of reso- nant behavior, the system seems to be out of any MMR. In the ∆ω ∼ 180◦ case, a small number of system con- figurations are consistent with a 2:1 period ratio, but these samples have statistically poorer − lnL and are thus unlikely to represent the HD 202696 system. Given the large fraction of nonresonant stable con- figurations, we concluded that the system HD 202696 is most likely dominated by secular interactions at low eccentricities. Figure 9 shows the orbital evolution of three confident stable configurations that have a mean period ratio of Prat. ∼ 1.83, exhibiting ∆ω ∼ 0◦ and ∆ω ∼ 180◦, and ∆ω circulating, respectively. The var- ious stable configurations show libration amplitudes in their eccentricities between 0 and 0.15 but with mean values consistent with the posterior eccentricity distri- bution. The secular time scales of these oscillations strongly depend on the initial orbital geometries and planetary masses, ranging mostly between ∼200 and ∼1000 yr. 6 For the m:n MMR, the resonance angles are defined as θk=1,m−n = nλb + (n − k)b − mλc − (1 − k)c, where  and λ are the planetary longitude of periatron and the mean longitude, respectively. At least one resonance angle θk must librate to claim a MMR configuration. 5. DISCUSSION 1.71.81.92.0Prat.306090120∆ωmax.-∆ωmin.306090120∆ωmax.-∆ωmin.1.71.81.92.0Prat.306090120∆ωmax.-∆ωmin.306090120∆ωmax.-∆ωmin. 11 Figure 9. Reconstruction of three possible outcomes of the orbital evolution of the HD 202696 system with Prat. ∼ 1.83, randomly chosen from the stable posterior distribution. Left panel: orbital evolution with ∆ω librating around 0◦. Middle panel: a case where ∆ω librates around 180◦. Right panel: a case where ∆ω circulates between 0◦ and 360◦. Note the large amplitude of Prat. in all three cases. 5.1. Tight Jovian Pairs around Evolved Stars The HD 202696 system is another example of a grow- ing number of multiplanetary systems around evolved intermediate-mass stars with semi-major axes close to the critical stability boundary. Including HD 202696, there are now seven tightly packed planetary systems with period ratios smaller than or about 2:1. Earlier examples are the 24 Sex and HD 200964 sys- tems (Johnson et al. 2011). A 1.5M(cid:12) subgiant 24 Sex is orbited by two Jovian planets with minimum dynamical masses of mb ∼ 2, and mc ∼ 0.9MJup, and periods of Pb ≈ 450 and Pc ≈ 883 days. The best-fit period ratio for the 24 Sex system is slightly below 2:1, but the planets are almost certainly involved in a 2:1 MMR (Johnson et al. 2011; Wittenmyer et al. 2012). The HD 200964 system is similar to 24 Sex in terms of stellar mass and minimum planetary masses, but it has a much more compact orbital configuration, with periods of Pb ≈ 610 and Pc ≈ 830 days. The HD 200964 system is only stable if the orbits are in a 4:3 MMR (Wittenmyer et al. 2012; Tadeu dos Santos et al. 2015). The 1.5M(cid:12) subgiant star HD 5319 is another 4:3 MMR candidate host (Robinson et al. 2007; Giguere et al. 2015). With minimum planetary masses of mb sin i ∼ 1.8MJup, mc sin i ∼ 1.2MJup, and periods of Pb ≈ 640 d and Pc ≈ 890 d, respectively, this system is dy- namically very challenging. Giguere et al. (2015) showed that, although most of the possible configurations of HD 5319 are unstable, stable solutions at the exact 4:3 MMR do exist. Therefore, the 4:3 MMR seems like the most reasonable explanation of the coherent Doppler sig- nal of HD 5319. A 1.7M(cid:12) slightly evolved star, HD 33844 was reported by Wittenmyer et al. (2016) to host two Jovian-mass planets with minimum dynamical masses mb ∼ 1.9, and mc ∼ 1.7MJup, and periods of Pb ≈ 550, Pc ≈ 920 days, respectively. These authors concluded that the plane- tary pair is most likely trapped in the 5:3 MMR, but stable solutions consistent with 8:5 and 12:7 MMR were also found within the 1σ uncertainty in the semi-major axes. A 1.8M(cid:12) giant, HD 47366 has two nearly equal-mass planetary companions (Sato et al. 2016). With mb sin i ∼ 1.8, mc sin i ∼ 1.9MJup and periods of Pb ≈ 363 and Pc ≈ 685 days, the system is dynamically challenging. Within the uncertainties of the orbital parameters given, the initial period ratio is close to 15:8, where the sys- tem could be stable if the planets are on circular orbits. Sato et al. (2016) also proposed that HD 47366 could be locked in a retrograde 2:1 MMR. Tightly packed retro- grade configurations are, however, difficult to reconcile with current planet formation scenarios. Marshall et al. (2018) showed that alternatives to these configurations for HD 47366 exist, including stable prograde solutions in the 2:1 MMR. The η Ceti system is composed of a giant star orbited by a planet pair with minimum dynamical masses of mb ∼ 2, mc ∼ 3MJup, and nominal best-fit periods of Pb ≈ 400 and Pc ≈ 750 days, respectively. Trifonov et al. (2014) found two possible stability regimes for η Ceti: ei- ther it is a strongly interacting two-planet system locked in an anti-aligned 2:1 MMR, or, with lower probability, it could be a low-eccentricity two-planet system domi- nated by secular interactions close to a high-order MMR like 9:5, 11:6, 13:7, or 15:8, resembling to some extent the HD 202696 system. Figure 10 shows all known RV multiple-planet systems around intermediate-mass stars with masses larger than 12 Figure 10. Period ratio of known multiplanetary systems discovered via the Doppler method and orbiting stars with estimated masses larger than 1.3 M(cid:12). Each planetary com- panion is plotted at the published best-fit period ratio for the system, with a symbol that is color-coded and scaled in size with its reported mass. Six dynamically challenging systems are known to orbit at or below the 2:1 MMR. A new mem- ber of this sample with a period ratio of 11:6, HD 202696 is indicated in the plots with a red star symbol. See text for details. 1.3M(cid:12)7. The stellar host mass is shown as a function of the logarithm of the system's period ratio. A total of 21 planetary8 systems are shown in Fig. 10. The seven planetary systems found between the 2:1 and 4:3 first-order MMRs (and briefly described above) repre- sent 33% of all known massive multiple-planet systems around intermediate-mass stars. This fraction is rather large given the broad range of the observed planetary period ratios. The planet pair occurrence rate at low period ratios, of course, could be a result of a selection effect, since these systems are likely to be discovered first by precise Doppler surveys. However, a more intrigu- ing possibility is that the increasing number of these tight planetary pairs indicates the existence of a mas- sive multiple-planet system population, some of which have successfully broken the 2:1 MMR during the planet migration phase. It should be noted that a strongly interacting 2:1 MMR system could, in principle, exhibit prominent os- cillations of their semi-major axis (and period ratio). 7 Taken from http://exoplanet.eu, regardless of their spectral type and evolutionary stage. 8 As can be seen from Fig. 10 one of the systems is in fact a brown dwarf pair. This is the ν Oph system locked in a 6:1 MMR (Quirrenbach et al. 2011; Sato et al. 2012, Quirrenbach et al. A&A in press.) The secular oscillation time scales are usually longer than the temporal baseline of the Doppler measurements and therefore normally only marginally detected in the best-fit modeling. An example is the η Ceti system, where best-fit dynamical modeling of the osculating or- bits suggests Prat. ∼ 1.86, but the stable solutions of the system have Prat. = 2:1, with an MMR libration ampli- tude of ∆Prat. ∼ 0.2 and time scales of ∼500 yr (Trifonov et al. 2014). In this context, the 24 Sex, HD 47366, and even the HD 202696 system could in fact be true 2:1 MMR systems with oscillating period ratios, observed slightly below the 2:1 period ratio at the present epoch. More precise Doppler data for these systems taken in the future, with an extended temporal baseline, could reveal the true orbital configurations of these systems; we encourage such extended RV campaigns. Still, the HD 33844 system is likely stable at or near the 5:3 MMR, while the HD 5319 and the HD 200964 planetary systems can only be stable if they are involved in the 4:3 MMR. The latter systems are dynamically not consistent with the 2:1 MMR and thus indicate that planet pairs could break the 2:1 MMR and settle at other stable orbits in lower-order MMRs. 5.2. Planetary migration below the 2:1 MMR The HD 202696 system, like the majority of the known Jovian-mass planetary systems, is currently located within the so-called "ice-line", where it is unlikely to form gas giants. This can be easily seen from Fig. 11, which shows the planetary systems from Fig. 10 vs. their planetary semi-major axes. The dashed blue line de- notes the approximate ice-line radius, beyond which it is commonly accepted that planetary embryos are able to accumulate icy material and grow massive enough to start gas accretion from the protoplanetary disk to eventually become Jovian-mass planets. While accreting disk material, massive planets undergo inward migration towards the stellar host before the disk dissipates, and planets end up at their observed semi-major axes. We note that the ice line plotted in Fig. 11 is calculated for the protoplanetary phase assuming only stellar irradia- tion, where the disk temperature would be Tirr = 150 K (Ida et al. 2016), but adopting a stellar mass-luminosity relation of the form L = M 3.5, which is reasonable for main-sequence stars. We are aware that this is a crude approximation, but it is sufficient to demonstrate the generally "warm orbits" of the discovered Jovian-mass planets orbiting intermediate stars. Fig. 11, in connection with current planet formation theories, suggests that the massive planet pairs have formed further out than observed today and have mi- grated inward together. If the migration is convergent, there is a high probability of planets being trapped in an orbital resonance, with a final stop at the strong 2:1 0.00.51.01.52.02.53.03.5ln(Prat.)1.01.52.02.53.03.5M(cid:12)4:32:13:14:15:16:1159131823Mjup 13 same time-scale, and it is very hard for the outer planet to catch up with the inner planet, let alone to break free from a resonance trapping via planet-disk interactions. However, if the outer planet is less massive and does not open a full gap, it can undergo type III migration (Masset & Papaloizou 2003). This migration mode is even faster than type I migration and can easily allow outer planets to catch up with inner, more massive plan- ets and be trapped in resonance (Masset & Snellgrove 2001). This situation has been studied in the framework of the Grand-Tack scenario for Jupiter and Saturn (Walsh et al. 2011; Pierens et al. 2014). In particular, Pierens et al. (2014) found that the inward-migrating Saturn could, under certain disk conditions, have broken free from the 2:1 resonance and migrated closer to Jupiter, where trapping could have occurred, e.g. in the 3:2 res- onance. While the planets are still embedded in the gas disk, they will continue to accrete gas and grow. The outer planet can feed from a larger gas reservoir than the inner one because it has direct access to the disk's gas supply with the full disk accretion rate. The inner planet can only accrete the gas that passed the outer planet, which is typically only up to 20% of the disk's accretion rate (Lubow & D'Angelo 2006). Eventually, the outer planet can then outgrow the inner planet if the disk lifetime is long enough. During their final growth, the dynami- cal interactions between the giant planets change, which could lead to small orbital disturbances breaking the fi- nal resonance. A possible escape from the 2:1 MMR could also be a resonance overstability, as in the context of Goldreich & Schlichting (2014). In this case, convergent planetary migration with strongly damped eccentricities may only lead to a temporal capture at the 2:1 MMR. However, these scenarios require further investigation, which is beyond the scope of this paper. 6. SUMMARY The star HD 202696 is an early K0 giant (M = 1.91+0.09−0.14 M(cid:12), log g = 3.24 ± 0.15 cm · s−2, [Fe/H] = 0.02 ± 0.04 dex), for which HIRES high-precision RV measurements were taken between 2007 and 2014. The RV data indicate the presence of at least two strong pe- riodicities near 520 and 940 days, with no evidence of stellar activity near these periods. Therefore, the most plausible explanation of the observed RVs is stellar re- flex motion induced by a pair of Jovian-mass planets, each with a minimum mass of ∼2 MJup. Dynamical modeling of the RV data is particularly important for the HD 202696 system, since planet- planet perturbations in this relatively compact system of massive planets cannot be neglected. For the or- Figure 11. Same as in Figure 10 but as a function of the planetary semi-major axes. The dashed blue curve shows the approximate position of the pre-main-sequence "ice line" for stellar masses between 1 and 3.5 M(cid:12). Most of the discovered Jovian-mass multiple-planet systems around intermediate- mass stars are within the ice line, suggesting that these pairs most likely migrated inward together during the disk phase. See text for details. MMR (e.g. Lee & Peale 2002; Beaug´e et al. 2003). Per- haps due to not yet well understood dynamical processes between the disk and the planets, pairs of planets may have broken the 2:1 MMR and migrated further in. Andr´e & Papaloizou (2016) studied systems similar to 24 Sex and η Ceti and were able to reproduce the 2:1 MMR in these systems. However, in their simulations, there were also cases in which the planets passed through the 2:1 MMR, which made them undergo capture into the destabilizing 5:3 resonance. Andr´e & Papaloizou (2016) concluded that finding systems in the 5:3 MMR is unlikely. In this context, the 4:3 MMR systems are also very challenging from the formation perspective. Rein et al. (2012) failed to reproduce the 4:3 MMR of the HD 200964 system using the standard formation sce- narios, such as convergent migration, scattering, and in situ formation. On the other hand, Tadeu dos Santos et al. (2015) were able to closely reproduce the 4:3 res- onant dynamics of HD 200964 by including interactions between type I and type II migration, planetary growth, and stellar evolution. As pointed out by Tadeu dos San- tos et al. (2015), however, the 4:3 MMRs capture of the HD 200964 system occurred only for a very narrow range of initial conditions. In general, massive planets that have reached the gap opening mass migrate via type II migration, which fol- lows the viscous disk evolution (Lin & Papaloizou 1986). In this case, a two-planet pair would migrate on the 0246810a[au]1.01.52.02.53.03.5M(cid:12)159131823Mjup159131823Mjup 14 bital analysis, we therefore rely on a self-consistent dynamical scheme for the RV modeling, coupled with an MCMC sampling algorithm for parameter analy- sis. Each MCMC orbital configuration was further inte- grated with the SyMBA N -body integrator for a max- imum of 1 Myr to analyze the long-term stability and dynamical properties of the system. Our long-term stability results yield that approxi- mately 46% of the edge-on coplanar (i = 90◦, ∆Ω = 0◦) MCMC samples survived 1 Myr. The stable parameter space of confident MCMC configurations suggests plan- ets on nearly circular orbits with (minimum) dynamical masses of mb = 2.0+0.22−0.10 and mc = 1.86+0.18−0.23 MJup. The planetary semi-major axes are ab = 1.57+0.02−0.01 and ac = 2.34+0.03−0.04 au, forming a period ratio of ∼ 11:6, which is preserved during the numerical orbital evolution. Over- all, we find that the HD 202696 system is most likely dominated by secular perturbations near the 11:6 mean- motion commensurability, with no direct evidence of res- onance motions. In this context, the HD 202696 system is not unique. It belongs to a population of massive multiplanet sys- tems orbiting evolved intermediate-mass stars with or- bital period ratios between the first-order MMRs of 2:1 and 4:3. These systems impose challenges for planet migration theories, since during the type II migration phase, they unavoidably must have passed through the strong 2:1 MMR, which requires special conditions for each case and thus in general has a low probability. Un- derstanding the giant-planet migration into high-order MMR commensurability after passing the strong 2:1 MMR as in the case of the HD 202696 system requires more sophisticated disk-planet migration simulations, which are beyond the scope of this study. The discovery of more multiplanet systems around G and K giants will be important to better understand planet formation and evolution around more massive and evolved stars. This work was supported by the DFG Research Unit FOR 2544, Blue Planets around Red Stars, project No. RE 2694/4-1. S.R. further acknowledges the sup- port of the DFG Priority Program SPP 1992, Explor- ing the Diversity of Extrasolar Planets (RE 2694/5- 1). M.H.L. was supported in part by Hong Kong RGC grant HKU 17305015. B.B. thanks the Euro- pean Research Council (ERC Starting grant 757448- PAMDORA) for their financial support. This research has made use of the SIMBAD database, operated at CDS, Strasbourg, France. This work has made use of data from the European Space Agency (ESA) mis- sion Gaia (https://www.cosmos.esa.int/gaia), pro- cessed by the Gaia Data Processing and Analysis Con- sortium (DPAC; https://www.cosmos.esa.int/web/ gaia/dpac/consortium). Funding for the DPAC has been provided by national in particular the institutions participating in the Gaia Multilateral Agreement. We also wish to extend our special thanks to those of Hawaiian ancestry on whose sacred mountain of Maunakea we are privileged to be guests. Without their generous hospitality, the Keck observations pre- sented herein would not have been possible. We thank the anonymous referee for the excellent comments that helped to improve this paper. institutions, Software: CERES pipeline (Brahm et al. 2017a), ZASPE (Brahm et al. 2017b), astroML (VanderPlas et al. 2012), Systemic Console (Meschiari et al. 2009), emcee (Foreman-Mackey et al. 2013) REFERENCES Andr´e, Q., & Papaloizou, J. C. B. 2016, MNRAS, 461, 4406 Anglada-Escud´e, G., Amado, P. J., Barnes, J., et al. 2016, Nature, 536, 437 Arenou, F., & Luri, X. 1999, in Astronomical Society of the Pacific Conference Series, Vol. 167, Harmonizing Cosmic Distance Scales in a Post-HIPPARCOS Era, ed. D. Egret & A. Heck, 13 -- 32 Bailer-Jones, C. A. L., Rybizki, J., Fouesneau, M., Mantelet, G., & Andrae, R. 2018, ArXiv e-prints, arXiv:1804.10121 Baluev, R. V. 2009, MNRAS, 393, 969 Beaug´e, C., Ferraz-Mello, S., & Michtchenko, T. A. 2003, ApJ, 593, 1124 Brahm, R., Jord´an, A., & Espinoza, N. 2017a, PASP, 129, 034002 Brahm, R., Jord´an, A., Hartman, J., & Bakos, G. 2017b, MNRAS, 467, 971 Bressan, A., Marigo, P., Girardi, L., et al. 2012, MNRAS, 427, 127 Butler, R. P., Marcy, G. W., Williams, E., et al. 1996, PASP, 108, 500 Castelli, F., & Kurucz, R. L. 2004, ArXiv Astrophysics e-prints, astro-ph/0405087 Duncan, M. J., Levison, H. F., & Lee, M. H. 1998, AJ, 116, 2067 ESA, ed. 1997, ESA Special Publication, Vol. 1200, The HIPPARCOS and TYCHO catalogues. Astrometric and photometric star catalogues derived from the ESA HIPPARCOS Space Astrometry Mission Foreman-Mackey, D., Hogg, D. W., Lang, D., & Goodman, J. 2013, PASP, 125, 306 Gaia Collaboration, Brown, A. G. A., Vallenari, A., et al. 2018, ArXiv e-prints, arXiv:1804.09365 Gaia Collaboration, Prusti, T., de Bruijne, J. H. J., et al. 2016, A&A, 595, A1 Giguere, M. J., Fischer, D. A., Payne, M. J., et al. 2015, ApJ, 799, 89 Goldreich, P., & Schlichting, H. E. 2014, AJ, 147, 32 Gontcharov, G. A., & Mosenkov, A. V. 2018, VizieR Online Data Catalog, 2354 Butler, R. P., Vogt, S. S., Laughlin, G., et al. 2017, AJ, 153, 208 Ida, S., Guillot, T., & Morbidelli, A. 2016, A&A, 591, A72 15 Johnson, J. A., Payne, M., Howard, A. W., et al. 2011, AJ, 141, Robinson, S. E., Laughlin, G., Vogt, S. S., et al. 2007, ApJ, 670, 16 1391 Kjeldsen, H., & Bedding, T. R. 2011, A&A, 529, L8 Lee, M. H., & Peale, S. J. 2002, ApJ, 567, 596 -- . 2003, ApJ, 592, 1201 Lin, D. N. C., & Papaloizou, J. 1986, ApJ, 309, 846 Lubow, S. H., & D'Angelo, G. 2006, ApJ, 641, 526 Marcy, G. W., & Butler, R. P. 1992, PASP, 104, 270 Marshall, J. P., Wittenmyer, R. A., Horner, J., et al. 2018, ArXiv e-prints, arXiv:1811.06476 Masset, F., & Snellgrove, M. 2001, MNRAS, 320, L55 Masset, F. S., & Papaloizou, J. C. B. 2003, ApJ, 588, 494 Meschiari, S., Wolf, A. S., Rivera, E., et al. 2009, PASP, 121, 1016 Nelder, J. A., & Mead, R. 1965, Computer Journal, 7, 308 Nelson, B. E., Robertson, P. M., Payne, M. J., et al. 2016, MNRAS, 455, 2484 Pierens, A., Raymond, S. N., Nesvorny, D., & Morbidelli, A. 2014, ApJL, 795, L11 Pojmanski, G., Pilecki, B., & Szczygiel, D. 2005, AcA, 55, 275 Press, W. H., Teukolsky, S. A., Vetterling, W. T., & Flannery, B. P. 1992, Numerical recipes in FORTRAN. The art of scientific computing (Cambridge University Press) Robitaille, T. P., Whitney, B. A., Indebetouw, R., & Wood, K. 2007, ApJS, 169, 328 Sato, B., Omiya, M., Harakawa, H., et al. 2012, PASJ, 64, 135 Sato, B., Wang, L., Liu, Y.-J., et al. 2016, ApJ, 819, 59 Stock, S., Reffert, S., & Quirrenbach, A. 2018, A&A, 616, A33 Tadeu dos Santos, M., Correa-Otto, J. A., Michtchenko, T. A., & Ferraz-Mello, S. 2015, A&A, 573, A94 Tan, X., Payne, M. J., Lee, M. H., et al. 2013, ApJ, 777, 101 Trifonov, T., Lee, M. H., Reffert, S., & Quirrenbach, A. 2018, AJ, 155, 174 Trifonov, T., Reffert, S., Tan, X., Lee, M. H., & Quirrenbach, A. 2014, A&A, 568, A64 Valenti, J. A., Butler, R. P., & Marcy, G. W. 1995, PASP, 107, 966 VanderPlas, J., Connolly, A. J., Ivezic, Z., & Gray, A. 2012, in Proceedings of Conference on Intelligent Data Understanding (CIDU), pp. 47-54, 2012., 47 Vogt, S. S., Allen, S. L., Bigelow, B. C., et al. 1994, in Proc. SPIE, Vol. 2198, Instrumentation in Astronomy VIII, ed. D. L. Crawford & E. R. Craine, 362 Walsh, K. J., Morbidelli, A., Raymond, S. N., O'Brien, D. P., & Quirrenbach, A., Reffert, S., & Bergmann, C. 2011, in American Mandell, A. M. 2011, Nature, 475, 206 Institute of Physics Conference Series, Vol. 1331, American Institute of Physics Conference Series, ed. S. Schuh, H. Drechsel, & U. Heber, 102 -- 109 Wittenmyer, R. A., Horner, J., & Tinney, C. G. 2012, ApJ, 761, 165 Wittenmyer, R. A., Johnson, J. A., Butler, R. P., et al. 2016, Rein, H., Payne, M. J., Veras, D., & Ford, E. B. 2012, MNRAS, ApJ, 818, 35 426, 187 Rivera, E. J., Laughlin, G., Butler, R. P., et al. 2010, ApJ, 719, Worthey, G., & Lee, H.-c. 2011, ApJS, 193, 1 Zechmeister, M., & Kurster, M. 2009, A&A, 496, 577 890 APPENDIX 16 Table A1. HIRES Doppler measurements for HD 202696 from Butler et al. (2017) and RV residuals of the Best-fit models from Table 2 Epoch (BJD) RV (m s−1) σRV (m s−1) S index H index 0.0312 2,454,287.927 0.1091 1.96 1.25 1.14 1.39 1.49 1.19 1.39 1.40 1.32 1.23 1.38 1.40 1.31 1.31 1.30 1.40 1.37 1.61 1.47 1.40 1.40 1.43 1.39 1.42 1.33 1.23 1.36 1.27 1.34 1.37 1.35 1.56 9.66 1.26 1.23 1.25 1.29 1.31 1.46 1.31 1.43 1.23 0.1072 0.0899 0.1066 0.0868 0.1132 0.1065 0.1138 0.1055 0.1075 0.1065 0.1114 0.1072 0.1061 0.1143 0.1074 0.1129 0.1251 0.1086 0.0936 0.1145 0.1069 0.1011 0.1075 0.1010 0.1137 0.1097 0.0975 0.1085 0.1062 0.1006 0.1099 0.0659 0.0882 0.1292 0.1071 0.0945 0.1084 0.1084 0.1098 0.1149 0.1113 0.0307 0.0312 0.0308 0.0308 0.0309 0.0308 0.0308 0.0308 0.0310 0.0311 0.0309 0.0308 0.0310 0.0309 0.0307 0.0306 0.0306 0.0307 0.0305 0.0307 0.0308 0.0308 0.0309 0.0309 0.0309 0.0309 0.0307 0.0307 0.0308 0.0309 0.0309 0.0285 0.0308 0.0312 0.0309 0.0307 0.0309 0.0308 0.0307 0.0309 0.0309 o−c Kep. (m s−1) −2.94 −2.53 −5.05 −0.79 −4.21 −4.38 −2.61 −1.13 16.03 o−c N−body (m s−1) −3.57 −4.50 −5.41 −0.85 −4.00 −3.84 −2.07 −0.41 16.72 7.25 0.58 −1.53 −15.66 −17.54 −1.89 1.18 1.10 8.42 6.10 7.88 1.14 −2.59 −2.85 −2.35 0.17 7.89 −1.22 −5.03 −5.25 16.87 −16.91 28.09 −6.39 5.86 −3.05 −6.85 5.29 −8.23 4.67 2.08 1.73 −5.09 7.84 0.93 −1.20 −15.70 −17.59 −1.97 0.87 0.74 8.08 5.92 7.96 1.44 −2.29 −2.57 −2.15 −1.29 7.22 −1.26 −4.34 −4.49 16.67 −17.80 27.74 −6.71 5.65 −3.06 −6.68 5.69 −7.59 5.64 3.36 0.45 −5.03 2,454,399.822 2,454,634.038 2,454,674.900 2,454,718.006 2,454,777.857 2,454,778.819 2,454,805.740 2,454,955.107 2,454,964.128 2,454,984.066 2,454,985.007 2,455,014.963 2,455,015.957 2,455,019.021 2,455,043.881 2,455,076.033 2,455,084.036 2,455,106.909 2,455,133.918 2,455,169.772 2,455,170.691 2,455,187.695 2,455,198.700 2,455,290.146 2,455,373.126 2,455,396.059 2,455,436.743 2,455,455.739 2,455,521.793 2,455,698.124 2,455,770.042 2,455,787.946 2,455,841.869 2,455,880.836 2,455,903.717 2,455,931.693 2,456,098.123 2,456,154.011 2,456,194.986 2,456,522.092 2,456,894.081 2.88 −63.90 −20.49 2.57 10.64 11.66 13.33 10.92 −5.77 −15.69 −23.92 −26.07 −39.57 −41.36 −25.42 −18.51 −9.97 0.00 5.94 17.65 21.46 17.93 20.55 21.99 2.47 −33.29 −52.38 −69.28 −73.42 −53.00 −10.58 53.71 20.04 26.89 8.95 −1.19 2.89 −28.24 −2.81 3.05 −57.44 19.32
1910.01554
1
1910
2019-10-03T15:35:01
The role of C/O in nitrile astrochemistry in PDRs and planet-forming disks
[ "astro-ph.EP", "astro-ph.GA", "astro-ph.SR" ]
Complex nitriles, such as HC3N, and CH3CN, are observed in a wide variety of astrophysical environments, including at relatively high abundances in photon-dominated regions (PDR) and the UV exposed atmospheres of planet-forming disks. The latter have been inferred to be oxygen-poor, suggesting that these observations may be explained by organic chemistry in C-rich environments. In this study we first explore if the PDR complex nitrile observations can be explained by gas-phase PDR chemistry alone if the elemental C/O ratio is elevated. In the case of the Horsehead PDR, we find that gas-phase chemistry with C/O $\gtrsim$ 0.9 can indeed explain the observed nitrile abundances, increasing predicted abundances by several orders of magnitude compared to standard C/O assumptions. We also find that the nitrile abundances are sensitive to the cosmic ray ionization treatment, and provide constraints on the branching ratios between CH3CN and CH3NC productions. In a fiducial disk model, an elevated C/O ratio increases the CH3CN and HC3N productions by more than an order of magnitude, bringing abundance predictions within an order of magnitude to what has been inferred from observations. The C/O ratio appears to be a key variable in predicting and interpreting complex organic molecule abundances in photon-dominated regions across a range of scales.
astro-ph.EP
astro-ph
Draft version October 4, 2019 Typeset using LATEX twocolumn style in AASTeX63 9 1 0 2 t c O 3 . ] P E h p - o r t s a [ 1 v 4 5 5 1 0 . 0 1 9 1 : v i X r a The role of C/O in nitrile astrochemistry in PDRs and planet-forming disks Romane Le Gal,1 Madison T. Brady,2, ∗ Karin I. Oberg,1 Evelyne Roueff,3 and Franck Le Petit3 1Harvard-Smithsonian Center for Astrophysics, 60 Garden St., Cambridge, MA 02138, USA 2California Institute of Technology, Pasadena, CA 91125, USA 3Sorbonne Universit´e, Observatoire de Paris, Universit´e PSL, CNRS, LERMA, F-92190, Meudon, France (Received July 30th, 2019; Revised October 1st, 2019; Accepted October 2nd, 2019) ABSTRACT Complex nitriles, such as HC3N, and CH3CN, are observed in a wide variety of astrophysical envi- ronments, including at relatively high abundances in photon-dominated regions (PDR) and the UV exposed atmospheres of planet-forming disks. The latter have been inferred to be oxygen-poor, sug- gesting that these observations may be explained by organic chemistry in C-rich environments. In this study we first explore if the PDR complex nitrile observations can be explained by gas-phase PDR chemistry alone if the elemental C/O ratio is elevated. In the case of the Horsehead PDR, we find that gas-phase chemistry with C/O (cid:38) 0.9 can indeed explain the observed nitrile abundances, increasing predicted abundances by several orders of magnitude compared to standard C/O assumptions. We also find that the nitrile abundances are sensitive to the cosmic ray ionization treatment, and pro- vide constraints on the branching ratios between CH3CN and CH3NC productions. In a fiducial disk model, an elevated C/O ratio increases the CH3CN and HC3N productions by more than an order of magnitude, bringing abundance predictions within an order of magnitude to what has been inferred from observations. The C/O ratio appears to be a key variable in predicting and interpreting complex organic molecule abundances in photon-dominated regions across a range of scales. Keywords: astrochemistry -- ISM: molecules -- methods: numerical -- photon-dominated region (PDR) -- protoplanetary disks 1. INTRODUCTION Origins of life on Earth must have been closely linked to the emergence of information-rich polymers such as DNA (deoxyribonucleic acid) or RNA (ribonucleic acid). While their initial formation on the early Earth re- mains mysterious, there are plausible chemical path- ways to their building blocks on the early Earth through a nitrile-centered UV-driven chemistry (Powner et al. 2009; Patel et al. 2015; Sutherland 2016). Simple and complex nitriles are abundantly found at all stages of star and planet formation, including in planet-forming disks, suggesting that the organic chemistry that pre- ceded life on Earth is not unique to the Solar System (Chapillon et al. 2012; Oberg et al. 2015; Bergner et al. Corresponding author: Romane Le Gal romane.le [email protected] ∗ Noland Internship at Harvard-Smithsonian Center for Astro- physics (summer 2018) 2018; Loomis et al. 2018). Perhaps surprisingly CH3CN and HC3N are two of the most commonly detected larger organic molecules in disks, and the origins of these high Oberg et al. (2015) and abundances are uncertain. Loomis et al. (2018) both invoke grain-surface chemical pathways to predict sufficient amounts of CH3CN, but these predictions are extremely uncertain due to lack of experimental data on ice nitrile chemistry and desorp- tion. One important observational constraint is that ob- served HC3N and CH3CN emissions appear to come from the upper most layer of disks or disk atmospheres ( Oberg et al. 2015; Bergner et al. 2018; Loomis et al. 2018). Disk atmospheres are proposed analogs to the more well-studied photon-dominated regions (PDR). In- terestingly, complex nitriles have also been detected at unexpectedly high abundances in the deeply character- ized PDR, the Horsehead nebula (Gratier et al. 2013). Located in the Orion constellation and seen almost edge- on (Abergel et al. 2003), the Horsehead nebula consti- 2 Le Gal et al. tutes a perfect template source to study in detail the physics and chemistry occurring in PDRs. With the WHISPER survey1 (Wideband High-resolution Iram- 30m Survey at two Positions with Emir Receivers, PI: J. Pety), the chemistry of this PDR has been surveyed at unprecedented detail, both at the edge of the PDR (defined by the HCO peak emission, Gerin et al. 2009), and toward an interior 'core' position (defined by the DCO+ peak, Pety et al. 2007). Of interest to this study, Gratier et al. (2013) found that the CH3CN emission is ∼ 40 times brighter at the PDR position than in the 'Core'. Similar to protoplanetary disks, this excess in CH3CN could not be explained by gas-phase chemistry alone, and Gratier et al. (2013) instead suggested that a combination of UV-mediated surface chemistry with surface desorption processes were responsible. However, models developed by Le Gal et al. (2017), coupling the Meudon PDR (Le Bourlot et al. 1993; Le Petit et al. 2006; Le Bourlot et al. 2012) and the Nautilus (Hersant et al. 2009; Ruaud et al. 2016) astrochemical codes, could not reproduce the abundance of CH3CN at the PDR posi- tion by about two orders of magnitude when taking these processes into account and advanced alternative expla- nations that (i) either CH3CN originates from deeper inside of the cloud than previously assumed; or, (ii) the photo-desorption rate is higher and ice photolysis rate lower than those currently implemented in models; or, (iii) critical chemical formation pathways are missing in current astrochemical networks. Another possible explanation for these high nitrile abundances could be the elemental gas-phase C/O ra- tio. Indeed, the relative elemental gas-phase abundances of oxygen and carbon are known to strongly impact the chemistry of star-forming regions (van Dishoeck & Blake 1998). For instance, small hydrocarbons, such as C2H, C3H, C3H2 and C4H, observed in a wide variety of as- trophysical objects including PDRs (Fuente et al. 2003; Pety et al. 2005, 2012; Cuadrado et al. 2015; Guzm´an et al. 2015) and protoplanetary disks (Dutrey et al. 1997; Fuente et al. 2010; Henning et al. 2010; Qi et al. 2013; Kastner et al. 2015; Guilloteau et al. 2016; Bergin et al. 2016; Kastner et al. 2018; Cleeves et al. 2018; Bergner et al. 2019; Loomis et al. 2019), are believed to be mainly formed from atomic carbon (i.e. C+ and/or C). How- ever, atomic carbon is readily converted into CO. There- fore, depending upon the UV-shielding and C/O ratio, more or less carbon can be locked into CO, hamper- ing the production of hydrocarbons and more complex 1 http://www.iram-institute.org/∼horsehead/ Horsehead Nebula/WHISPER.html carbon-containing molecules such as CH3CN and HC3N. In planet-forming disks, a super-solar C/O ratio ((cid:38) 0.8) explains the hydrocarbon observations well and is rea- sonably justified by oxygen removal through water for- mation and other non-volatile O-bearing species (e.g. Hogerheijde et al. 2011; Cleeves et al. 2018). Here, we explore whether the observed CH3CN and other complex nitriles in the Horsehead PDR can be ex- plained by pure gas-phase chemistry when taking into account a revised understanding of the cosmic-ray (CR) ionization rate, a more complex gas-phase chemistry network, and most importantly, a C-rich environment. We then carry out a smaller study of complex nitrile production in planet-forming disks with elevated C/O In § 2, we describe the physical and chemical ratios. properties we used and developed within the Meudon PDR Code as well as our fiducial protoplanetary disk model. The resulting molecular abundances and their dependence upon the CR ionization rate, C/O ratio, and complex nitrile formation pathways are presented in § 3. In § 4, we discuss the dominant reaction pathways for the four nitrile molecules detected toward the Horsehead nebula - C3N, HC3N, CH3CN and CH3NC - as well as which parameters affect these nitrile abundances. Our conclusions are summarized in § 5. 2. MODELING For the PDR chemical investigations we use the Meudon PDR Code, tuned to the physical conditions of the Horsehead nebula, and extended to incorporate a more complete gas-phase chemical network for nitriles up to CH3CN and CH3NC in complexity. In the sec- ond, smaller part of this paper we use a fiducial pro- toplanetary disk model previously described in Le Gal et al. (2019) to test whether our nitrile-optimized PDR chemistry can also explain high abundance of complex nitriles in disks. 2.1. PDR Physical structure The Meudon PDR Code is a 1D astrochemical modeling code which considers a stationary plane-parallel slab of gas and dust illuminated by a radiation field (Le Petit et al. 2006), which can be introduced at will. Assuming a cloud at steady-state, it solves the physical and chemical conditions at different visual extinction throughout the cloud, taking into account radiative transfer from UV absorption, cooling emissions, and heating processes. Figure 1 displays the typical physical structure we computed for the present study, assuming that the cloud has a fixed pressure of 4 × 106 K cm−3 in the PDR re- gion, and a constant density of 2× 105 cm−3 in the core (Habart et al. 2005), i.e. in our model for AV (cid:38) 2 mag. The role of C/O in nitrile astrochemistry in PDRs and planet-forming disks 3 Table 1. Initial gas-phase elemental abundances Species He O(b) C N S Si Fe (a) ni/nH 0.1 3.02 × 10−4 1.38 × 10−4 7.95 × 10−5 3.50 × 10−6 1.73 × 10−8 1.70 × 10−9 Note -- (a) from Pety et al. (2005) & Goicoechea et al. (2006) (b) To test the impact of the C/O ratio, we varied the oxygen elemental abundance in the range [3.45 − 0.92] × 10−4 (see § 3.2). per H2 being a typical dense cloud value (e.g. Goicoechea et al. 2009). We highlight here that what we labeled as ζ in the present study is the CR ionization rate per H2, which corresponds to approximately twice the value of the CR ionization rate per H atom (Glassgold & Langer 1974). In a previous modeling study of the Horsehead nebula, Rimmer et al. (2012) found that chemical pre- dictions are in better agreement with observations when ζ is allowed to vary across the cloud, considering the fol- lowing equation adapted from Nath & Biermann (1994) by Rimmer et al. (2012): ζ = 3.05 × 10−16(AV)−0.6 + 10−17 s−1per H2. (1) For the PDR position, where AV ≈ 2 mag, Eq. (1) gives ζ ≈ 2 × 10−16 s−1 per H2. In § 3, we test the impact of this higher value of ζ on the nitrile chemistry, by comparison to the canonical value of ζ = 5×10−17 s−1 per H2 used in Pety et al. (2005) and Goicoechea et al. (2006). Lastly, while the public version of the Meudon PDR Code (v.1.5.2) does not include grain chemistry, it does model the formation of H2 on grains and computes the charge and temperature distribution of grains. In this study, we kept the default grain size distribution, i.e with grain radius from 1 × 10−3 to 0.3 µm and their relative abundances described by the MRN distribution (Mathis et al. 1977). 2.2. PDR model Chemistry Each model was performed using the same initial abundance set as in Pety et al. (2005) and Goicoechea et al. (2006), except for the oxygen abundance that we varied in some models to explore the impact of the C/O ratio on the chemistry (Table 1). We updated and extended the PDR Meudon Code (v.1.5.2) chemi- Figure 1. Horsehead nebula profiles of the temperature (top panel), the density (middle panel) and the UV flux (bottom panel) as function of the visual extinction, AV. The incident radiation upon this cloud is that of σ Ori, an O 9.5 V star system, which results in an incident FUV intensity upon the cloud of about χ = 60 (i.e 60 × the ISRF in Draine's units ≈ 60 × 2.7 × 10−3 erg s−1 cm−2, Draine 1978; Habart et al. 2005). The physical structure shown here was built considering standard ini- tial gas-phase elemental abundances, see Table 1, i.e. a C/O ratio of ≈ 0.46 (Pety et al. 2005; Goicoechea et al. 2006). Another parameter to consider is the CR ionization rate ζ. Low energy cosmic rays (10-100 MeV, e.g Gre- nier et al. 2015) can penetrate deep into dense clouds, producing ions that drive the gas-phase chemistry via fast ion-neutral reactions. Diffuse clouds usually present higher values of ζ than denser clouds (e.g. Indriolo et al. 2015; Le Petit et al. 2016), with ζ ≈ (1−5)×10−17 s−1 101102103Temperature (K)104105Density (cm3)102101100101AV (mag)1014109104101UV flux (Draine's units) 4 Le Gal et al. cal network with 39 species and 913 reactions relevant to the chemistry of C3N, HC3N, and CH3CN, that we extracted from the KIDA database2 for most of them. We also extended the chemical network to the chem- istry of CH3NC, based on theoretical studies (e.g. De- frees et al. 1985) and the chemistry of its isomer CH3CN (see § 3.3). In total, our network is composed of 191 species and 3616 chemical reactions, including gas-phase bi-molecular reactions (i.e. radiative associations, ion- neutral and neutral-neutral reactions), recombinations with electrons, ionization and dissociation reactions by direct cosmic rays and secondary photons (i.e. photons induced by cosmic rays), and by UV-photons (see Le Petit et al. 2006, for rate formulae details). The criti- cal reactions discussed in this paper are summarized in Table A1, with rates and references. UV-photo-reactions are expected to play a crucial role in PDR chemistry. The Meudon PDR Code allows the choice between two different methods to compute the photo-reaction rates: 1) if the photo-ionization and/or photo-dissociation cross-sections of the molecule is known, the most accurate approach consists in inte- grating this cross-section over the radiation field at each given position in the cloud; 2) if the photo-cross-section of the molecule is unknown, an analytical expression as function of the visual extinction is estimated, i.e. consid- ering fixed fitted parameters for each molecules and com- puting their photo-rates as function of the visual extinc- tion (see for an example Eq. (14) of Heays et al. 2017). We updated the cross-sections of all the molecules avail- able in the Leiden database3 that are included in our chemical network in the Meudon PDR Code (Heays et al. 2017). 2.3. Protoplanetary disk physical structure Our fiducial protoplanetary disk astrochemical model is based on the MWC 480 disk model of Le Gal et al. (2019), which consists in a 2D parametric physical struc- ture onto which the chemistry is post-processed (see §2.4). The disk physical structure assumes a disk that is symmetric azimuthally and with respect to the mid- plane. Thus, it can be described in cylindrical coordi- nates centered on the inner star along two perpendicu- lar axes characterizing the radius and height in the disk. Figure 2 represents the profiles of the gas temperature, density, visual extinction and UV flux throughout the disk, for which the parameterization is briefly summa- rized below and further described in Le Gal et al. (2019). 2 http://kida.obs.u-bordeaux1.fr/ 3 https://home.strw.leidenuniv.nl/∼ewine/photo/ For a given radius r from the central star, the vertical temperature profile is computed following the formalism of Rosenfeld et al. (2013) and Williams & Best (2014), originally developed by Dartois et al. (2003):  Tmid + (Tatm − Tmid) Tatm (cid:104) (cid:17)(cid:105)2δ (cid:16) πz 2zq sin T (z) = if z < zq if z ≥ zq, (2) where Tmid and Tatm are respectively the midplane and atmosphere temperatures that vary as power law of the radii (Beckwith et al. 1990; Pi´etu et al. 2007; Le Gal et al. 2019). zq = 4H with H the pressure scale height that, assuming vertical static equilibrium, can be ex- pressed as follows: (cid:115) H = kB Tmid r3 µ mH G M(cid:63) , (3) with kB the Boltzmann constant, µ = 2.4 the reduced mass of the gas, mH the proton mass, G the gravitational constant, and M(cid:63) the mass of the central star. The mid- plane temperature Tmid is estimated following a simple irradiated passive flared disk approximation (e.g. Chi- ang & Goldreich 1997; Dullemond et al. 2001; Huang et al. 2018): (cid:18) ϕL(cid:63) (cid:19)1/4 8πr2σSB Tmid(r) ≈ , (4) with L(cid:63) = 24 L(cid:12) the stellar luminosity (Andrews et al. 2013), σSB the Stefan-Boltzman constant and ϕ = 0.05 a typical flaring angle. The atmosphere tem- perature, Tatm, is based on observational constraints. So here we consider Tatm = Tatm,100 au( 100 au ), with Tatm,100 au=48 K from Guilloteau et al. (2011). r The disk is assumed to be in hydrostatic equilibrium. Thus, for a given vertical temperature profile, the verti- cal density structure is determined by solving the equa- tion of hydrostatic equilibrium, as described from Eq. (17) to (20) in Le Gal et al. (2019). The surface density of the disk is assumed to follow a simple power law vary- ing as r−3/2 (Shakura & Sunyaev 1973; Hersant et al. 2009): Σ(r) = ΣRc , (5) (cid:18) r (cid:19)−3/2 Rc where ΣRc is the surface density at the characteristic radius that can be expressed as function of the mass of the disk, Mdisk, and its outer radius, Rout: ΣRc = −3/2 c √ MdiskR 4π Rout , (6) with here Mdisk = 0.18 M(cid:12) (Guilloteau et al. 2011). The role of C/O in nitrile astrochemistry in PDRs and planet-forming disks 5 Figure 2. Disk physical structure fed in our fiducial protoplanetary disk astrochemical model. The 2D temperature (first panel), density (second panel), visual extinction (third panel) and UV flux (fourth panel) profiles are represented as functions of disk radius versus height, both in au. The dashed black line, on the densities panels, delineates 1 scale height. The visual extinction profile is derived from the hy- drostatic density profile using the gas-to-extinction ratio of NH/AV = 1.6× 1021 (Wagenblast & Hartquist 1989), with NH = N (H) + 2N (H2) the vertical hydrogen col- umn density of hydrogen nuclei. This gas-to-extinction ratio assumes a typical mean grain radius size of 0.1 µm and dust-to-mass ratio of 0.01, consistent with model assumptions. Finally, the UV flux profile is computed considering the UV flux impinging the disk convolved with the vi- sual extinction profile. The unattenuated UV flux fac- tor, fUV, at a given radius r depends on both the pho- tons coming directly from the central embedded star and on the photons that are downward-scattered by small grains in the upper atmosphere of the disk. Thus, fol- lowing Wakelam et al. (2016), we consider: fUV = fUV,Rc/2 (cid:17)2 (cid:16) r Rc (cid:17)2 . (cid:16) 4H Rc + (7) 2.4. Protoplanetary disk chemical model The disk chemistry is computed time-dependently in 1+1D using the gas-grain astrochemical model Nautilus (v.1.1) (Hersant et al. 2009; Wakelam et al. 2016) in three phase mode (Ruaud et al. 2016), i.e. in- cluding gas-phase, grain-surface and grain-bulk chem- istry (see Le Gal et al. 2019, for more details). First, the chemical evolution of a representative starless dense molecular cloud is modeled up to a characteristic age of 1 × 106 years (e.g. Elmegreen 2000; Hartmann et al. 2001). For this 0D model we use typical constant phys- ical conditions: grain and gas temperatures of 10 K, a gas density of 2 × 104 cm−3 and ζ = 5 × 10−17 s−1 per H2; this parent molecular cloud is also considered to be shielded from external UV photons by a visual extinc- tion of 30 mag. For consistency, we use the same initial abundances as for our PDR model (see Table 1) for this first simulation step. The outcoming chemical gas and ice compositions of this parent molecular cloud are then used as initial chemistry for our 1+1D disk model. Sec- ond, we ran the chemistry of our 1+1D disk model up to one million years, typical chemical age of a disk when grain growth is not considered (e.g. Cleeves et al. 2015). While the disk chemistry has not reached steady state at that time, its evolution is slow enough that the re- sults presented here hold for a disk twice younger or older. Note that in contrast to the PDR model, the disk chemical code does include grain surface reactions. How- ever the grain-surface reactions pathways to CH3CN and HC3N, the two molecules or particular interest for this study, remain poorly constrained. 3. RESULTS 3.1. Impact of cosmic-ray treatment Figure 3 presents the abundances of C3N, HC3N and CH3CN computed with the Meudon PDR Code as func- tion of the visual extinction AV, for two models. Both models consider our new chemical network and the ini- tial gas-phase elemental abundances prescribed in Pety et al. (2005) and Goicoechea et al. (2006) (see Table 1) but each model uses a different CR ionization rate. The standard model uses the CR ionization rate canonical value of ζ = 5 × 10−17 s−1 per H2 (Pety et al. 2005; Goicoechea et al. 2006), and the high-ζ model uses a higher CR ionization rate of ζ = 2 × 10−16 s−1 per H2, as calculated from (Eq. 1). By impacting the ion abun- dances in molecular clouds, the CR ionization rate in- directly drives the abundances of their daughter neutral molecules (see § 2.1). The nitrile abundances are in- deed higher with the high-ζ model than with the stan- dard model, but both models under-predict by several orders of magnitude the abundances observed toward the Horsehead nebula. Rimmer et al. (2012) showed that a varying ζ across the cloud tends to produce more accurate results. How- 100200300400Radius (au)0255075100125150175Height (au)1720253550678910-101231.0-5.00.02.02.5-10.00102030405060708090Temperature (K)56789101112log(Density) (cm3)21012345log(AV) (mag)108642024log(UV flux) (Draine's units) 6 Le Gal et al. phase. In order to mimic the differential freeze-out of volatiles on grains, we varied the oxygen gas-phase ele- mental abundance from 3.45 × 10−4 to 9.2 × 10−5 while keeping the carbon abundance fixed. This led to a varia- tion of the C/O ratio from 0.4 to 1.5. The lowest consid- ered O abundance is a factor of two higher than the CO abundance derived in the Horsehead PDR (5.6 × 10−5, Pety et al. 2005), while the highest considered O abun- dances is below the cosmic O abundance of 4.9 × 10−4 (Asplund et al. 2009) to 5.75 × 10−4 (Przybilla et al. 2008). The choice of fixing the carbon elemental abun- dance and varying the oxygen one, is also justified by the fact that, between these two elements, the elemen- tal gas-phase abundance of oxygen is the less constrained (Jenkins 2009; Whittet 2010; Jones & Ysrad 2019), as discussed in Le Gal et al. (2014). Figure 4 shows the impact of the gas-phase C/O ratio on the abundances of C3N, HC3N and CH3CN and on the gas temperature as function of the visual extinction AV in our model of the Horsehead nebula. An O-poor chemistry (i.e. a high C/O ratio) results in higher abun- dances of the three nitriles. For a CR ionization rate of ζ = 2 × 10−16 s−1 per H2 and a gas-phase C/O ratio in the range 0.9 − 1.5, our new gas-phase chemistry model can reproduce the three nitrile observations at the PDR position within an order of magnitude. As for the Core position, our best fit models are found for lower C/O ratios, in the range 0.6 − 0.9. This lowering of C/O with increasing visual extinction could be explained by photon-mediated release of refractory carbon into gas phase in the PDR region, and/or the onset of freeze out of carbon species in the core region. Whatever the mechanism, the decrease of C/O with increasing visual extinction suggests that the gas-phase C/O ratio vary across astrophysical objects. It is also important to mention that the chemical rates used in astrochemical models sometimes present large uncertainties. We ran two additional models to test the impact of such uncertainties on the major reaction rates listed in Table A1 which are driving the complex nitrile chemistry. These additional simulations compute the chemistry with (i) the maximum allowed rates, and (ii) the minimum allowed rates. The results are that the nitrile abundances of interest for this study vary by less than a factor of three in the PDR and Core regions, which is small compared to the more than two orders of magnitude mismatch between observations and models using the standard C/O value. Though this does not constitute a rigorous detailed sensitivity analysis such as those developed for instance by Vasyunin et al. (2004, 2008) and Wakelam et al. (2005, 2006, 2010), our simple analysis suggests that our results are robust. Figure 3. Computed C3N (dark blue) HC3N (purple) and CH3CN (orange) abundances with respect to H nuclei, as function of the visual extinction AV obtained with the stan- dard model (solid lines) and with the high-ζ model (dashed lines), see § 3.1. These model results are compared to the observations from Gratier et al. (2013) (dashed boxes and arrow). 50% error bars are included on the observations. The PDR (1 mag < AV < 2 mag) and Core (AV (cid:38) 8 mag) regions are shaded. ever, for the visual extinctions associated with the PDR region and for molecules of interest studied here, vary- ing ζ across the cloud does not significantly impact the results compared to a constant-ζ model. In the Core region, the varying-versus-constant ζ model abundances are about half an order of magnitude different. Changing the CR ionization rate also impacts the gas temperature, as shown in the bottom panels of Fig. 4. Typically, increasing ζ shifts the temperature gradient closer to the PDR's edge. As a result, a higher ζ leading to higher temperatures and thus higher reaction rates, the absolute abundances of nitriles are slightly increased in the PDR with a high-ζ model. In addition, likely due to the shift in temperature, the abundance patterns are shifted outwards (i.e. closer to the illuminated edge of the PDR) in the case of a higher ζ. In summary, while the CR treatment has an impact on the complex ni- trile chemistry, increasing nitrile abundances by almost an order of magnitude in the PDR region, the effect is small compared to the mismatch between models and observations (see Fig. 3). 3.2. Impact of the C/O ratio Major carriers of O and C are expected to freeze out In particular a substantial under different conditions. amount of O can become incorporated into water ice which is one of the least volatile common interstellar molecules, resulting in an elevated C/O ratio in the gas- 101100101AV (mag)10171016101510141013101210111010109n(X)/nHC3NHC3NCH3CNC3NHC3NCH3CN The role of C/O in nitrile astrochemistry in PDRs and planet-forming disks 7 Figure 4. C3N, HC3N and CH3CN abundances with respect to H nuclei, as well as the gas temperature, computed with our model of the Horsehead nebula as function of the visual extinction, AV, for varying C/O and the standard- (left column) and high- (right column) CR ionization rates considered in this study. The observations from Gratier et al. (2013) are represented by the black hatched boxes, which consider 50% error bars, and the downward arrow. The core (AV > 8 mag) and PDR (1 mag < AV < 2 mag) regions are shaded. n(C3N)/nHPDRCore=5×1017 s1 per H2PDRCore=2×1016 s1 per H2n(HC3N)/nH10181016101410121010108106n(CH3CN)/nHC/O = 0.400.550.700.800.901.001.201.50C/O = 0.400.550.700.800.901.001.201.50101100101AV (mag)101102Temperature (K) 8 Le Gal et al. Since we are mainly interested in the PDR nitrile chemistry, we consider our best-fit model the model with the lowest C/O ratio that reproduces at the PDR posi- tion the three complex nitrile abundances shown Fig. 4. The model with ζ = 2×10−16 s−1 per H2 and C/O = 1.0 fulfills these criteria. Figure 5 shows the modeled abundances, obtained with our best-fit model, of other typical oxygenated and carbonated molecules that were also observed toward the Horsehead nebula, i.e CO (Pety et al. 2005), HCO+ (Goicoechea et al. 2009), H2CO (Guzm´an et al. 2011), and the hydrocarbons C2H, c-C3H, and C4H (Pety et al. 2005; Guzm´an et al. 2015). For comparison, the stan- dard model results are also presented in the same fig- ure. The C/O variation does not significantly impact the CO abundance throughout the cloud, which is con- sistent with the fact that CO is the main reservoir of carbon and the carbon elemental abundance is fixed in our models. Our best fit model displays better agree- ment between model and observations in the PDR for the other O-bearing molecules we consider, H2CO and HCO+. With regards to the hydrocarbons, our best model better matches the observations at the PDR position, generally by orders of magnitude, compared to the stan- dard model. This is an expected results because atomic O is a main destroyer for small hydrocarbons such as C2H, c-C3H and C4H (Millar et al. 1987; Millar & Herbst 1990). Thus, while C is kept constant, diminishing O increases the hydrocarbon abundances. However, even our best fit model does not fully reproduce the observed abundances (see for instance C4H in Fig. 5), indicative of that the C/O ratio does not provide a complete expla- nation for the abundant hydrocarbon chemistry in the Horsehead PDR. Regarding the Core position, our best fit model gener- ally over-predicts the observations, but this might sim- ply be explained by the fact that freeze-out on grains is not included in our model. It could also be that the gas-phase C/O ratio decreases within the nebula, as sug- gested by the C/O grid results shown in Fig. 4 when compared to the observations in each observed positions. 3.3. CH3NC vs CH3CN chemical pathways isocyanide (CH3NC), Another interesting nitrile molecule to study is the methyl the isomer of methyl cyanide (CH3CN). First detected toward Sgr B2 (Cer- nicharo et al. 1988; Remijan et al. 2005), CH3NC was also detected toward the Horsehead nebula (Gratier et al. 2013), Orion KL (L´opez et al. 2014), and more recently toward the solar-type binary protostar IRAS 16293-2422 (Calcutt et al. 2018). A few theo- retical and experimental studies have investigated the isomers' chemistry and their abundance ratio (Huntress & Mitchell 1979; Defrees et al. 1985; Anicich et al. 1995), and converged on the same major gas-phase production pathways for both via the reaction: CH3 + + HCN −→ (CH3NCH+)∗ k2−→ CH3NCH+ + hν, (8) k3−→ CH3CNH+ + hν, (9) followed by the with k2 and k3 given in Table A1, dissociative recombinations of both protonated ions CH3NCH+ and its isomer CH3CNH+ to form CH3NC and CH3CN, respectively (see reaction rates in Ta- ble A1). However, the branching ratio is poorly con- strained and depends on the stabilization processes of the intermediate complex (CH3NCH+)∗ (e.g. Anicich et al. 1995). Due to its lower energy state, CH3CNH+ is + + found to be the major product of the reaction CH3 HCN (9). However, its formation requires the isomeriza- tion of the intermediate complex (CH3NCH+)∗, which likely happens due to collisions with a third body. Thus, the ratio between the two isomeric ions depends on the competition between the relaxation and isomeriza- tion rates of the intermediary complex. The resulting CH3NCH+/CH3CNH+ ratio was estimated to lie in the range 0.1 -- 0.4 by one theoretical study and assumed to propagate to a CH3NC/CH3CN ratio of 0.1 -- 0.4 via the respective subsequent dissociative recombinations (De- frees et al. 1985). Here, we investigated the impact of the branching ratios in between the pathways (8) and (9) on the re- sulting CH3NC/CH3CN ratio in our PDR model. Fig- ure 6 presents the results obtained using our best fit model and three different branching ratios leading to: 100%, 80% and 0% of isomerization. The best fit re- sults are obtained for a branching ratio of 80% (i.e. CH3NC/CH3CN ∼ 0.2), in agreement with the theo- retical calculation of Defrees et al. (1985). However, to our knowledge the CH3NCH+ dissociative recombina- tion has not been studied yet and even though the rate of the CH3CND+ dissociative recombination was mea- sured (Vigren et al. 2008) its branching ratio remains un- certain (e.g. Plessis et al. 2010, 2012; Loison et al. 2014). It would thus be interesting to study whether the dis- sociative recombination of CH3CNH+ and CH3NCH+ could lead to disproportionate prevalence of each initial isomer. Further theoretical and experimental studies are therefore needed to assess the validity of our astrochem- ically motivated branching ratios. The role of C/O in nitrile astrochemistry in PDRs and planet-forming disks 9 Figure 5. CO, HCO+, H2CO, C2H, c-C3H and C4H abundances with respect to H nuclei, as function of the visual extinction AV computed with our best-fit model (solid lines) and standard model (dashed lines) of the Horsehead nebula, compared to published observations (Pety et al. 2005; Goicoechea et al. 2009; Guzm´an et al. 2011, 2015) represented by the filled and hatched boxes and the downward arrows. 50% error bars are included on the observations. The core (AV > 8 mag) and PDR (1 mag < AV < 2 mag) regions are shaded. 3.4. Complex nitrile production in a protoplanetary disk with a high C/O ratio To test if our new understanding of the complex ni- trile PDR chemistry can be generalized to disks, we used a fiducial protoplanetary disk astrochemical model, loosely based on the disk around MWC 480, from Le Gal et al. (2019) described § 2.3 and § 2.4. We ran the chemical post-processing for two different C/O ratios: (i) C/O= 0.46, as in our PDR standard model, and (ii) C/O= 1.0 as in our PDR best fit model. For each of these C/O ratios we ran two disk models, a full gas- grain model and a gas-grain model where CH3CN and HC3N are only formed in the gas phase. The results of this total of four disk models on the abundances of HC3N and CH3CN are shown in Fig. 7. A huge gap is observed from ∼ 25 to ∼ 200 au in the computed nitrile column densities for the standard C/O ratio disk mod- els, and disappears for higher C/O. The prevalence of grains' pathway formation increases with C/O, and even becomes negligible in the formation of HC3N for stan- dard C/O. Even though our disk model also includes grain chemistry, the main result from our PDR study holds for disk astrochemistry, i.e. that an elevated C/O ratio better reproduce the nitrile observations. Without any tuning of our disk model, our best-fit model predic- tions are within an order of magnitude for the CH3CN case. For the HC3N, the results are in agreement, at the order of magnitude level, for the inner 100 au of Figure 6. CH3CN (dark blue) and CH3NC (orange) abun- dances with respect to H nuclei, as function of the visual extinction, AV, in the modeled Horsehead nebula. Three different models, based on our best-fit model (see § 3.2), are depicted here, testing the izomerization branching ratio of the reaction pathway (9): (i) 100% (dotted lines); (iii) 80% (solid lines - best model); (iii) 0% (dashed lines). The model results are also compared to published observations from Gratier et al. (2013), where 50% error bars are included (dashed boxes and downward arrow). The core (AV > 8 mag) and PDR (1 mag < AV < 2 mag) regions are shaded. 101100101AV (mag)10191017101510131011109107105103n(X)/nHPDRCoreCOHCO+H2COPDRCoreC2Hc-C3HC4Hbest-fit modelstandard model101100101AV (mag)10181016101410121010108106n(X)/nHPDRCore=2×1016 s1, C/O = 1.0CH3CNCH3NCCH3CNCH3NC 10 Le Gal et al. Figure 7. Radial profile of the column density of CH3CN and HC3N computed with our fiducial protoplanetary disk astrochemical model for four different models, differing in (i) their C/O ratio, with (a) C/O= 0.46 (in purple) as in our PDR standard model and (b) C/O= 1.0 (in orange) as in our PDR best fit model; (ii) their grain chemical network that includes (solid lines) or not (dashes lines) the formation of CH3CN and HC3N on grains. The horizontal gray lines represent the column density derived from the MWC 480 disk observations of (Bergner et al. 2018). the disk, where likely most of the emission originates (Bergner et al. 2018). 4. DISCUSSION 4.1. Nitrile formation pathways In our PDR models, the C3N formation is dominated by the reaction: c-C3H + N k4−−→ H + C3N, (10) with k4 given in Table A1. c-C3H being itself mainly +. produced by the electronic recombination of c-C3H2 Thus, the under-prediction of C3N by our model in the PDR region could be explained by the under-prediction of c-C3H (see Fig. 5). Therefore, for a fixed abun- dance of N, a carbon-enriched medium would enhance the production of C3N. As for its destruction, photo- dissociation dominates in the PDR (see the correspond- ing reaction rate Table A1), but depending on the C/O ratio two other reactions also participate in the C3N destruction, i.e.: O + C3N k5−−→ CO + CCN, C + C3N k6−−→ C3 + CN, (11) (12) with k5 and k6 given in Table A1. Reaction (11) even becomes the primary destruction pathway of C3N in O- rich (C/O ≈ 0.4) PDR. While, a priori, it may seem odd for a photo-dissociation process to not be the pri- mary destruction mechanism in PDR, in our model this is explained by the high concentration of atomic oxygen in this region for low C/O ratios. In O-poor (C/O (cid:38) 1) PDR, C3N is primarily destroyed through a combination of photo-dissociation and by atomic carbon (12). HC3N is formed from a variety of different reactions involving carbon- and nitrogen-containing molecules: HC3NH+ + e− k7−−→ H + HC3N, C4H + N k8−−→ C + HC3N, C + H2CCN k9−−→ H + HC3N, c-C3H2 + N k10−−→ H + HC3N, C2H2 + CN k11−−→ H + HC3N, (13) (14) (15) (16) (17) with k7 to k11 given in Table A1. Thus, with a lower amount of one of the main hydrocarbon destroyer, atomic O, and a higher amount of 'free' carbon in the gas phase, more reaction pathways meaningfully con- tribute to the formation of HC3N, via these diverse carbon-rich intermediates. HC3N is mainly destroyed by UV-photons up to an AV ≈ 4 mag, via the reaction: HC3N + hν k12−−→ C2H + CN, (18) with k12 given in Table A1. For AV (cid:38) 4 mag, the im- pact of destruction by dominant ions (e.g. H+, H3 +, 0100200300400Radius (au)10510710910111013N[X] (cm2)CH3CNCH3CNHC3NHC3NC/O = 0.46C/O = 1.0full gas-grain modelgas-grain model w/o CH3CN nor HC3N formation on grains The role of C/O in nitrile astrochemistry in PDRs and planet-forming disks 11 H3O+, C+, HCO+) gradually increases with the opti- cal depth, since UV-photon penetration diminishes and most of the ion abundances increase. Other destruc- tion pathways involving atomic carbon forming bigger carbon chain molecules appear with increasing optical depth, but these are typically far less common in our grid models and are only relevant in dense clouds with higher carbon abundances. The formation of CH3CN is dominated by the dis- sociative recombination of CH3CNH+ with electrons, where CH3CNH+ itself is primarily formed by the ra- diative association (9) and the following: HNC + CH3 + k13−−→ CH3CNH+ + hν, (19) with k13 given in Table A1. Similarly, and as already presented in § 3.3, the formation of CH3NC is dom- inated by the dissociative recombination of CH3NCH+ with electrons, with CH3NCH+ primarily formed by the + is formed via successive radiative association (8). CH3 hydrogenation from C+. HCN and HNC also descend from atomic carbon (Le Gal et al. 2014; Loison et al. 2014). As a consequence, the formations of CH3CN and its isomer CH3NC seem to be guided mostly by the carbon abundance, explaining that the abundance of CH3CN increases with C/O (see Fig. 4). The destruc- tion of the isomers are dominated by photo-dissociation: CH3CN/CH3NC + hν k14−−→ CH3 + CN, (20) with k14 given in Table A1. To summarize, the enhancement in nitrile abundances appear to be tightly correlated with the C/O ratio. More interestingly however, our study highlights the im- portance of the relative elemental gas-phase abundances with respect to one another, and emphasizes the indirect role of oxygen in nitrile chemistry. The latter has a dra- matic impact on the carbon chemistry in O-rich molec- ular environments, where most of the carbon is rapidly locked in CO and atomic O is a main destroyer for hydro- carbons, hampering the development of more complex carbon chemistry. In addition, we have shown in § 3.2 that the nitrile abundances increase with ζ (Fig. 4). In- deed, the ionization fraction is directly linked to ζ, which +, HC3NH+ governs the production of ions, such as CH3 +, parent molecules of the complex nitriles and c-C3H2 + react with HCN and HNC to here under study. CH3 produce CH3CNH+, that readily recombines with elec- trons to form CH3CN; HC3NH+ recombines with elec- + recombines with trons to form HC3N (13); and c-C3H2 electrons to form c-C3H which reacts with N to produce C3N (10). 4.2. The role of C/O in PDR and disk atmosphere nitrile chemistry The strong impact of the C/O ratio on the nitrile chemistry can likely be explained by the fact that in the gas phase, for a fixed amount of C element, O-removal decreases one of the main destruction pathways of hy- drocarbons, which are the parent molecules of nitriles. To this effect adds the reactions of small carbon and oxy- gen species to form CO. Typically, in molecular clouds, most of the carbon hastily reacts with all available oxy- gen to form CO, effectively removing it from the reaction pathways that build up more complex molecules, such as complex nitriles. O-removal thus leaves more 'free' car- bon available in the gas phase to form carbon-containing molecules such as hydrocarbons, carbon chains and ni- triles. As a comparison, the standard model results in a CO abundance of almost 1.38 × 10−4, i.e. quasi all the carbon available in our models (see Table 1), whereas our best-fit model produces a CO abundance of ∼ 1.34× 10−4. This leaves ∼ 6.0× 10−6 'free' carbon to build more complex carbon-containing molecules. Varying the C/O ratio also impacts the gas tempera- ture for AV (cid:46) 1.5 mag, as shown in the last panel row of Fig. 4. An increase in C/O ratio increases the gas temperature. This is due to the fact that, in the present work, we varied the abundance of atomic O to change the C/O ratio, thus reducing one of the main gas coolant in this region of the PDR. To test the impact of varying the C/O ratio via the carbon elemental abundance in- stead of the oxygen, we ran additional models. The main difference is in the resulting CO abundance. For a fixed C/O ratio, it increases with the amount of carbon. As a consequence, the gas temperature decreases at the edge of the PDR, since CO is another important gas coolant. However, these differences diminish with the increase of the visual extinction and are minor at the PDR position and in particular on the nitrile abundances. Thus, the main results found in the present work on the C/O im- pact on the Horsehead nebula chemistry is agnostic to whether C or O is varied to achieve a C/O ratio of ∼1. In disks, the C/O ratio is also strongly impacting the nitrile chemistry with the additional effect of changing the morphology of the nitrile abundance throughout the disk, as shown in Fig. 7. Whether or not CH3CN forms in gas or through gas-grain chemistry in disks, an el- evated C/O in disks also helps in better reproducing the observations. However, grain chemistry seems to be needed to better reproduce the observations, in agree- ment with Oberg et al. (2015) and Loomis et al. (2018). Our finding of a C/O(cid:38) 1 needed to reproduce the com- plex nitriles chemistry observed in disk atmospheres is in good agreement with the results of Bergin et al. (2016) 12 Le Gal et al. that also find that a C/O ratio exceeding unity is re- quired to reproduce the observations of C2H in disks. These authors therefore proposed C2H as probe of C/O- enriched disk layers. Here we propose that complex ni- triles could also serve as such probe in both PDR and disks, with the vantage of also probing the internal parts of the latter ((cid:46) 100 au) as predicted by our models (see Fig. 7). Moreover, the fact that complex nitriles, such as HC3N and CH3CN, are much more commonly ob- served in disks than O-containing complex molecules, such as CH3OH, highly suggests that organic chemistry is regulated by the C/O ratio and that disk atmosphere chemistry seems to be more generally C-rich than O- rich. 5. CONCLUSIONS We tackled the chemistry of nitriles in PDR, in or- der to, first, understand the observations found for the nitriles C3N, HC3N, CH3CN and CH3NC toward the Horsehead PDR (Gratier et al. 2013), and second, test if our improved PDR chemistry could help to explain the recent observations of HC3N and CH3CN in disk atmospheres ( Oberg et al. 2015; Bergner et al. 2018). To this aim, we extended the chemistry of the Meudon PDR code (v.1.5.2) up to these four nitriles and ex- plored the impact of some key parameters, such as the cosmic-ray ionization rate ζ and the gas-phase elemen- tal C/O ratio, on our modeled nitrile results. Our main conclusions are summarized below: 1. Varying the C/O ratio between 0.4 and 1.5 in a model of the Horsehead PDR results in orders of magnitude changes in nitrile abundances. 2. A gas-phase C/O ratio of (cid:38) 0.9 can reproduce the C3N, HC3N, CH3CN and CH3NC abundances within an order of magnitude in the Horsehead PDR, without any grain-surface chemistry. 3. The cosmic-ray ionization rate moderately affects the nitrile chemistry through its impact on elec- +). trons and hydrocarbon ions (e.g. CH3 A ζ = 2× 10−16 s−1 per H2 better fits the observa- tions than the standard value of ζ = 5× 10−17 s−1 per H2. +, C3H2 4. Our best fit PDR model (i.e. with ζ = 2 × 10−16 s−1 per H2 and C/O= 1) can reproduce both the relative abundances of CH3NC and CH3CN when adopting a branching ratio of 0.8 isomeriza- tion for the reaction CH3 + + HCN (9). 5. An elevated C/O ratio (∼ 1.0) could also be the key for understanding complex nitrile disk chemistry. Using a fiducial protoplanetary disk astrochemical model, we find that disk observa- tions of CH3CN and HC3N are reproduced within an order of magnitude, while our standard model under-predicts the same molecules by 2-3 orders of magnitude. While the good agreement between observations and models in both a classic PDR and a planet-forming disk, it is important to note that the nitrile grain chemistry is still poorly constrained, and it may contribute to both kinds of regions (e.g. Bertin et al. 2017a,b; Calcutt et al. 2018; Nguyen et al. 2019). Further experiments and the- ory on nitrile grain surface chemistry and desorption are needed to make progress here. Thus, it would be inter- esting to add grain chemistry, and in particular grain nitrile chemistry, in the Meudon PDR code to test how the results presented here would be affected. Further- more, other parameters would be worth testing in fu- ture model developments, such as the impact of stellar X-ray irradiation on disk chemistry which could affect its ionization (e.g. Glassgold et al. 1997; Rab et al. 2018; Waggoner & Cleeves 2019), and the sensitivity of disk chemistry to grain sizes (e.g. Wakelam et al. 2019) as smaller grains provide a higher surface area relative to their volume and thus more reaction sites and also tend to have temperatures closer to that of the gas. In the meantime we note that the strong impact of C/O on ni- trile chemistry may enable us to use nitriles to constrain this important parameter in disks and PDR analogs. ACKNOWLEDGMENTS The authors would like to thank the anonymous ref- eree for constructive suggestions and comments. R. LG. also thanks Tom J. Millar for useful discussion. This work was supported by an award from the Simons Foun- dation (SCOL # 321183, KO). M. T. B. acknowledges support from the Noland Internship at Caltech. Software: Pandas (McKinney 2010), Matplotlib (Hunter 2007), NumPy (van der Walt et al. 2011), SciPy (Jones et al. 2001 -- ), Meudon PDR Code (Le Bourlot et al. 1993; Le Petit et al. 2006; Le Bourlot et al. 2012), Nautilus-v1.1 (Hersant et al. 2009; Ruaud et al. 2016; Wakelam et al. 2016). The role of C/O in nitrile astrochemistry in PDRs and planet-forming disks 13 APPENDIX A. CRITICAL CHEMICAL REACTIONS ADDED TO THE MEUDON PDR CODE NETWORK Table A1. Rates of the critical chemical reactions discussed in this study. Chemical reactions α (cm3 s−1) β γ k rate type(a) Main formation and destruction pathways for CH3CNH+ & CH3CN HCN → CH3CNH+ HCN → CH3CNH+ HNC → CH3CNH+ e -- e -- e -- e -- hν HNC HCN CN → → → → → H H HCN → CH3NCH+ HCN → CH3NCH+ e -- e -- hν HCN CN → → → H hν hν hν H CH3CN CH3 CH3 CH3 H2CCN 7.20(-9) 8.00(-11) 9.00(-9) 8.00(-8) 1.30(-7) 6.00(-8) 6.00(-8) 2.95(-9) -0.50 -3.00 -0.50 -0.50 -0.50 -0.50 -0.50 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 3.07 hν hν CH3NC CH3 CH3 1.80(-9) 2.00(-11) 1.30(-7) 6.00(-8) 2.95(-9) -0.50 -3.00 -0.50 -0.50 0.00 0.00 0.00 0.00 0.00 3.07 Main formation and destruction pathways for CH3NCH+ & CH3NC + + + CH3 CH3 CH3 CH3CNH+ CH3CNH+ CH3CNH+ CH3CNH+ CH3CN + + CH3 CH3 CH3NCH+ CH3NCH+ CH3NC N C3N C3N C3N HC3NH+ C4H C N C2H2 HC3N T range Ref.(b) Rate (K) 10 − 299 300 − 800 10 − 300 10 − 300 10 − 300 10 − 300 10 − 300 -- 10 − 299 300 − 800 10 − 300 10 − 300 -- 10 − 300 -- 10 − 298 10 − 300 10 − 800 10 − 300 10 − 300 10 − 800 10 − 280 -- uncertainties(c) F0 = 10, g = 0 F0 = 2, g = 0 F0 = 10, g = 0 F0 = 3, g = 0 F0 = 3, g = 0 F0 = 3, g = 0 F0 = 3, g = 0 F0 = 1.3, g = 0 [1], [2] [2], [3] [4] [4] [4] [4] [4] [5] [1], [2] [2], [3] [4]((cid:63)) [4]((cid:63)) [5]((cid:63)) F0 = 10, g = 0 F0 = 2, g = 0 F0 = 3, g = 0 F0 = 3, g = 0 F0 = 1.3, g = 0 [4] [6] [7] [4] [8] [4] [4] [8] [6] [5] F0 = 3, g = 0 F0 = 2, g = 0 F0 = 3, g = 2.97 F0 = 3, g = 0 F0 = 1.4, g = 0 F0 = 3, g = 0 F0 = 3, g = 0 F0 = 3, g = 500 F0 = 1.25, g = 0 F0 = 1.3, g = 0 (1) (1) (1) (2) (2) (2) (2) (3) (1) (1) (2) (2) (3) (2) (3) (2) (2) (2) (2) (2) (2) (2) (3) c-C3H → → hν → O → C H C2 CO CN Main formation and destruction pathways for C3N C3N CN CCN C3 1.10(-10) 5.00(-10) 1.00(-10) 2.40(-10) 0.17 0.00 0.00 0.00 0.00 1.80 0.00 0.00 e -- N → → H2CCN → c-C3H2 → → → CN hν Main formation and destruction pathways for HC3N H C H H H CN HC3N HC3N HC3N HC3N HC3N C2H 6.00(-7) 7.00(-11) 1.00(-10) 1.00(-11) 2.72(-10) 7.13(-9) -0.58 0.17 0.00 0.00 -0.52 0.00 0.00 0.00 0.00 2000.00 19.00 2.59 Note -- Numbers in parentheses are power of 10 (a) Rate formulae: (1) Radiative associations and (2) bimolecular gas-phase reactions are computed from the same rate formula k = α(T /300)βe−γ/T , (3) photo-reactions for which the photo-cross-sections are unknown are computed with k = α , with χ− and χ+ scaling factors of the radiation field with respect to that of Draine on the left and right side of the cloud, respectively (Le Petit et al. 2006); (b) [1] Herbst (1985); [2] Anicich (2003); [3] Harada et al. (2010); [4] Loison et al. (2014); [5] Heays et al. (2017); [6] OSU database; [7] Smith database; [8] Loison et al. (2017); ((cid:63)) Assumed from the referenced publication; (c) Rate uncertainties are from the KIDA database and are considered to follow a lognormal distribution, i.e. with a probability of ≈ 68% for the rate k to be in , k × F0] and g being an expansion parameter used to parameterize a possible temperature dependence of the the range [ k F0 ) with T0 = 300 K (see Wakelam et al. 2012, for more details). uncertainty, according to the formula F (T ) = F0 exp(g 1 χ−e−γAV + χ+e−γ(Amax V −AV)(cid:17) (cid:16) T − 1 T0 14 Le Gal et al. REFERENCES Abergel, A., Teyssier, D., Bernard, J. P., et al. 2003, A&A, Gerin, M., Goicoechea, J. R., Pety, J., & Hily-Blant, P. 410, 577 2009, A&A, 494, 977 Andrews, S. M., Rosenfeld, K. A., Kraus, A. L., & Wilner, D. J. 2013, ApJ, 771, 129 Anicich, V. G. 2003, JPL Publication 03-19, 1 Anicich, V. G., Sen, A. D., Huntress, Wesley T., J., & McEwan, M. J. 1995, JChPh, 102, 3256 Glassgold, A. E., & Langer, W. D. 1974, ApJ, 193, 73 Glassgold, A. E., Najita, J., & Igea, J. 1997, ApJ, 480, 344 Goicoechea, J. R., Pety, J., Gerin, M., Hily-Blant, P., & Le Bourlot, J. 2009, A&A, 498, 771 Goicoechea, J. R., Pety, J., Gerin, M., et al. 2006, A&A, Asplund, M., Grevesse, N., Sauval, A. J., & Scott, P. 2009, 456, 565 ARA&A, 47, 481 Gratier, P., Pety, J., Guzm´an, V., et al. 2013, A&A, 557, Beckwith, S. V. W., Sargent, A. I., Chini, R. S., & Guesten, A101 R. 1990, AJ, 99, 924 Grenier, I. A., Black, J. H., & Strong, A. W. 2015, Bergin, E. A., Du, F., Cleeves, L. I., et al. 2016, ApJ, 831, ARA&A, 53, 199 101 Bergner, J. B., Guzm´an, V. G., Oberg, K. I., Loomis, R. A., & Pegues, J. 2018, ApJ, 857, 69 Bergner, J. B., Oberg, K. I., Bergin, E. A., et al. 2019, ApJ, 876, 25 Bertin, M., Doronin, M., Fillion, J. H., et al. 2017a, A&A, 598, A18 Bertin, M., Doronin, M., Michaut, X., et al. 2017b, A&A, 608, A50 Calcutt, H., Fiechter, M. R., Willis, E. R., et al. 2018, A&A, 617, A95 Cernicharo, J., Kahane, C., Guelin, M., & Gomez-Gonzalez, J. 1988, A&A, 189, L1 Chapillon, E., Dutrey, A., Guilloteau, S., et al. 2012, ApJ, 756, 58 Chiang, E. I., & Goldreich, P. 1997, ApJ, 490, 368 Cleeves, L. I., Bergin, E. A., Qi, C., Adams, F. C., & Oberg, K. I. 2015, ApJ, 799, 204 Cleeves, L. I., Oberg, K. I., Wilner, D. J., et al. 2018, ApJ, 865, 155 Cuadrado, S., Goicoechea, J. R., Pilleri, P., et al. 2015, A&A, 575, A82 Dartois, E., Dutrey, A., & Guilloteau, S. 2003, A&A, 399, 773 Defrees, D. J., McLean, A. D., & Herbst, E. 1985, ApJ, 293, 236 Guilloteau, S., Dutrey, A., Pi´etu, V., & Boehler, Y. 2011, A&A, 529, A105 Guilloteau, S., Reboussin, L., Dutrey, A., et al. 2016, A&A, 592, A124 Guzm´an, V., Pety, J., Goicoechea, J. R., Gerin, M., & Roueff, E. 2011, A&A, 534, A49 Guzm´an, V. V., Pety, J., Goicoechea, J. R., et al. 2015, ApJL, 800, L33 Habart, E., Abergel, A., Walmsley, C. M., Teyssier, D., & Pety, J. 2005, A&A, 437, 177 Harada, N., Herbst, E., & Wakelam, V. 2010, Astrophysical Journal, 721, 1570 Hartmann, L., Ballesteros-Paredes, J., & Bergin, E. A. 2001, ApJ, 562, 852 Heays, A. N., Bosman, A. D., & van Dishoeck, E. F. 2017, A&A, 602, A105 Henning, T., Semenov, D., Guilloteau, S., et al. 2010, ApJ, 714, 1511 Herbst, E. 1985, Astrophysical Journal, 291, 226 Hersant, F., Wakelam, V., Dutrey, A., Guilloteau, S., & Herbst, E. 2009, A&A, 493, L49 Hogerheijde, M. R., Bergin, E. A., Brinch, C., et al. 2011, Science, 334, 338 Huang, J., Andrews, S. M., Dullemond, C. P., et al. 2018, arXiv e-prints, arXiv:1812.04041 Hunter, J. D. 2007, Computing in Science Engineering, 9, 90 Draine, B. T. 1978, ApJS, 36, 595 Dullemond, C. P., Dominik, C., & Natta, A. 2001, ApJ, Huntress, W. T., J., & Mitchell, G. F. 1979, ApJ, 231, 456 Indriolo, N., Neufeld, D. A., Gerin, M., et al. 2015, ApJ, 560, 957 Dutrey, A., Guilloteau, S., & Guelin, M. 1997, A&A, 317, L55 Elmegreen, B. G. 2000, ApJ, 530, 277 Fuente, A., Cernicharo, J., Ag´undez, M., et al. 2010, A&A, 524, A19 Fuente, A., Rodrıguez-Franco, A., Garcıa-Burillo, S., 800, 40 Jenkins, E. B. 2009, ApJ, 700, 1299 Jones, A. P., & Ysrad, N. 2019, arXiv e-prints, arXiv:1906.01382 Jones, E., Oliphant, T., Peterson, P., et al. 2001 -- , SciPy: Open source scientific tools for Python, , . http://www.scipy.org/ Martın-Pintado, J., & Black, J. H. 2003, A&A, 406, 899 Kastner, J. H., Qi, C., Gorti, U., et al. 2015, ApJ, 806, 75 The role of C/O in nitrile astrochemistry in PDRs and planet-forming disks 15 Kastner, J. H., Qi, C., Dickson-Vandervelde, D. A., et al. Plessis, S., Carrasco, N., Dobrijevic, M., & Pernot, P. 2012, 2018, ApJ, 863, 106 Icarus, 219, 254 Le Bourlot, J., Le Petit, F., Pinto, C., Roueff, E., & Roy, Plessis, S., Carrasco, N., & Pernot, P. 2010, JChPh, 133, F. 2012, A&A, 541, A76 134110 Le Bourlot, J., Pineau Des Forets, G., Roueff, E., & Flower, Powner, M. W., Gerland, B., & Sutherland, J. D. 2009, D. R. 1993, A&A, 267, 233 Le Gal, R., Herbst, E., Dufour, G., et al. 2017, A&A, 605, Nature, 459, 239 EP . http://dx.doi.org/10.1038/nature08013 A88 Le Gal, R., Hily-Blant, P., Faure, A., et al. 2014, A&A, 562, A83 Le Gal, R., Oberg, K. I., Loomis, R. A., Pegues, J., & Bergner, J. B. 2019, ApJ, 876, 72 Le Petit, F., Nehm´e, C., Le Bourlot, J., & Roueff, E. 2006, ApJS, 164, 506 Le Petit, F., Ruaud, M., Bron, E., et al. 2016, A&A, 585, A105 Loison, J.-C., Wakelam, V., & Hickson, K. M. 2014, MNRAS, 443, 398 Przybilla, N., Nieva, M.-F., & Butler, K. 2008, ApJL, 688, L103 Qi, C., Oberg, K. I., Wilner, D. J., & Rosenfeld, K. A. 2013, ApJL, 765, L14 Rab, C., Gudel, M., Woitke, P., et al. 2018, A&A, 609, A91 Remijan, A. J., Hollis, J. M., Lovas, F. J., Plusquellic, D. F., & Jewell, P. R. 2005, ApJ, 632, 333 Rimmer, P. B., Herbst, E., Morata, O., & Roueff, E. 2012, A&A, 537, A7 Rosenfeld, K. A., Andrews, S. M., Wilner, D. J., Kastner, J. H., & McClure, M. K. 2013, ApJ, 775, 136 Ruaud, M., Wakelam, V., & Hersant, F. 2016, MNRAS, Loison, J.-C., Ag´undez, M., Wakelam, V., et al. 2017, 459, 3756 MNRAS, 470, 4075 Loomis, R., Oberg, K. I., Andrews, S., & et al. 2019, submitted to ApJ Loomis, R. A., Cleeves, L. I., Oberg, K. I., et al. 2018, ApJ, 859, 131 Shakura, N. I., & Sunyaev, R. A. 1973, A&A, 24, 337 Sutherland, J. D. 2016, Angewandte Chemie International Edition, 55, 104. https://onlinelibrary.wiley.com/doi/ abs/10.1002/anie.201506585 van der Walt, S., Colbert, S. C., & Varoquaux, G. 2011, L´opez, A., Tercero, B., Kisiel, Z., et al. 2014, A&A, 572, Computing in Science and Engineering, 13, 22 A44 Mathis, J. S., Rumpl, W., & Nordsieck, K. H. 1977, ApJ, 217, 425 McKinney, W. 2010, in Proceedings of the 9th Python in Science Conference, ed. S. van der Walt & J. Millman, 51 -- 56 Millar, T. J., & Herbst, E. 1990, MNRAS, 242, 92 Millar, T. J., Leung, C. M., & Herbst, E. 1987, A&A, 183, 109 Nath, B. B., & Biermann, P. L. 1994, MNRAS, 270, L33 Nguyen, T., Fourr´e, I., Favre, C., et al. 2019, A&A, 628, A15 Oberg, K. I., Guzm´an, V. V., Furuya, K., et al. 2015, Nature, 520, 198 Patel, B. H., Percivalle, C., Ritson, D. J., Duffy, C. D., & Sutherland, J. D. 2015, Nature Chemistry, 7, 301 EP . https://doi.org/10.1038/nchem.2202 van Dishoeck, E. F., & Blake, G. A. 1998, ARA&A, 36, 317 Vasyunin, A. I., Semenov, D., Henning, T., et al. 2008, ApJ, 672, 629 Vasyunin, A. I., Sobolev, A. M., Wiebe, D. S., & Semenov, D. A. 2004, Astronomy Letters, 30, 566 Vigren, E., Kami´nska, M., Hamberg, M., et al. 2008, Physical Chemistry Chemical Physics (Incorporating Faraday Transactions), 10, 4014 Wagenblast, R., & Hartquist, T. W. 1989, MNRAS, 237, 1019 Waggoner, A. R., & Cleeves, L. I. 2019, arXiv e-prints, arXiv:1908.08048 Wakelam, V., Chaietillon, E., Dutrey, A., et al. 2019, MNRAS, 484, 1563 Wakelam, V., Herbst, E., Le Bourlot, J., et al. 2010, A&A, 517, A21 Wakelam, V., Herbst, E., & Selsis, F. 2006, A&A, 451, 551 Wakelam, V., Ruaud, M., Hersant, F., et al. 2016, A&A, Pety, J., Goicoechea, J. R., Hily-Blant, P., Gerin, M., & 594, A35 Teyssier, D. 2007, A&A, 464, L41 Wakelam, V., Selsis, F., Herbst, E., & Caselli, P. 2005, Pety, J., Teyssier, D., Foss´e, D., et al. 2005, A&A, 435, 885 A&A, 444, 883 Pety, J., Gratier, P., Guzm´an, V., et al. 2012, A&A, 548, Wakelam, V., Herbst, E., Loison, J. C., et al. 2012, ApJS, A68 Pi´etu, V., Dutrey, A., & Guilloteau, S. 2007, A&A, 467, 163 199, 21 Whittet, D. C. B. 2010, ApJ, 710, 1009 Williams, J. P., & Best, W. M. J. 2014, ApJ, 788, 59
0901.1843
1
0901
2009-01-13T17:57:02
Mercury's capture into the 3/2 spin-orbit resonance including the effect of core-mantle friction
[ "astro-ph.EP" ]
The rotation of Mercury is presently captured in a 3/2 spin-orbit resonance with the orbital mean motion. The capture mechanism is well understood as the result of tidal interactions with the Sun combined with planetary perturbations. However, it is now almost certain that Mercury has a liquid core, which should induce a contribution of viscous friction at the core-mantle boundary to the spin evolution. This last effect greatly increases the chances of capture in all spin-orbit resonances, being 100% for the 2/1 resonance, and thus preventing the planet from evolving to the presently observed configuration. Here we show that for a given resonance, as the chaotic evolution of Mercury's orbit can drive its eccentricity to very low values during the planet's history, any previous capture can be destabilized whenever the eccentricity becomes lower than a critical value. In our numerical integrations of 1000 orbits of Mercury over 4 Gyr, the spin ends 99.8% of the time captured in a spin-orbit resonance, in particular in one of the following three configurations: 5/2 (22%), 2/1 (32%) and 3/2 (26%). Although the present 3/2 spin-orbit resonance is not the most probable outcome, we also show that the capture probability in this resonance can be increased up to 55% or 73%, if the eccentricity of Mercury in the past has descended below the critical values 0.025 or 0.005, respectively.
astro-ph.EP
astro-ph
Mercury’s capture into the 3/2 spin-orbit resonance including the effect of core-mantle friction Alexandre C. M. Correiaa,b,∗, Jacques Laskarb bAstronomie et Syst`emes Dynamiques, IMCCE-CNRS UMR8028, Observatoire de Paris, UPMC, 77 Av. Denfert-Rochereau, 75014 Paris, France aDepartamento de F´ısica, Universidade de Aveiro, Campus de Santiago, 3810-193 Aveiro, Portugal 9 0 0 2 n a J 3 1 . ] P E h p - o r t s a [ 1 v 3 4 8 1 . 1 0 9 0 : v i X r a Abstract The rotation of Mercury is presently captured in a 3/2 spin-orbit resonance with the orbital mean motion. The capture mechanism is well understood as the result of tidal interactions with the Sun combined with planetary perturbations (Goldreich and Peale, 1966; Correia and Laskar, 2004). However, it is now almost certain that Mercury has a liquid core (Margot et al., 2007) which should induce a contribution of viscous friction at the core-mantle boundary to the spin evolution. According to Peale and Boss (1977) this last effect greatly increases the chances of capture in all spin-orbit resonances, being 100% for the 2/1 resonance, and thus preventing the planet from evolving to the presently observed configuration. Here we show that for a given resonance, as the chaotic evolution of Mercury’s orbit can drive its eccentricity to very low values during the planet’s history, any previous capture can be destabilized whenever the eccentricity becomes lower than a critical value. In our numerical integrations of 1000 orbits of Mercury over 4 Gyr, the spin ends 99.8% of the time captured in a spin-orbit resonance, in particular in one of the following three configurations: 5/2 (22%), 2/1 (32%) and 3/2 (26%). Although the present 3/2 spin-orbit resonance is not the most probable outcome, we also show that the capture probability in this resonance can be increased up to 55% or 73%, if the eccentricity of Mercury in the past has descended below the critical values 0.025 or 0.005, respectively. Key words: Mercury, spin dynamics, tides, core-mantle friction, resonance 1. Introduction Mercury’s present rotation is locked in a 3/2 spin-orbit res- onance, with its spin axis nearly perpendicular to the orbital plane (Pettengill and Dyce, 1965). The stability of this ro- tation results from the solar torque on Mercury’s quadrupo- lar moment of inertia, combined with an eccentric orbit: the axis of minimum moment of inertia is always aligned with the direction to the Sun when Mercury is at the perihe- lion of its orbit (Colombo, 1965; Goldreich and Peale, 1966; Counselman and Shapiro, 1970). The initial rotation of Mer- cury was presumably faster than today, but tidal dissipation along with core-mantle friction brought the planet rotation to the present configuration, where capture can occur. However, the exact mechanism on how this state initially arose is not com- pletely understood. In their seminal work, Goldreich and Peale (1966) have shown that since the tidal strength depends on the planet’s ro- tation rate, it creates an asymmetry in the tidal potential that allows capture into spin-orbit resonances. They also computed the capture probability into these resonances for a single cross- ing, and found that for the present eccentricity value of Mer- cury (e = 0.206), and unless one uses an unrealistic tidal model with constant torques (which cannot account for the observed damping of the planet’s libration), the probability of capture ∗Corresponding author, Email addresses: [email protected] (Alexandre C. M. Correia) Preprint submitted to Icarus into the present 3/2 spin-orbit resonance is on the low side, at most about 7%, which remained somewhat unsatisfactory. Goldreich and Peale (1967) nevertheless pointed out that the probability of capture could be greatly enhanced if a planet has a molten core. In 1974, the discovery of an intrinsic magnetic field by the Mariner 10 spacecraft (Ness et al., 1974), seemed to imply the existence of a conducting liquid core and conse- quently an increment in the capture probability in the 3/2 reso- nance. However, according to Goldreich and Peale (1967), this also increases the capture probability in all the previous reso- nances. Peale and Boss (1977) indeed remarked that only very specific values of the core viscosity allow to avoid the 2/1 reso- nance and permit the capture in the 3/2 configuration. More recently, Correia and Laskar (2004) (hereafter denoted by Paper I) have shown that as the orbital eccentricity of Mer- cury is varying chaotically, from near zero to more than 0.45, the capture probability is substantially increased. Indeed, when the large eccentricity variation is factored into the capture, the rotation rate of the planet can be accelerated, and the 3/2 reso- nance could have been crossed many times in the past. Perform- ing a statistical study of the past evolutions of Mercury’s orbit, over 1000 cases, it was demonstrated that capture into the 3/2 spin-orbit resonant state is in fact, and without the need of spe- cific core-mantle effect, the most probable final outcome of the planet’s evolution, occurring about 55.4% of the time. In con- trast, because the eccentricity can decrease to near zero, all res- onances except the 1/1 become unstable, allowing the planet to escape from resonance. This mechanism suggests that in pres- April 21, 2014 ence of core-mantle friction the planet can escape to a previous capture in the 2/1 or higher order spin-orbit resonances. The present paper continues the work started in Paper I. In addition to the effects of tides and planetary perturbations, we will consider here also the core-mantle friction effect as de- scribed by Goldreich and Peale (1967), since the presence of a liquid core inside Mercury is now confirmed by radar obser- vations (Margot et al., 2007). In the next section we give the averaged equations of motion in a suitable form for simula- tions of the long-term variations of Mercury’s spin, including the resonant motion, viscous tidal effects, core-mantle coupling and planetary perturbations. In section 3 we discuss the conse- quences of each effect into the spin evolution and evaluate the capture probabilities in resonance. Finally, in last section we perform some numerical simulations to illustrate the different effects described in section 3. 2. Equations of motion We will adopt here a model for Mercury which is an exten- sion of the model from Poincar´e (1910) of a perfect incom- pressible and homogeneous liquid core with moments of inertia Ac = Bc < Cc inside an homogeneous rigid body with moments of inertia Am ≤ Bm < Cm, supported by the reference frame (~i,~j, ~k), fixed with respect to the planet’s figure. Tidal dissi- pation and core-mantle friction drive the obliquity close to zero (Yoder, 1997; Correia et al., 2003). Since we are only interested here in the study of the final stages of evolution (where capture in spin-orbit resonance may occur), we will neglect the effect of the small obliquity variations on the equations of motion. Moreover, since for a long-term study we are not interested in diurnal nutations, we can average over fast rotation angles and merge the axis of principal inertia and the axis of rotation (Bou´e and Laskar, 2006). Therefore, the mantle rotation rate is simply given by ω = θ + ϕ, where θ is the rotation angle and ϕ the precession angle. 2.1. Precession torque The gravitational potential V generated at a generic point of the space ~r is given by (e.g. Tisserand, 1891; Smart, 1953): V(~r) = − Gm r + + G(B − A) r3 G(C − A) r3 P2(~ur ·~j) P2(~ur · ~k) , (1) where terms in (R/r)3 were neglected. G is the gravitational constant, m the mass of Mercury, ~ur = ~r/r and P2(x) = (3x2 − 1)/2 is the Legendre polynomial of degree two. When interacting with the Sun’s mass, m , the spin of Mercury will undergo important changes. The middle term in the above po- tential will be responsible for a libration in the spin, while the last term causes the spin axis ~k to precess around ~K, the nor- mal to the orbit. Since we are only interested in the study of the long term motion, we will average the potential over the rotation angle θ and the mean anomaly M, after expanding the ⊙ true anomaly v in series of the eccentricity e and mean anomaly. However, when the rotation rate ω and the mean motion n = M are close to resonance (ω ≃ pn, for a semi-integer1 value p), we must retain the terms with argument (2θ − 2pM) in the ex- pansion cos(2θ − 2v) r3 = 1 a3 +∞ Xp=−∞ H(p, e) cos(2θ − 2pM) , (2) where a is the semi-major axis of the planet’s orbit and the func- tion H(p, e) is a power series in e (Tab. 1). The exact averaged contributions to the spin are given in a suitable form for our study by expression (15) in Correia (2006)2. For zero obliquity we have: dω dt β cm = − H(p, e) sin 2(θ + ϕ − pM − ) , (3) where is the longitude of the perihelion, cm = Cm/C = 0.55 (Margot et al., 2007), and β = 3Gm 2a3 ⊙ B − A C ≃ 3 2 n2 B − A C . (4) The Mariner 10 flyby of Mercury provided information on the internal structure of the planet, though subject to some uncertainty. For the gravity field it has been measured J2 = (6.0 ± 2.0)× 10−5 and C22 = (1.0 ± 0.5)× 10−5 (Anderson et al., 1987). Modeling the interior structure of Mercury, it has been estimated for the structure constant ξ = C/(mR2) ≃ 0.3359 (Spohn et al., 2001). Thus, for the moments of inertia we com- pute (B − A)/C = 4 C22/ξ ≃ 1.2 × 10−4. 2.2. Tidal torques Tidal effects arise from differential and inelastic deforma- tions of Mercury due to the gravitational effect of the Sun. Their contributions to the spin variations are based on a very gen- eral formulation of the tidal potential, initiated by George H. Darwin (1880). The distortion of the planet gives rise to a tidal potential, Vg(~r, ~r′) = −k2 Gm r !3 R R ⊙ 3 (cid:18) R r′(cid:19) P2(cid:0)~ur · ~ur′(cid:1) . (5) where R is the average radius of the planet and ~r′ the radial distance from the planet’s center to the Sun. In general, imper- fect elasticity will cause the phase angle of Vg to lag behind the perturbation (Kaula, 1964), because there is is a time delay ∆t between the perturbation of the Sun and the maximal defor- mation of Mercury. During that time, the planet rotates by an angle ω∆t, while the Sun also changes its position. Assuming a constant time delay allows us to linearize the tidal potential and simplify the tidal equations (Mignard, 1979, 1980; Hut, 1981). 1We have retained the use of semi-integers for better comparison with pre- vious results. 2There is a misprint in the sign of φ in Correia (2006). 2 Table 1: Coefficients of H(p, e) to e7. The exact expression of these coefficients is given by H(p, e) = 1 r(cid:17)3 π R π 0 (cid:16) a exp(i 2ν) exp(i 2pM) dM. p 1/1 3/2 2/1 5/2 3/1 7/2 4/1 9/2 H(p, e) 1 − 7 e 2 5 2 e2 − 17 e2 2 + 123 16 e3 − 845 48 e3 13 16 e4 + 115 6 e4 − 533 16 e4 − 489 128 + e5 32525 768 e5 − 228347 3840 e5 35 288 e6 − 601 48 + e6 13827 160 e6 − 73369 720 e6 1763 2048 e7 208225 6144 e7 3071075 18432 e7 12144273 71680 e7 For zero obliquity, the averaged contributions to the spin are then given by: dω dt = − where Ω(e) K cm  ω n − N(e) , 1 + 3e2 + 3e4/8 (1 − e2)9/2 , 1 + 15e2/2 + 45e4/8 + 5e6/16 Ω(e) = N(e) = (1 − e2)6 m (cid:19)(cid:18) R a(cid:19) ⊙ 3 , K = n2 3 k2 ξ Q (cid:18) m (6) (7) (8) (9) , and Q−1 = n∆t. This tidal model is particularly adapted to de- scribe the planets behavior in slow rotating regimes (ω ∼ n), which is the case of Mercury during the spin-orbit resonance crossing. In the present work we will adopt k2 = 0.4 and Q = 50, which yields K = 2.2 × 10−5 yr−2. This choice is somewhat arbitrary, but based on the parameter values of the other terrestrial planets (Goldreich and Soter, 1966). 2.3. Core-mantle friction effect The Mariner 10 flyby of Mercury revealed the presence of an intrinsic magnetic field, which is most likely due to motions in a conducting fluid core (for a review see Ness, 1978). Subsequent observations made with Earth-based radar provided strong evi- dence that the mantle of Mercury is decoupled from a core that is at least partially molten (Margot et al., 2007). 3 If there is slippage between the liquid core and the man- tle, a second source of dissipation of rotational energy results from friction occurring at the core-mantle boundary. Indeed, because of their different shapes and densities, the core and the mantle do not have the same dynamical ellipticity and the two parts tend to precess at different rates (Poincar´e, 1910). This tendency is more or less counteracted by different inter- actions produced at their interface: the torque of non-radial in- ertial pressure forces of the mantle over the core provoked by the non-spherical shape of the interface; the torque of the vis- cous friction (or turbulent) between the core and the mantle; and the torque of the electromagnetic friction, caused by the in- teraction between electrical currents of the core and the bottom of the magnetized mantle. The two types of friction torques (viscous and electromagnetic) depend on the differential rota- tion between the core and the mantle and can be expressed by a single effective friction torque, Υ (Mathews and Guo, 2005; Deleplace and Cardin, 2006): Υ = Ccκ δ ; δ = ω − ωc , (10) where κ is an effective coupling parameter, ωc the core’s ro- tation rate and cc = Cc/C = 1 − cm = 0.45 (Margot et al., 2007). According to Mathews and Guo (2005) we may write κ ≃ 2.62 √νω/Rc, where Rc is the core radius and ν is the effective kinematic viscosity of the core. Adopting Rc/R = 0.77 ± 0.04 (Spohn et al., 2001) and ν = 10−6m2s−1, we com- pute for the 3/2 resonance: κ = 5 × 10−5 yr−1. The uncertainty over ν is very large, according to Lumb and Aldridge (1991) it can cover about 13 orders of magnitude. As in former studies on planetary evolution (Yoder, 1997; Correia and Laskar, 2001, 2003), we used the same value as the best estimated so far for the Earth (Gans, 1972; Poirier, 1988). Another important consequence of core-mantle friction is to tilt the equator of the planet to the orbital plane, which results in a secular decrease of the obliquity (Rochester, 1976; Correia, 2006), reinforcing the previous assumption that the obliquity remains close to zero during the last evolutionary stages. Since the core and the mantle are decoupled, we need a dif- ferential equation for each rotation rate, the coupling equation being given by the friction torque (Peale, 2005; Correia, 2006): dω dt ccκ cm = − δ and dωc dt = κ δ . (11) 2.4. Planetary perturbations When considering the perturbations of the other planets, the eccentricity of Mercury is no longer constant, but undergoes strong chaotic variations in time (Laskar, 1990, 1994, 2008). These variations can be modeled using the averaging of the equations for the motion of the Solar System, that have been compared to recent numerical integrations, with very good agreement (Laskar et al., 2004a,b). These equations are ob- tained by averaging the equations of motion over the fast angles that are the mean longitudes of the planets. The averaging of the equation of motion is obtained by expanding the perturbations of the Keplerian orbits in Fourier series of the angles, where the coefficients themselves are expanded in series of the eccentric- ities and inclinations. This averaging process was conducted in a very extensive way, up to second order with respect to the masses, and through degree five in eccentricity and inclination, leading to truncated secular equations of the Solar System of the form dw dt = i [Γw + Φ3(w, ¯w) + Φ5(w, ¯w)] , (12) where w is the column vector (z1, ..., z8, ζ1, ..., ζ8), with z j = e j exp(i j), ζ j = sin(I j/2) exp(iΩ j) and and Ω, respectively the longitude of the perihelion and node. The 16 × 16 matrix Γ is the linear Lagrange-Laplace system, while Φ3(w, ¯w) and Φ5(w, ¯w) gather terms of degree 3 and 5 respectively. The system of equations thus obtained contains some 150000 terms, but its main frequencies are now the precession frequen- cies of the orbits of the planets. The full system can thus be numerically integrated with a very large step-size of 250 years. Contributions due to the secular perturbation of the Moon and general relativity are also included (for more details and refer- ences see Laskar, 1990, 1996; Laskar et al., 2004a). This secular system is then simplified and reduced to about 50000 terms, after neglecting terms of very small value (Laskar, 1994). Finally, a small correction of the terms appearing in Γ (Eq.12), after diagonalization, is performed in order to adjust the linear frequencies, in a similar way as it was done by Laskar (1990). The secular solutions are very close to the direct nu- merical integration La2004 (Laskar et al., 2004a,b) over about 35 Myr, the time over which the direct numerical solution itself is valid because of the imperfections of the model. It is thus legitimate to investigate the diffusion of the orbital motion of Mercury over long times using the secular equations. Over several millions years, the eccentricity of Mercury presents significant variations (Fig.1) that need to be taken into account in the resonance capture process (Paper I). 3. Spin evolution and capture probabilities The evolution of the spin for zero obliquity is completely described when we put together the contributions from the dif- ferent effects presented in section 2:   ω = P + T − Υ/Cm , ωc = Υ/Cc , (13) where P is the precession torques (Eq.3), T the tidal torque (Eq.6) and Υ the friction torque (Eq.10). In order to better study the capture probabilities in spin-orbit resonances, we will adopt a change of variables that is valid around each resonance p: γ = θ + ϕ − pM − , and thus, γ = ω − pn − and γ = ω − . (14) (15) In absence of planetary perturbations, for a Keplerian orbit, = 0 and thus γ = ω. Instead of the core rotation ωc we will also adopt the differential rotation δ = ω − ωc as variable. The above system of equations (13) becomes then: γ = P(γ) + T(γ) − ccκmδ − , δ = P(γ) + T(γ) − κmδ , (16) where κm = κ/cm. In general, we like to express γ as the sum of the precession torque P(γ) and a global dissipative torque D(γ) such that γ = −βm sin 2γ + D(γ) , where βm = β cm H(p, e) and D(γ) = −D0 V + µ , γ n (17) (18) (19) D0, V and µ being “constant” quantities. Indeed, under this linearized form we are able to estimate the capture probabili- ties using a simple expression derived by Goldreich and Peale (1966): Pcap = 2"1 + π 2 n ∆ω V µ#−1 , (20) where ∆ω is the maximal amplitude of libration in resonance, i.e., ∆ω = p2βm. 4 where χ = ccκn/[KΩ(e)]. It allows us easily to compute the capture probabilities using expression (20) provided that we are able to estimate correctly γc0 just before the planet crosses the resonance: , (21) Pcap = 2 1 + . (27) 3.1. Effect of tides Let us consider first the simplified case where the spin of the planet is only subject to the precession and tidal torques (Goldreich and Peale, 1966). Thus, we may use expression (17) to express γ, where D(γ) is given by expression (6): D(γ) = T = − K cm p − E(e) + Ω(e) γ n with E(e) = N(e)/Ω(e). The limit solution of the rotation for a constant value of the eccentricity is obtained when D(γ) = 0, that is, for γ/n = E(e) − p ⇔ ω/n = E(e). If the planet encounters a spin-orbit resonance in the way to the equilibrium position, capture may occur with a probability computed from expression (20): Pcap = 2(cid:20)1 + π 2 n ∆ω (p − E(e))(cid:21)−1 .  p − E(e) − χ 1 + χ π 2 n ∆ω γc0 n −1  According to expression (11) we have ωc = κ δ. Since κ ≪ ∆ω, the core is unable to follow the periodic variations in the mantle’s rotation rate. Thus, only the secular variations on the mantle will be followed by the core and we may write: ωc = ω − δ, where ω is the averaged rotation of the mantle and (22) δ ≃ e−κmtZ t t0 T eκmt′dt′ ≃ T κm . (28) Using the present value of the eccentricity of Mercury and cm = 1 we compute for the p = 3/2 resonance a probability of capture of about 7%, which is unsatisfactory given that this is the presently observed resonant state. 3.2. Effect of core-mantle friction We now add the effect of core-mantle friction to the effects already considered in the previous section. In this case we must take into account the rotation of the core, and the complete ro- tation rate evolution is given by system (16) with = 0. The general solution for the differential rotation in system (16) is given by: δ(t) = δ0 e−κmt + e−κmtZ t t0 (cid:2)P(γ) + T(γ)(cid:3) eκmt′dt′ , (23) where δ0 is the initial value of δ(t) for t = t0. In section 2.3 we estimate for the present rotation of Mercury 1/κm ≃ 104 yr, which can be seen as the time scale needed to damp the initial differential rotation δ0 to zero. Thus, for t ≫ 1/κm the evolution of the differential rotation is given only by the last term is ex- pression (23). It also means that for time intervals ∆t ≪ 1/κm, we can write for t < t0 + ∆t: δ(t) ≃ δ0 +Z t t0 γ dt′ , that is, δ(t) ≃ δ0 + γ − γ0 = γ − γc0 , (24) (25) where γc0 = γ0 − δ0 = ωc0 − pn, and ωc0 is the value of the core rotation for t = t0. This approximation is very useful because we can express γ by means of expression (17) with D(γ) = − = − K cm K cm Ω(e) Ω(e) p − E(e) + p − E(e) − χ γ n − γc0 n ccκ cm (cid:0)γ − γc0(cid:1) n + (1 + χ) γ , (26) 5 Just before crossing the pth spin-orbit resonance we have ω0 = pn + 2∆ω/π, and thus γc0 = ωc0 − pn = ω0 − δ − pn = 2 π ∆ω − T κm ≃ 2 π ∆ω + KΩ(e) κ (cid:2)p − E(e)(cid:3) . (29) When using the present eccentricity of Mercury and the val- ues of κ and K estimated in previous sections, we compute χ = 19.5. Substituting this into expression (27) we obtain a probability of capture of 100% in the 3/2 spin-orbit resonance. However, as noticed by Peale and Boss (1977), if we compute the probability for the 2/1 resonance we also get 100%. Thus, either the planet started its rotation below the 2/1 resonance, which is unlikely, or there must be another mechanism to avoid capture in the 2/1 and higher order resonances. 3.3. Effect of planetary perturbations The orbital eccentricity of Mercury undergoes important sec- ular perturbations from the other planets (Fig.1) and its contri- bution needs to be taken into account. The mean value of the eccentricity is ¯e = 0.198, slightly lower than the present value e ≃ 0.206, but we also observe a wide range for the eccen- tricity variations, from nearly zero to more than 0.45 (Paper I, Laskar, 2008). Even if some of these episodes do not last for a long time, they will allow additional capture into and escape from spin-orbit resonances. Moreover, the capture probabilities are also modified for different eccentricities: for the same reso- nance we can have zero or 100% of captures depending on the eccentricity value (Fig.2). For all resonances, the capture prob- ability is 100% whenever the eccentricity is close to the equi- librium value for the rotation rate, E(e) = N(e)/Ω(e) (Eq.21), but it tends to decrease as the eccentricity moves away from this equilibrium value. If the eccentricity is too high (or too low if the spin is increasing from lower rotation rates) some reso- nances cannot be reached and the probability of capture sud- denly drops to zero (Fig.2). Table 2: Critical eccentricity ec for the resonance p. If e < ec, the resonance p becomes unstable, and the solution may escape the resonance (Paper I). The critical eccentricity ec is obtained when βH(p, e) < K[Ω(e)p − N(e)]. p 5/1 9/2 4/1 7/2 3/1 5/2 2/1 3/2 1/1 ec 0.211334 0.174269 0.135506 0.095959 0.057675 0.024877 0.004602 0.000026 − For a non-constant eccentricity e(t), the limit solution of the rotation rate when D(γ) = 0 (Eq.17) is no longer ω/n = E(e), but more generally: 1 g(t) Z t 0 K cm (cid:20)N(e(τ)) − ccκ K δ(τ)(cid:21) g(τ) dτ , ω(t) = where g(t) = exp K cmn Z t 0 Ω(e(τ)) dτ! . (30) (31) The dissipation torques can thus drive the rotation rate sev- eral times across the same spin-orbit resonance, increasing the chances of capture. Another important consequence of a non-constant eccentric- ity is that all resonances but the 1/1 may become unstable. In- deed, the amplitude of the libration torque depends on the co- efficient H(p, e) (Eq.3), which goes to zero with the eccentric- ity, except for the 1/1 resonance (Tab.1). Whenever the am- plitude of the libration restoration torque becomes smaller than the amplitude of the dissipation torque, equilibrium in the spin- orbit resonance can no longer be sustained and the resonance is destabilized. Critical eccentricities for each resonance are listed in Table 2, obtained when the torques become equivalent (Eqs.3,6): Ω(e)p − N(e) H(p, e) = β K . (32) 4. Numerical simulations We will now use the dynamical equations established in sec- tion 2 to simulate the final evolution of Mercury’s spin by per- forming massive numerical integrations. The main goal is to illustrate the effects described in section 3, in particular the probabilities of capture and escape from spin-orbit resonances. Mercury geophysical models and parameters in use are those listed in section 2. We recall here the most uncertain values: k2 = 0.4, Q = 50 and ν = 10−6m2s−1. 6 Table 3: Capture probabilities in several spin-orbit resonances (in percentage). In the left panel (T only) we consider the effect of tides alone, while in the right panel (T + CMF) both tides and core-mantle friction effects are taken into account. The first column (Pcap) refers to the theoretical estimation given by expression (20), while the next column (num.) refers to the estimation obtained running a numerical simulation with 2000 close initial conditions, differing by π/1000 in the libration angle. Planetary perturbations are not considered and we used a constant eccentricity e = 0.206. p 5/1 9/2 4/1 7/2 3/1 5/2 2/1 3/2 T only Pcap (%) − − 0.1 0.1 0.3 0.7 1.8 7.7 num. (%) − − − 0.1 0.4 1.4 1.7 7.2 T + CMF Pcap (%) 1.6 3.1 5.9 11.4 22.6 46.6 100.0 100.0 num. (%) 0.3 1.3 4.8 10.9 22.8 46.2 100.0 100.0 4.1. Simulations without planetary perturbations Before considering the effect of planetary perturbations we can test numerically the theoretical estimates of the capture probability given by expressions (22) and (27). Since cap- ture in resonance is a statistical process we need to perform many integrations with slightly different initial conditions. For that purpose we ran 2000 simulations using a fixed eccentric- ity (e = 0.206), initial rotation period of 2 days, zero obliq- uity and different initial libration phase angles with step-size of π/1000 rad. Results are listed in Table 3. We can see that there is a good agreement between the theoretical previsions and the numerical estimation of the probabilities. As discussed in sections 3.1 and 3.2 the probability of capture when con- sidering only the effect of tides is very small (∼ 7% for the 3/2 resonance), while it becomes very important when core- mantle friction is added (100% for the 3/2 resonance). They are also in conformity with those obtained in the previous stud- ies by Goldreich and Peale (1967) and Peale and Boss (1977). As they all noticed, when the effect from core-mantle friction is considered, the probabilities of capture are greatly enhanced for all spin-orbit resonances. In particular, capture in the 2/1 res- onance also becomes 100%, preventing a subsequent evolution to the 3/2 resonance. 4.2. Inclusion of planetary perturbations When planetary perturbations are taken into account, the ec- centricity presents chaotic variations with many excursions to higher and lower values than today (Laskar, 1990, 1994, 2008, Paper I). It is then impossible to know its exact evolution at the time the planet first encountered the spin-orbit resonances. A statistical study with many different orbital solutions is the only possibility to get a global picture of the past evolution of the spin of Mercury. In Paper I we performed such a study by in- tegrating 1000 orbits over 4 Gyr in the past starting with very close initial conditions. This statistical study was only made possible by the use of the averaged equations for the motion of the Solar System (Laskar, 1990, 1994). In figure 3 we show five examples of the eccentricity evo- lution through the 4.0 Gyr. We choose some cases illustrative of the chaotic behavior, where we can see that the eccentricity can be as small as zero, but it can also reach values as high as 0.5. In some cases the eccentricity can remain within [0.1, 0.3] throughout the evolution, while in other cases it can span the whole interval [0, 0.5] (Laskar, 2008). Owing to the chaotic evolution, the density function of the 1000 solutions over 4 Gyr is a smooth function (Fig.4), well approximated by a Rice probability distribution (Laskar, 2008). The eccentricity excursions to higher values allow the planet to cross the 3/2 resonance several times, and thus increase the probability of capture. This behavior becomes very important if the evolution is driven by tidal friction alone. Even though the probability of capture in a single crossing of the 3/2 spin-orbit resonance is only around 7%, multiple crossings increase it up to 55% (Paper I). 4.3. Planetary perturbations with core-mantle friction In presence of an efficient core-mantle friction the multiple crossings of the 3/2 resonance are no longer needed, since the capture in this resonance after a single crossing is already 100% (Fig.2). Nevertheless, eccentricity excursions to lower values can destabilize the equilibrium in any spin-orbit resonance dif- ferent from the 1/1 (Tab.2). This effect was already present when the core-mantle friction was not considered (Paper I), but with small influence on the results, while here it becomes of capital importance. Indeed, it may allow the evasion from pre- vious captures in higher order resonances than the 3/2 and per- mit subsequent evolution to the present observed spin state. In order to check this new scenario, we have performed a sta- tistical study of the past evolutions of Mercury’s orbit, with the integration of the same 1000 orbits over 4 Gyr in the past used in Paper I. We now additionally include the effect of core-mantle friction as described by Goldreich and Peale (1967), i.e., we will consider the full dynamics of the spin governed by Eq.(13). Assuming an initial rotation period of Mercury of 10 h, we estimated that the time needed to de-spin the planet to the slow rotations would be about 300 million years. We will then start our integrations already in the slow-rotation regime, with a rota- tion period of 10 days (ω ≃ 8.8n), zero obliquity and a starting time of −4 Gyr, although these values are not critical. In Table 4 we show the amount of captures for each resonance at the end of the simulations (column “final”). We also list the resonances in which the spin was first captured before being destabilized (col- umn “1stcap.”) and we recall the results obtained for a constant eccentricity (e = 0.206) and in Paper I, with a model without core-mantle friction. After running 1000 trajectories we observe that the spin of Mercury preferably chooses one of the three final configura- tions: 5/2, 2/1 or 3/2 (Tab.4). With 26% of captures, the present configuration no longer represents the most probable final out- come, as it was in absence of core-mantle friction (Paper I). However, it is still among the most probable scenarios, the al- ternatives receiving comparable amounts of captures (22% and Table 4: Capture probabilities in several spin-orbit resonances (in percentage). We performed a statistical study of the past evolutions of Mercury’s spin, with the integration of 1000 orbits over 4 Gyr, a initial rotation period of 10 days and zero obliquity. In the “1stcap.” column we list the resonances in which the spin was first captured (before being destabilized). In the “final” column we list the results after the full 4 Gyr of simulations. For comparison we also list the results obtained with a constant eccentricity e = 0.206 (“const.”) and the final results obtained in Paper I (“C&L04” ), with a model without core-mantle friction. p 6/1 11/2 5/1 9/2 4/1 7/2 3/1 5/2 2/1 3/2 1/1 none const. (%) − − 0.3 1.3 4.7 10.3 19.0 29.8 34.6 − − − number of captures 1stcap. (%) 0.1 0.4 1.3 2.7 5.3 8.7 15.5 26.5 31.2 8.1 0.2 − final (%) − − − − − 4.7 11.6 22.1 31.6 25.9 3.9 0.2 C&L04 (%) − − − − − − − − 3.6 55.4 2.2 38.3 32% respectively for the 5/2 and the 2/1 resonances). The 5/2 and the 2/1 spin-orbit resonances benefit from the fact that the planet must cross them first. On the other hand, the 3/2 reso- nance is more stable and the chances of capture are higher when crossed. Since the eccentricity of Mercury 4.0 Gyr ago can be around 0.4 (e.g. Fig.3), at the moment of the first encounter with the spin-orbit resonances, capture in resonances as high as the 6/1 can occur (Fig.2). Because the probability of capture is small and because there are not many orbital solutions reaching such high values for the eccentricity, we only count about 10% of captures in resonances above or equal to the 4/1 (Tab.4). Once captured, these equilibria can be maintained as long as the ec- centricity remains above the respective critical values (Tab.2). Contrary to the results predicted for a constant eccentricity (Tab.4), we also registered a few trajectories directly captured in the 3/2 resonance just after the first passage through the reso- nance area. When we used the present value of the eccentricity (e = 0.206), the 3/2 resonance could not be attained because for that value capture probability in the 2/1 resonance is 100% (Fig.2). However, for eccentricity values lower than about 0.19, the probability of capture in this resonance decreases, as well as for higher order resonances. For instance, when the eccentricity is 0.09, capture in the 2/1 resonance drops to 50%. Thus, since the eccentricity of Mercury is varying, it may happen that about 4.0 Gyr ago its value was much lower than today and the spin managed to avoid all the spin-orbit resonances higher than the 3/2 and was directly captured in the present observed config- uration. We estimate nevertheless that the probability for this scenario to occur is very low, only about 8% (Tab.4). 7 4.4. Critical eccentricities Over 4.0 Gyr of evolution the eccentricity has many chances of experiencing a period of very small values (e.g. Fig.3). Even when a period of low eccentricity does not last for a long time, a single passage of the eccentricity below a critical value (Tab.2) can be enough to destabilize the corresponding spin-orbit reso- nance. All orbital solutions were generated starting from initial con- ditions close to the present values (see Laskar, 2008), and there- fore converge to the same final evolution. The eccentricity behavior is thus identical for the last 50-60 Myr (Fig.1), be- fore which the chaotic diffusion dominates (Fig.5). During the last 50 Myr the eccentricity certainly reached values lower than 0.13, thus the 4/1 and above spin-orbit resonances cannot rep- resent a possible final outcome for Mercury (e4/1 ≈ 0.136). For the 7/2 and lower order spin-orbit resonances, capture until the present day is not forbidden by the last 50 Myr of Mercury’s evolution, but depends on the true orbital evolution of the ec- centricity (Fig.5). The higher is the critical eccentricity, the lower is the probability of remaining trapped, because more or- bital solutions will come below this value. In figure 6 we plot the cumulative distribution of the minimal eccentricities attained for each one of the 1000 orbital solutions that we used. We also mark with straight lines the critical values of the eccentricity for each spin-orbit resonance (Tab.2) and use dots to represent the amount of captures obtained numerically for spin-orbit resonances that are still stable below each critical value of the eccentricity. Since a large amount of the orbital solutions experience at least one episode with an eccentricity below 0.05 (log10 e ≈ −1.3), about 84% of the final evolutions will end in the 5/2 spin-orbit resonance or lower (Tab.4). By comparing the eccentricity instability thresholds for each spin- orbit resonance with the amount of captures obtained numeri- cally below that resonance we see that there is a good agree- ment, suggesting a strong correlation between the orbital evo- lution of the eccentricity and the percentage of captures in each resonance. The reason why there is not full agreement between the two is because spin-orbit resonances below critical values of the eccentricity can also be attained by trajectories that escaped capture in higher order resonances, that is, they can be attained even if the eccentricity is never below the critical value for that resonance (Fig.2). When comparing the results after 4.0 Gyr with those after the first capture, we verify that the 5/2 resonance (and above) lose a significant amount of previously captured solutions. The amount of orbits captured in the 2/1 resonance remains roughly the same, because the number of trajectories quitting this reso- nance is more or less compensated by the incoming trajectories from higher order resonances. The 3/2 is the real winner of this transition process, as the amount of trajectories that end in this last configuration is about 4 times larger than it was initially. An identical scenario was already observed in Paper I, except that only a few captures occurred in spin-orbit resonances higher than the 2/1 and they were all subsequently destabilized (Tab.4). As in Paper I, we also noticed about 4% of the trajectories captured in the 1/1 spin-orbit resonance (Tab.4). Since the prob- ability of capture in the 3/2 spin-orbit resonance is almost 100% even for very low values of the eccentricity (Fig.2), the major possibility of evolving into the 1/1 resonance is by destabilizing the 3/2. This becomes a possibility if the eccentricity is almost zero, that is, for e < 3 × 10−5 (Tab.2). 4.5. Different scenarios of evolution The critical eccentricity needed to destabilize the 2/1 spin- orbit resonance is e2/1 ≈ 0.0046 (Tab.2). Whenever the orbital eccentricity is below this value, the spin will then evolve to- wards the 3/2 resonance or below (Fig.7). However, this is not the only possibility of achieving this last configuration if the planet was first captured in a higher-order resonance. Indeed, as discussed for the 1st capture column (Tab.4), if the eccen- tricity is lower than 0.19 at the time the planet crosses the 2/1 resonance (Fig.2), the chances of capture are lower than 100%, opening some space for subsequent evolution to the 3/2 reso- nance. For instance, for a previous capture in the 5/2 resonance, an eccentricity of e5/2 ≈ 0.025 will destabilize it and produce a capture probability of only 14.4% in the 2/1 resonance, i.e., when the 5/2 resonance is destabilized there is about 85% of chance of ending in the 3/2 present configuration (Fig.7). In order to exemplify the multitude of possible evolutionary scenarios, we performed another kind of experiment. Adopt- ing a particular orbital solution, which presents a gradual de- crease in the eccentricity (Fig.3a), we integrated close initial conditions for the spin. Since for this orbital solution the ec- centricity is high at the time of the first encounter, there is a great chance of capturing the spin in a spin-orbit resonance with p > 5/2. In figure 7 we plot the behavior of four trajectories, each one initially captured in a different spin-orbit resonance, when the eccentricity approaches a zone of very low values. As expected, the spin-orbit resonances are sequentially aban- doned as the eccentricity assumes small values. In particular, we observe that the resonances are quit immediately after the eccentricity is below the critical values listed in Table 2. After being destabilized, the spin can evolve directly to the present 3/2 configuration, or can be trapped in an intermediate spin- orbit resonance. In this example, the eccentricity become lower than e2/1 ≈ 0.0046 around −1.79 Gyr and captures in the 2/1 resonance become destabilized after that date. We purposely plot one situation, where the spin does not end in the 3/2 resonance, however. In this case, at the moment the eccentricity becomes lower than e2/1, the spin is still captured in the 5/2 resonance. This resonance logically becomes desta- bilized and the rotation rate decreases. Nevertheless, at the mo- ment the spin encounters the 2/1 resonance the eccentricity is again higher than e2/1, and therefore there is a chance of cap- ture in this resonance, preventing a subsequent evolution toward the 3/2 state (Fig.7). 4.6. Constraints on the orbital evolution We have seen in previous sections that there is an important correlation between the minimal eccentricity attained by Mer- cury through its orbital evolution and the probability of capture in a given resonance (Fig.6). The lower is the minimal eccen- tricity, the higher is the probability of achieving a low order 8 Table 5: Capture probabilities in spin-orbit resonances (in percentage), when the eccentricity descends below a given critical value (Tab.2). p 7/2 3/1 5/2 2/1 3/2 1/1 none e3/2 − − − 3.1 25.0 68.8 3.1 e2/1 − 2.0 3.2 5.4 73.2 15.4 0.7 e5/2 1.2 4.2 5.2 23.9 55.1 9.7 0.5 e3/1 2.6 7.2 16.0 30.5 37.2 6.1 0.3 e7/2 3.8 10.7 20.7 32.6 27.7 4.2 0.2 e4/1 4.7 11.6 22.1 31.6 25.9 3.9 0.2 spin-orbit resonance. In particular, each time the eccentricity descends below a given critical value for a spin-orbit resonance (Tab.2), the spin will evolve into a lower resonance. Since we know the distribution of the minimal eccentricities (Fig.6), we can estimate the probability of ending in a specific spin-orbit resonance given the value of the minimal eccentric- ity of a considered orbit, Pcap/ec. For that purpose we eliminate all the trajectories for which the minimal eccentricity is above the critical value (Tab.2) and then count the number of captures in each resonance for the remaining orbital solutions. Results are listed in Table 5. While for an arbitrary orbital solution the probability of capture in the present 3/2 spin-orbit resonance is only 25.9%, this value rises to 55.1% if we assume that the ec- centricity of Mercury was below e5/2 ≈ 0.025 at some time in the past, or even up to 73.2% if the eccentricity descends below e2/1 ≈ 0.0046. Results for the critical eccentricity e4/1 ≈ 0.136 are the same as the global results shown in Table 4, because the eccentricity of Mercury was below that value in the most recent 50 Myr, where the chaotic behavior is not significant (Fig.5). Notice also that there are always a few captures left in resonances above the corresponding critical value. This can be explained by the same effect described in the last paragraph of section 4.5 and illustrated in figure 7. Inversely, since we know that the rotation of Mercury is presently captured in the 3/2 spin-orbit resonance, we can esti- mate the probability for the eccentricity to have descended dur- ing its past evolution below a specific critical level, Pec. Using conditional probabilities, we have then Pec = Pcap/ec × Porb Pcap , (33) where Pcap/ec is the probability of ending in a specific spin-orbit resonance given the value of the minimal eccentricity (Tab.5), Porb the probability for the eccentricity to reach that minimal eccentricity (Fig.6) and Pcap the global capture probability in the specific spin-orbit resonance (Tab.4). Results for the 3/2 spin-orbit resonance are given in Table 6. These probabilities for the orbital evolution of the eccentricity are the same as if we select only the evolutions that finished in the 3/2 spin-orbit resonance and then look at the minimal eccentricity distribu- tion. From the above analysis we conclude that there is a strong probability that the eccentricity of Mercury reached very low values; in particular there is about a 77% chance that it de- Table 6: Probability for the eccentricity of Mercury to have descended below a specific critical level (Pec ) given that its rotation is captured in the 3/2 spin-orbit resonance today. Results for each critical eccentricity ec (Tab.2) are obtained from Eq.(33) with Pcap = 25.9% (Tab.4). P Pcap/ec Porb Pec e3/2 25.0 3.2 3.1 e2/1 73.2 27.1 76.6 e5/2 55.1 40.3 85.7 e3/1 37.2 64.5 92.6 e7/2 27.7 99.7 98.8 e4/1 25.9 100.0 100.0 scended below e2/1 ≈ 0.0046 (but only a 3% chance of going below e3/2 ≈ 0.00003). 5. Conclusions Due to the increasing evidence of a molten core inside Mer- cury (Ness et al., 1974; Margot et al., 2007), viscous friction at the core-mantle boundary is expected and its consequences to the spin must be taken into account. An important consequence is a considerable increase in the probability of capture for all spin-orbit resonances; in particular, for the 2/1 and the 3/2 it can reach 100% (Peale and Boss, 1977). Since it is believed that Mercury’s initial rotation was much faster than today, a destabilization mechanism is then required to allow the planet to escape from the 2/1 and higher order resonances and subse- quently evolve to the present observed 3/2 configuration. With the consideration of the chaotic evolution of the eccen- tricity of Mercury we show that such destabilization mechanism exists whenever the eccentricity becomes smaller than a critical value for each spin-orbit resonance (Tab.2). This mechanism was already described in Paper I, but becomes of capital impor- tance when core-mantle friction is taken into account. There are two main possibilities to evolve into the 3/2 configuration: • The eccentricity becomes lower than the critical value for the 2/1 spin-orbit resonance (e2/1 ≈ 0.005) and evolves into the 3/2. • The eccentricity becomes lower than the critical value for a higher order resonance than the 2/1, and then crosses this resonance with an eccentricity lower than e < 0.19. This allows a non zero probability of escaping the 2/1 res- onance, and subsequent evolution into the 3/2. The other mechanism of capturing in the 3/2 resonance de- scribed in Paper I, consisted in a returning to the 3/2 spin-orbit resonance after an increase in the eccentricity. This effect is not as important when we take into account core-mantle friction, since the most part of the trajectories are captured in resonance after a single passage. After running 1000 orbital solutions, starting from 4 Gyr in the past until they reached the present date, the spin of the planet was captured in a spin-orbit resonance 99.8% of the time. The main resonances to be filled and the respective probability were (Tab.4): P5/2 = 22.1%, P2/1 = 31.6%, P3/2 = 25.9% . (34) 9 Although in this case the present configuration no longer repre- sents the most probable final outcome, as it was in absence of core-mantle friction (Paper I), it is still among the most proba- ble scenarios. Moreover, if we assume that at some time in the past, the eccentricity of Mercury becomes lower than e5/2 ≈ 0.025 or e2/1 ≈ 0.005 respectively, the probability of reaching the 3/2 spin-orbit resonance rises to 55% and 73% respectively (Table 5). Given that Mercury is presently trapped in the 3/2 config- uration, we can also estimate that the eccentricity of Mercury has known at least one period of very low eccentricity during its past evolution, with about 86% and 77% of chances of being below e5/2 ≈ 0.025 and e2/1 ≈ 0.005 respectively (Table 6). The probability of capture in the 3/2 resonance can also be increased if the orbital eccentricity experiences more periods near zero. This can be achieved if we use direct integration of the Solar System instead of the averaged equations, because the true eccentricity is expected to undergo some additional small variations around the value obtained for the averaged equations (Laskar, 2008). Alternatively, the probability of capture in the 3/2 resonance can still be increased if we are able to increase the critical eccentricities that destabilize spin-orbit resonances (Tab.2). This can be achieved if the tidal dissipation is stronger (k2 > 0.4 and/or Q < 50) or if C22 < 1.0 × 10−5 (Eq.32). Lower values for the core effective viscosity, ν, will not change the critical eccentricities, but will decrease the amount of captures for all spin-orbit resonances. As a consequence, it becomes easier to escape from the capture in spin-orbit reso- nances, and all those trajectories that also escape the 3/2 reso- nance can be later trapped there when the eccentricity experi- ences a period with e > 0.325 (Paper I). We then believe that the true scenario for the evolution of the spin of Mercury may be somewhere between the scenario described here, with an ef- ficient core-mantle friction effect, and the scenario described in Paper I, for a total absence of core-mantle friction. In the future, different dissipative parameters and models could be tested as well as the effect of the obliquity, that was supposed to be zero in the present study. Acknowledgments The authors thank S.J. Peale for discussions. This work was supported by the Fundac¸ao Calouste Gulbenkian (Portu- gal), Fundac¸ao para a Ciencia e a Tecnologia (Portugal), and by PNP-CNRS (France). References Anderson, J. D., Colombo, G., Espsitio, P. B., Lau, E. L., Trager, G. B., Sep. 1987. The mass, gravity field, and ephemeris of Mercury. Icarus 71, 337– 349. Bou´e, G., Laskar, J., Dec. 2006. Precession of a planet with a satellite. Icarus 185, 312–330. Colombo, G., 1965. Rotational Period of the Planet Mercury. Nature 208, 575– 578. Correia, A. C. M., Dec. 2006. The core-mantle friction effect on the secular spin evolution of terrestrial planets. Earth Planet. Sci. Lett. 252, 398–412. Correia, A. C. M., Laskar, J., Jun. 2001. The four final rotation states of Venus. Nature 411, 767–770. Correia, A. C. M., Laskar, J., May 2003. Long-term evolution of the spin of Venus II. Numerical simulations. Icarus 163, 24–45. Correia, A. C. M., Laskar, J., Jun. 2004. Mercury’s capture into the 3/2 spin- orbit resonance as a result of its chaotic dynamics. Nature 429, 848–850. Correia, A. C. M., Laskar, J., N´eron de Surgy, O., May 2003. Long-term evo- lution of the spin of Venus I. Theory. Icarus 163, 1–23. Counselman, C. C., Shapiro, I. I., 1970. Spin-Orbit resonance of Mercury. Sym- posia Mathematica 3, 121–169. Darwin, G. H., 1880. On the secular change in the elements of a satellite re- volving around a tidally distorted planet. Philos. Trans. R. Soc. London 171, 713–891. Deleplace, B., Cardin, P., Nov. 2006. Viscomagnetic torque at the core mantle boundary. Geophys. J. Int. 167, 557–566. Gans, R. F., 1972. Viscosity of the Earth’s core. J. Geophys. Res. 77, 360–366. Goldreich, P., Peale, S., Aug. 1966. Spin-orbit coupling in the solar system. Astron. J. 71, 425–438. Goldreich, P., Peale, S., Jun. 1967. Spin-orbit coupling in the solar system. II. The resonant rotation of Venus. Astron. J. 72, 662–668. Goldreich, P., Soter, S., 1966. Q in the Solar System. Icarus 5, 375–389. Hut, P., Jun. 1981. Tidal evolution in close binary systems. Astron. Astrophys. 99, 126–140. Kaula, W. M., 1964. Tidal dissipation by solid friction and the resulting orbital evolution. Rev. Geophys. 2, 661–685. Laskar, J., Dec. 1990. The chaotic motion of the solar system - A numerical estimate of the size of the chaotic zones. Icarus 88, 266–291. Laskar, J., Jul. 1994. Large-scale chaos in the solar system. Astron. Astrophys. 287, L9–L12. Laskar, J., 1996. Large Scale Chaos and Marginal Stability in the Solar System. Celestial Mechanics and Dynamical Astronomy 64, 115–162. Laskar, J., Jul. 2008. Chaotic diffusion in the Solar System. Icarus 196, 1–15. Laskar, J., Correia, A. C. M., Gastineau, M., Joutel, F., Levrard, B., Robutel, P., Aug. 2004a. Long term evolution and chaotic diffusion of the insolation quantities of Mars. Icarus 170, 343–364. Laskar, J., Robutel, P., Joutel, F., Gastineau, M., Correia, A. C. M., Levrard, B., Dec. 2004b. A long-term numerical solution for the insolation quantities of the Earth. Astron. Astrophys. 428, 261–285. Lumb, L. I., Aldridge, K. D., 1991. On viscosity estimates for the Earth’s fluid outer core-mantle coupling. J. Geophys. Geoelectr. 43, 93–110. Margot, J. L., Peale, S. J., Jurgens, R. F., Slade, M. A., Holin, I. V., May 2007. Large Longitude Libration of Mercury Reveals a Molten Core. Science 316, 710–714. Mathews, P. M., Guo, J. Y., Feb. 2005. Viscoelectromagnetic coupling in precession-nutation theory. J. Geophys. Res. (Solid Earth) 110, B02402–16. Mignard, F., May 1979. The evolution of the lunar orbit revisited. I. Moon and Planets 20, 301–315. Mignard, F., Oct. 1980. The evolution of the lunar orbit revisited. II. Moon and Planets 23, 185–201. Ness, N. F., Mar. 1978. Mercury - Magnetic field and interior. Space Science Reviews 21, 527–553. Ness, N. F., Behannon, K. W., Lepping, R. P., Whang, Y. C., Schatten, K. H., Jul. 1974. Magnetic field observations near Mercury: Preliminary results from Mariner 10. Science 185, 153–162. Peale, S. J., Nov. 2005. The free precession and libration of Mercury. Icarus 178, 4–18. Peale, S. J., Boss, A. P., Aug. 1977. Spin-orbit constraint on the viscosity of a Mercurian liquid core. J. Geophys. Res. 82, 743–749. Pettengill, G. H., Dyce, R. B., 1965. A Radar Determination of the Rotation of the Planet Mercury. Nature 206, 1240. Poincar´e, H., 1910. Sur la pr´ecession des corps d´eformables. Bull. Astron. 27, 321–356. Poirier, J. P., Jan. 1988. Transport properties of liquid metals and viscosity of the earth’s core. Geophysical Journal 92, 99–105. Rochester, M. G., 1976. The secular decrease of obliquity due to dissipative core-mantle coupling. Geophys. J.R.A.S. 46, 109–126. Smart, W. M., 1953. Celestial Mechanics. London, New York, Longmans, Green. Spohn, T., Sohl, F., Wieczerkowski, K., Conzelmann, V., Dec. 2001. The in- terior structure of Mercury: what we know, what we expect from Bepi- Colombo. Plan. Space Sci. 49, 1561–1570. Tisserand, F., 1891. Trait´e de M´ecanique C´eleste (Tome II). Gauthier-Villars, Paris. 10 y t i c i r t n e c c e 0.32 0.30 0.28 0.26 0.24 0.22 0.20 0.18 0.16 0.14 0.12 -50 -40 -30 -20 -10 0 time (Myr) Figure 1: Evolution of Mercury’s eccentricity over 50 Myr in the past (Laskar et al., 2004a,b). Yoder, C. F., 1997. Venusian Spin Dynamics. In: Venus II: Geology, Geo- physics, Atmosphere, and Solar Wind Environment. pp. 1087–1124. 0.5 0.4 0.3 0.2 0.1 0 0.4 0.3 0.2 0.1 0 0.3 0.2 0.1 0 0.3 0.2 0.1 0 0.3 0.2 0.1 y t i c i r t n e c c e y t i c i r t n e c c e y t i c i r t n e c c e y t i c i r t n e c c e y t i c i r t n e c c e (a) (b) (c) (d) (e) 0 -4 -3.5 -3 -2.5 -2 -1.5 -1 -0.5 0 time [Gyr] Figure 3: Some examples of the possible variations of the eccentricity of Mer- cury through the past 4.0 Gyr. The eccentricity can be as small as zero, but it can also reach values as high as 0.5. In some cases the eccentricity can re- main within [0.1, 0.3] throughout the evolution, while in other cases it can span the whole interval [0, 0.5]. All these solutions converge to the known recent evolution of the planet’s orbit (Fig.5). Figure 4: Probability density function of Mercury’s eccentricity (Paper I). Val- ues are computed over 4 Gyr for the numerical integration of the secular equa- tions (Laskar et al., 2004a,b; Laskar, 2008) for 1000 close initial conditions (LA04). The mean value of the eccentricity is ¯e = 0.198. 11 Figure 2: Probability of capture in some spin-orbit resonances for different values of the eccentricity, under the effect of tides and core-mantle friction (Eq.27). The dashed line corresponds to a planet increasing its spin from slower rotation rates, while the solid line corresponds to a planet de-spinning from faster rotation rates. For all resonances, capture probability is 100% whenever the eccentricity is close to the equilibrium value for the rotation rate, ω/n = E(e) (Eq.21). It suddenly decays to zero when the equilibrium rotation rate falls outside the resonance width, i.e., the tidal evolution prevents the planet from crossing the resonance. 12 / / n n w w n n / / w w y t i c i r t n e c c e 5 5 4.5 4.5 4 4 3.5 3.5 3 3 2.5 2.5 2 2 1.5 1.5 1 1 4.5 4.5 4 4 3.5 3.5 3 3 2.5 2.5 2 2 1.5 1.5 1 1 0.3 0.25 0.2 0.15 0.1 0.05 0 -1.9 -1.88 -1.86 -1.84 time [Gyr] -1.82 -1.8 -1.78 Figure 7: Four possible final evolutions for the spin of Mercury. Adopting a particular orbital solution, which presents a gradual decrease in the eccentricity (Fig.3a), we integrate close initial conditions for the spin. We observe that spin-orbit resonances are quit immediately after the eccentricity is below the critical values listed in Table 2. After being destabilized, the spin can evolve directly to the present 3/2 configuration, or can be trapped in an intermediate spin-orbit resonance. We purposely left one situation where the spin does not end in the 3/2 resonance. At the moment the eccentricity becomes lower than e2/1 ≈ 0.0046, the spin is still captured in the 5/2 resonance. This resonance is then destabilized, but when the spin encounters the 2/1 resonance the eccentricity is already higher than e2/1. Thus, there is a chance of capture in this resonance, preventing a subsequent evolution toward the 3/2 state. 13 0.3 y t i c i r t n e c c e 0.2 0.1 -100 -80 -60 -40 -20 time [Myr] 5/1 9/2 4/1 7/2 3/1 0 Figure 5: Some examples of the recent evolution of the eccentricity of Mer- cury. During the last 50-60 Myr all the orbits present the same evolution, before which the chaos effect takes place. Horizontal lines correspond to the critical eccentricities that destabilize the equilibrium in a given spin-orbit resonance (Tab.2). During the last 50 Myr the eccentricity was certainly below 0.13, thus the 4/1 resonance is not a possible final outcome for Mercury. For some orbits the eccentricity is below 0.09 in the last 100 Myr and the 7/2 resonance was also destabilized. However, since this scenario is not true for all orbital solutions, we may expect a few final evolutions captured in this last configuration (Tab.4). e 0 1 g o l -0.5 -1.0 -1.5 -2.0 -2.5 -3.0 -3.5 -4.0 -4.5 -5.0 -5.5 4/1 7/2 3/1 5/2 2/1 3/2 0 20 40 60 80 100 number of orbital solutions (%) Figure 6: Cumulative distribution of the minimal eccentricities attained for the 1000 orbital solutions that we used. Straight lines represent the critical values of the eccentricity for each spin-orbit resonance (Tab.2), while dots represent the amount of captures obtained numerically for spin-orbit resonances that are still stable below each critical value of the eccentricity (Tab.4). 14
1006.5492
2
1006
2011-04-29T06:29:19
A comparison of spectroscopic methods for detecting starlight scattered by transiting hot Jupiters, with application to Subaru data for HD 209458b and HD 189733b
[ "astro-ph.EP" ]
The measurement of the light scattered from extrasolar planets informs atmospheric and formation models. With the discovery of many hot Jupiter planets orbiting nearby stars, this motivates the development of robust methods of characterisation from follow up observations. In this paper we discuss two methods for determining the planetary albedo in transiting systems. First, the most widely used method for measuring the light scattered by hot Jupiters (Collier Cameron et al.) is investigated for application for typical echelle spectra of a transiting planet system, showing that detection requires high signal-to-noise ratio data of bright planets. Secondly a new Fourier analysis method is also presented, which is model-independent and utilises the benefits of the reduced number of unknown parameters in transiting systems. This approach involves solving for the planet and stellar spectra in Fourier space by least-squares. The sensitivities of the methods are determined via Monte Carlo simulations for a range of planet-to-star fluxes. We find the Fourier analysis method to be better suited to the ideal case of typical observations of a well constrained transiting system than the Collier Cameron et al. method. We apply the Fourier analysis method for extracting the light scattered by transiting hot Jupiters from high resolution spectra to echelle spectra of HD 209458 and HD 189733. Unfortunately we are unable to improve on the previous upper limit of the planet-to-star flux for HD 209458b set by space-based observations. A 1{\sigma}upper limit on the planet-to-star flux of HD 189733b is measured in the wavelength range of 558.83-599.56 nm yielding {\epsilon} < 4.5 \times 10-4. Improvement in the measurement of the upper limit of the planet-to-star flux of this system, with ground-based capabilities, requires data with a higher signal-to-noise ratio, and increased stability of the telescope.
astro-ph.EP
astro-ph
Mon. Not. R. Astron. Soc. 000, 000 -- 000 (0000) Printed 2 October 2018 (MN LATEX style file v2.2) A comparison of spectroscopic methods for detecting starlight scattered by transiting hot Jupiters, with application to Subaru data for HD 209458b and HD 189733b Sally V. Langford,1,2 J. Stuart B. Wyithe,1 Edwin L. Turner,2,3 Edward B. Jenkins,2 Norio Narita,4 Xin Liu,2 Yasushi Suto2,5,6 and Toru Yamada7 1School of Physics, University of Melbourne, Parkville, Victoria 3010, Australia 2Princeton University Observatory, Princeton NJ 08540, USA 3Institute for the Physics and Mathematics of the Universe, University of Tokyo, Kashiwa 277-8568, Japan 4National Astronomical Observatory of Japan, 2-21-1 Osawa, Mitaka, Tokyo, 181-8588, Japan 5Department of Physics, The University of Tokyo, Tokyo 113-0033, Japan 6Research Center for the Early Universe, School of Science, University of Tokyo, Tokyo 113-0033, Japan 7Astronomical Institute, Tohoku University, Sendai 980-8578, Japan 2 October 2018 ABSTRACT The measurement of the light scattered from extrasolar planets informs atmospheric and formation models. With the discovery of many hot Jupiter planets orbiting nearby stars, this motivates the development of robust methods of characterisation from follow up observations. In this paper we discuss two methods for determining the planetary albedo in transiting sys- tems. First, the most widely used method for measuring the light scattered by hot Jupiters (Collier Cameron et al. 1999) is investigated for application for typical ´echelle spectra of a transiting planet system, showing that a detection requires high signal-to-noise ratio data of bright planets. Secondly a new Fourier analysis method is also presented, which is model- independent and utilises the benefits of the reduced number of unknown parameters in tran- siting systems. This approach involves solving for the planet and stellar spectra in Fourier space by least-squares. The sensitivities of the methods are determined via Monte Carlo sim- ulations for a range of planet-to-star fluxes. We find the Fourier analysis method to be better suited to the ideal case of typical observations of a well constrained transiting system than the Collier Cameron et al. (1999) method. To guide future observations of transiting planets with ground-based capabilities, the expected sensitivity to the planet-to-star flux fraction is quantified as a function of signal-to-noise ratio and wavelength range. We apply the Fourier analysis method for extracting the light scattered by transiting hot Jupiters from high resolu- tion spectra to ´echelle spectra of HD 209458 and HD 189733. Unfortunately we are unable to improve on the previous upper limit of the planet-to-star flux for HD 209458b set by space- based observations. A 1σ upper limit on the planet-to-star flux of HD 189733b is measured in the wavelength range of 558.83 -- 599.56 nm yielding ǫ < 4.5 × 10−4. This limit is not sufficiently strong to constrain models. Improvement in the measurement of the upper limit of the planet-to-star flux of this system, with ground-based capabilities, requires data with a higher signal-to-noise ratio, and increased stability of the telescope. 1 INTRODUCTION More than 500 planets have been discovered outside the Solar Sys- tem in the past two decades. Among these planets is a class of ob- jects known as hot Jupiters. These are roughly Jupiter-mass ob- jects that orbit their host stars within 0.05 AU (Seager et al. 2000), but not close enough to be considered very hot Jupiters, where thermal emission dominates the visible scattered light. It has been predicted that hot Jupiter planets will allow the starlight scattered by their upper atmospheres to be directly detected in the optical, with an expected flux of less than 10−4 times the direct stellar flux (Collier Cameron et al. 1999; Charbonneau et al. 1999). With increasing numbers of hot Jupiter planets currently being discovered, follow up observations become very important for char- acterisation that requires longer observation times, and for measur- ing the accuracy of the orbital parameters. The space-based CoRoT and Kepler missions, which are currently monitoring many thou- sands of nearby stars, will yield a wealth of new transiting planets. These planets will also be followed up with radial velocity mea- surements to constrain the orbital parameters. In order to fully re- alise the scientific potential of these systems, robust methods of 2 Langford et al. analysis of the starlight scattered by the transiting hot Jupiter plan- ets are required, as well as an understanding of their application to various systems. There has been no definite detection of the broadband un- polarised visible light scattered by a hot Jupiter. Extrasolar plan- ets have only recently been directly imaged in the optical. These planets orbit their parent stars at ∼100 AU (Kalas et al. 2008). In addition, visible polarised scattered light has been detected from the extended atmosphere of the transiting planet HD 189733b (Berdyugina et al. 2008, 2011). However, Wiktorowicz (2009) was unable to reproduce this result. Measuring the light scattered by a hot Jupiter planet is challenging due to the small angular sepa- ration, and the small planet-to-star flux ratio, which can be of the order of the systematic noise of ground-based observations. Ex- tracting the planet signal from the combined noisy stellar and scat- tered flux therefore generally requires making assumptions about the system. Upper limits on the visible light scattered by a hot Jupiter planet have been measured with ground-based and space- based capabilities (Charbonneau et al. 1999; Collier Cameron et al. 2002; Leigh et al. 2003a; Rowe et al. 2008; Kipping & Bakos 2011; Alonso et al. 2009a,b; Rogers et al. 2009; Christiansen et al. 2010; Sing & L´opez-Morales 2009). The light scattered by an extrasolar planet will depend on the composition of the atmosphere as well as the orbital parameters, allowing the testing of atmospheric models (Marley et al. 1999; Green et al. 2003). Models predict a range of planet-to-star fluxes for different classes of planetary atmospheres, which can guide the search for thermal emission and reflected light from hot Jupiter planets (Sudarsky et al. 2000, 2003). The non-detections of scat- tered light and resulting upper limits on the albedos of hot Jupiter planets aids the development of theories of extrasolar planet at- mospheres. The lack of a detection of starlight scattered by the planet HD 209458b suggests that the atmosphere is much less reflective than previously expected, and lacks highly scattering clouds (Seager et al. 2000; Rowe et al. 2006; Burrows et al. 2008). Increasingly sophisticated models are attempting to explain the in- flated radius and absence of a verifiable detection of light scattered by this hot Jupiter planet (Hood et al. 2008; Burrows et al. 2008). Further investigation and observation of known hot Jupiter planets will continue to guide the models, and potentially yield a definite detection of scattered broadband unpolarised visible light. 1.1 Current Methods Prior to the discovery of transiting extrasolar planets, two meth- ods of measuring the planet-to-star flux of hot Jupiter planets were concurrently developed by Charbonneau et al. (1999) and Collier Cameron et al. (1999). Both methods require modelling the direct stellar contribution, which is then removed, leaving the planet signal buried in noise. They model the system over a range of orbital phases, inclinations and velocities in order to find the signa- ture of the planet's scattered light, and the likelihood of the best fit orbital parameters. Both methods been important in beginning to constrain models of planetary atmospheres (Sudarsky et al. 2000, 2003). An upper limit on the planet-to-star flux of the extrasolar planet τ Bootis b was measured to be ≤ 5 × 10−5 times the direct stellar flux at the 99 per cent confidence level (Charbonneau et al. 1999). This measurement corresponds to a wavelength range of 465.8 -- 498.7 nm, and assumes an inclination of i ≥ 70o. The result was limited by the photon noise of the data, not systematic effects. In disagreement with this, Collier Cameron et al. (1999) mea- sured a best fit to the planet-to-star flux of the same system to be (7.5 ± 3) × 10−5 at the 97.8 per cent confidence level. This de- tection contains systematic uncertainty due to some dependence on wavelength of the albedo of τ Bootis b. For the wavelength re- gion of 456 -- 524 nm the planet-to-star flux ratio was found to be (1.9 ± 0.4) × 10−4, which is brighter than the upper limit set by Charbonneau et al. (1999). Outside this wavelength region there was no signal detected. The wavelength dependence of detected scattered light is useful in investigating the classes of hot Jupiter planet atmospheres. Disagreement in planet-to-star flux measure- ments may be due to the different phase functions adopted for the planets, and the methods used for subtracting the stellar contribu- tion. For example, Collier Cameron et al. (1999) adopt a Venus-like phase function of the planet to account for higher backscattering from clouds, as compared to a simple Lambert-law phase func- tion. Leigh et al. (2003a) failed to confirm the tentative detection of light scattered by τ Bootis b with subsequent reanalysis and addi- tional data. Collier Cameron et al. (2000) later revised their result, setting an upper limit on the planet-to-star flux of ≤ 3.5 × 10−5. Rodler et al. (2010) recently measured an upper limit on the light reflected by τ Bootis b to be ≤ 5.7 × 10−5 times the direct stellar flux, at the 99.9 per cent significance level. This upper limit was based on the method devised by Charbonneau et al. (1999). Collier Cameron et al. (2002) set an upper limit on the υ An- dromeda system with a 0.1 per cent false alarm probability of ≤ 5.84 × 10−5. With the same method, Leigh et al. (2003b) placed an upper limit on the planet-to-star flux of HD 75289b of ≤ 4.18 × 10−5 at the 99.9 per cent level. However, Rodler et al. (2008) disputed this result, stating that it was based on incorrect orbital phases for the planet. They rederived the upper limit on the planet-to-star flux of HD 75289b to be ∼ 60 per cent higher, via the approach taken by Charbonneau et al. (1999). These methods yield varying results for the same systems, most likely due to the difficulty in measuring the dim planet-to- star flux, their model dependence and the unconstrained orbital pa- rameters. However, the upper limits set are lower than predicted by theoretical considerations (Collier Cameron et al. 1999). This sug- gests that hot Jupiter planet atmospheres are less reflective than ex- pected based on the solar system gaseous planets. Constraining at- mospheric models further requires deeper upper limits on the scat- tered light, and minimising the model-dependence of measuring the albedos of hot Jupiter planets. In order to reduce the parameter space and number of assump- tions, Liu et al. (2007) developed a method specific to detecting the starlight scattered by transiting hot Jupiter planets. Knowing the planet's velocity from orbital constraints allows a shift-and-add method to be adopted, to verify the presence of the scattered light. Using the equivalent width ratios of the scattered and direct com- ponents of the spectra, Liu et al. (2007) found the planet to star ratio of HD 209458b to be (1.4 ± 2.9) × 10−4 in the wavelength range 544 -- 681 nm. This result is expected to be updated, following a correction for the non-linearity of the HDS CCDs (Tajitsu et al. 2010). Space-based satellites can use transit photometry to mea- sure the starlight scattered by hot Jupiters at a higher signal-to- noise ratio than possible with ground-based capabilities. To de- termine the amount of starlight scattered by HD 209458b from the depth of the secondary eclipse, Rowe et al. (2008) measured light curves of HD 209458 for the duration of the planetary orbit with the Microvariability and Oscillations of Stars (MOST) satel- lite. Their observations yielded a 1σ upper limit on the planet- to-star flux ratio of ≤ 1.6 × 10−5. In the wavelength region of Detecting starlight scattered by transiting hot Jupiters 3 400 -- 700 nm this corresponds to a geometric albedo of 0.038 ± 0.045 which is much less reflective than the solar system's gi- ant planets. This rules out bright clouds at a high altitude, and is consistent with the low albedo upper limits of υ Andromeda and HD 75289 (Collier Cameron et al. 2002; Leigh et al. 2003b). While searching for transiting planets, the CoRoT and Kepler missions have also yielded a range of planetary upper limits on the albedos of the hot Jupiter planets such as Kepler-7b, Kepler- 5b, CoRoT-1b, CoRoT-2b and OgleTr56(Kipping & Bakos 2011; Alonso et al. 2009a,b; Rogers et al. 2009; Christiansen et al. 2010; Sing & L´opez-Morales 2009). Further studies and repeated obser- vations of these bright, transiting systems will continue to constrain atmospheric models and theories of the formation of solar systems. However, space based missions are costly and it is not feasible to consistently follow up on the expected yield of hot Jupiter plan- ets while the search continues for Earth-like extrasolar planets. It is therefore beneficial, and timely to develop a suitable model- independent method of extracting the planet-to-star flux from tran- siting hot Jupiters in ground-based observations. It is also advan- tageous to measure phase independent planet-to-star fluxes with- out assuming a grey albedo. This motivates a model-independent Fourier based method for measuring the light scattered by hot Jupiter planets, which we introduce in this paper. This method re- quires minimal assumptions about the system and is suited to tran- siting planets where parameters are constrained by survey observa- tions. In this paper we quantify the precision of the existing method of Collier Cameron et al. (1999) in measuring the planet-to-star flux of transiting hot Jupiter planets via Monte Carlo simulations. We also introduce and quantify a new method specifically tailored for application to transiting planets in order to constrain the orbit and detect smaller signals with ground-based capabilities. We begin with a review of the method of measuring the light scattered by ex- trasolar planets developed by Collier Cameron et al. (1999). This method adopts a detailed approach to extracting the signal of the planet from the residual noise of the spectrum by least-squares de- convolution, following the stellar spectrum removal. Monte Carlo simulations are adopted to determine the sensitivity of this method for application to mock high resolution ´echelle spectra of a transit- ing hot Jupiter planet system, observed with ground-based capabil- ities. A new method is then developed which is model-independent and utilises the benefits of reduced parameters for a transiting sys- tem. This Fourier analysis method involves solving for the planet and stellar spectra in Fourier space by least-squares, and is applica- ble to transiting hot Jupiter planets. To guide future observations, the expected sensitivity to the planet-to-star flux fraction is quan- tified as a function of signal-to-noise ratio and wavelength range. This method is then tested on Subaru HDS spectra of HD 209458 and HD 189733, which host a transiting hot Jupiter planet. 2 SYSTEM PARAMETERS A planet will reflect an amount fp of its host star's flux f⋆; fp(α, λ) f⋆(λ) a (cid:19)2 = p(λ)g(α, λ)(cid:18) Rp(λ) ǫ(α, λ) ≡ where a is the semi-major axis, Rp(λ) is the radius of the planet, p(λ) is the geometric albedo, and g(α, λ) is the phase func- tion at planetary phase angle α and wavelength λ (Leigh et al. 2003c). From theoretical considerations a planet-to-star flux of , (1) ǫ ∼ 10−4 is expected for a hot Jupiter planet at full phase (Collier Cameron et al. 1999). 2.1 Phase Function The flux scattered by a hot Jupiter planet is a phase dependent frac- tion of the maximum reflectance at full phase. Unfortunately, due to a lack of signal and phase coverage it is not yet possible to de- termine the exact phase function g(α) for a hot Jupiter planet being observed, and if required, it must therefore be modelled as a func- tion of the phase angle α. It is simplest to adopt a Lambert-law sphere which assumes isotropic scattering over 2π steradians. This typically has the form; g(α) = [sin α + (π − α) cos α]/π, where the phase angle α is determined from the inclination i and the orbital phase φ; (2) cos α = - sin i cos 2πφ. (3) For an orbital phase of φ = 0 the planet is closest to the observer, and has the smallest planetary phase. For an orbital phase of φ = 0.5, the planet is at the furthest point in its orbit from the observer, with maximum reflectance. To account for stronger back-scattering than the Lambert-law phase function, it may be more realistic to adopt the phase function of a similar planet in the Solar System, that has a cloud-covered surface. For example; g(α) = 10−0.4∆m(α), (4) where; ∆m(α) = 0.09(cid:16) α 100o(cid:17) + 2.39(cid:16) α 100o(cid:17)2 − 0.65(cid:16) α 100o(cid:17)3 , (5) which is a polynomial approximation to the empirically determined Venus-like phase function (Collier Cameron et al. 2002). For tran- siting planets, which have orbital inclinations close to 90o, these two functions have a similar variation in flux of less than 10 per cent over the orbital phases adopted for the mock spectra presented here, and the Subaru HDS spectra of HD 189733b and HD 209458b. 2.2 Velocity A planet's velocity relative to its host star Vp(φ) at an orbital phase φ is given by; Vp(φ) = Kp sin(2πφ), (6) where φ = (t − T0)/Porb at the time t, where T0 is the transit epoch and Porb is the orbital period, for a circular orbit (ellipticity e=0). The apparent radial-velocity amplitude Kp about the centre of mass of the system is; Kp = 2πa Porb sin i 1 + q , (7) where a is the orbital distance, i is the inclination and q = Mp/M⋆ is the mass ratio of the planet to the star. For a system under con- sideration, a, i and Rp can be determined from radial-velocity mea- surements and transit photometry. For non -- transiting planets these parameters have greater uncertainty, and Kp is determined from fit- ting observational light curves. 4 Langford et al. A B Figure 1. Mock spectra constructed from high signal-to-noise ratio solar spectrum. Panel A - Example mock spectra for red wavelength range of 560 -- 608 nm. Panel B - Example mock spectra for blue wavelength range of 440 -- 515 nm. Statistical Gaussian noise added to the spectrum is distributed with standard deviation √N for N counts per pixel. The wavelength is logarithmically binned such that the Doppler shifted scattered light can be added as an integer pixel shifted copy of the star, with amplitude scaled by ǫg(α). The orders of the spectra were combined and normalised with an 11-knot cubic-spline. 3 CONSTRUCTION OF MOCK SPECTRA Mock spectra were constructed in order to quantify the precision of methods for measuring the light scattered by extrasolar planets, including that by Collier Cameron et al. (1999), as well as the new Fourier analysis method, introduced in Section 5. Spectra were con- structed from high resolution solar spectra (λ/∆λ ∼ 300, 000), obtained with the Fourier Transform Spectrometer at the Mc- Math/Pierce Solar Telescope situated on Kitt Peak, Arizona1. The mock spectra were constructed to be analogous to observing a bright hot Jupiter transiting system, such as HD 209458, with typ- ical ground-based 8 m telescope capabilities, such as Subaru HDS. An example mock spectrum is shown in Fig. 1. The high resolution solar spectra were scaled to the counts expected from HD 209458 with Subaru HDS for a typical exposure time of 500 seconds. A spectral resolution of R = λ/∆λ = 45, 000 was mimicked by Gaussian smoothing, consistent with twice the critical oversam- pling. Noise due to statistical fluctuation in photons from the hy- pothetical measurement with Subaru HDS was added after includ- ing the planet contribution and smoothing. The noise is Gaussian distributed with a standard deviation of √N for N counts per pixel. The planet velocity dictates the wavelength shift of the scattered spectrum, taken as a scaled copy of the stationary stellar spectrum. During the length of an exposure the planet velocity varies by less than the width of the absorption features. Planet velocities were randomly selected within the range 30 -- 80 km s−1 towards the observer. This is consistent with the range of velocities of HD 209458b when the planet is just out of sec- ondary eclipse and the orbital phase is in the range 0.55 -- 0.60. There is less than 10 per cent variation in the flux scattered from the planet within this phase range based on either a Lambert-sphere or a Venus-like phase function. This is a benefit of using transiting systems to measure planet-to-star fluxes. As the planet velocity am- 1 Operated by the National Solar Observatory, a Division of the National Optical Astronomy Observatories. NOAO is administered by the Associ- ation of Universities for Research in Astronomy, Inc., under cooperative agreement with the National Science Foundation. plitude can be measured from the constrained parameters, a smaller range of the orbit can be observed and selected to be where the planet is close to full phase and maximum reflectance. The veloc- ities must be large enough to shift the planet signal clear of the stellar absorption lines. Instrumental variations, such as flexure of the spectrograph and guiding errors, can cause wavelength variations between the spectra during a night of observations (Winn et al. 2004). This can be accounted for in data reduction by adjusting the wavelength cal- ibration using the star's spectral absorption lines. Variation in the flux calibration between exposures is removed by continuum fitting and normalisation. The mock spectra are taken to be ideal, with no variation in the shape of the response of the CCD with time and the host star spectra in all exposures are assumed to be perfectly aligned. 4 COLLIER CAMERON METHOD The most widely used current method (henceforth the Collier Cameron method) was first presented by Collier Cameron et al. (1999), and further developed by Collier Cameron et al. (2002) and Leigh et al. (2003b). The Collier Cameron method has previously been adopted to attempt to measure the light scattered by three different, non-transiting hot Jupiter planets (Collier Cameron et al. 1999, 2002; Leigh et al. 2003b). It involves detailed modelling of the stellar contribution and matched-filter analysis to extract the planet signal from the noise. It is a useful method for the case where the hot Jupiter planet's orbit is unknown, and has allowed constraints to be placed on highly reflective planetary atmospheres (Rodler et al. 2010). In this Section, the Collier Cameron method is reviewed for application to a transiting planet system, using the mock spectra described in Section 3, within Monte Carlo simula- tions. For the analysis of non-transiting planets, the Collier Cameron method requires observations of as much of a full orbit as possible, such that the most likely orbital phase function and planet velocity can be measured. For transiting systems the velocity of the planet Detecting starlight scattered by transiting hot Jupiters 5 There are typically ripples in the deconvolved stellar profile at low velocities that may cause spurious effects on measuring the planet contribution (Collier Cameron et al. 2002). For the Monte Carlo simulations the planet signal was extracted from the range of velocities with magnitudes greater than 30 km s−1. To extract the planet signal, a matched-filter is constructed, and fit to the data to determine the planet-to-star flux. The velocity profile of the planet signal is modelled as a moving Gaussian, based on the form used in Collier Cameron et al. (2002). The planet-to-star flux is de- termined by scaling the matched filter to match the deconvolved planet line profile, and finding the best fit with a χ2 minimisation (Collier Cameron et al. 1999). 4.1 Monte Carlo Analysis The sensitivity of the Collier Cameron method was determined for a range of signal-to-noise ratio strengths and planet-to-star fluxes via Monte Carlo simulations. The typical signal-to-noise ratio per pixel of the mock spectra was varied by increasing the counts per pixel, corresponding to a longer exposure time. A larger telescope collecting area or observing a brighter star would also increase the counts per pixel. The number of standard deviations separating a detection from zero was determined by the mean of the distribution of all Monte Carlo simulation solutions, divided by the standard deviation. This defines the distance of the mean value from zero in units of the spread of the distribution. For example, a result quoted at 1σ has a distribution with a mean value that is one standard de- viation away from zero, as shown in Fig. 2. In the red wavelength range of the mock spectra of 560 -- 680 nm, with 1465 absorption lines, a signal-to-noise ratio of ∼ 800 is required for a 1σ detection of a planet-to-star flux of 1.6 × 10−4. The blue wavelength range of 440 -- 515 nm has many more absorption lines (2570), such that the method is more sensi- tive to smaller planet-to-star fluxes for data with the same signal- to-noise ratio, assuming the stellar spectra is perfectly removed. Planet-to-star fluxes greater than 10−4 are detectable at the 1σ level. A planet-to-star flux of 1.6 × 10−5 requires a signal-to-noise of around ∼ 800 to be detected at the 3σ level, for the ideal case. The results of the Monte Carlo simulations with varying signal-to- noise ratio for the red wavelength range of 560 -- 680 nm and blue wavelength range of 440 -- 515 nm are presented in Fig. 3. There is a limit on the possible signal-to-noise ratio of the spectra due to the capability of the CCD and variations in the planet velocity during longer exposure times smearing the scattered signal. The results of the Monte Carlo simulations for ideal spec- tra are consistent with detections via this method for real spec- tra. Probable detections have been made for planet-to-star fluxes brighter than ǫ = 10−4, and upper limits are set with low false- alarm probabilities. For example, Collier Cameron et al. (1999) de- tected a best-fit to the signal of the light scattered by τ Bootis b of ǫ = 1.9 × 10−4 for data with a typical signal-to-noise ratio of ∼1000, in a central wavelength range. As the Collier Cameron method was developed prior to the discovery of transiting hot Jupiter planets, it did not take advantage of the minimisation of unknown parameters in these systems. This method can benefit from the known planet velocity in extracting the scattered signal. However in constructing the template spectrum, the spectra are summed to smear out the planet signal. This requires a large range of velocities, or isolating spectra with the planet at minimal illumination. Observing a small section of the transiting planet orbit may therefore affect the accuracy of the stellar model, and reduce the planet signal in the residual spectrum. Figure 2. Example probability distribution. An example probability dis- tribution with a mean planet-to-star flux of ǫ = 1.2 × 10−4 marked by the vertical solid line. The distance of one, two and three standard deviations from the mean are marked by the vertical dashed lines. This distribution has a statistical significance of > 1σ as the mean is greater than one standard deviation, but less than two standard deviations from zero. is known, and a small section of the orbit can be observed. As we are evaluating the Collier Cameron method for the case where the planet is transiting, the thirty mock spectra that were used to mea- sure a planet-to-star flux are in a small orbital phase range, either side of the secondary eclipse, when the majority of the dayside of the planet is facing the observer. This corresponds to velocities with magnitudes of 30 -- 80 km s−1. The planet's spectral features are therefore shifted clear of the corresponding stellar features, and the planet is close to maximum brightness. The planet is assumed to follow a Venus-like phase function as defined in Section 2.1. The initial input spectra do not need to be merged for this method, but can remain as separate orders, without continuum normalisa- tion. The spectra are in the red wavelength range of 560 -- 680 nm, and the blue wavelength of range of 440 -- 515 nm. We note that the mock spectra are ideal, in that they are all aligned, without cosmic ray hits or vignetting, and the stellar contribution is constant. The Collier Cameron method requires modelling and remov- ing the direct starlight from the total spectrum, leaving the planet signal buried in noise. It is very difficult to accurately remove the stellar contribution, as small distortions in the strong stellar lines can produce changes larger than the planet signature. Problems can arise due to shifting of the spectrum on the detector, changes in the telescope focus and atmospheric seeing. In the blue wavelength range, the density of strong lines makes removing the stellar sig- nal particularly difficult, and observations for previous upper limit measurements have generally been in a central wavelength range. In this paper, results are presented for the ideal mock spectra, and the stellar spectrum does not vary with exposure. This aids the removal of the modelled stellar spectrum, since we do not have to account for sources of distortion. This simplification of the removal of the direct light is equivalent to perfectly modelling the stellar spectrum, and is the best case scenario for the Collier Cameron method described in Collier Cameron et al. (2002). The sensitivity of the method to the planet-to-star flux will be impeded by instru- mental variations, and the difficulty in modelling the stellar spec- trum, particularly in the blue wavelength. The planet signal strength is extracted from the residual after the direct flux subtraction via least-squares deconvolution with a set of known weights of stellar absorption lines. The method is de- scribed in detail in Collier Cameron et al. (2002). This results in a deconvolved velocity profile which represents the lines present in the planet spectrum. It is similar to a cross-correlation, but neigh- bouring blended lines are dealt with more accurately. For the mock spectra, a library of spectral lines in the required wavelength range was compiled from data for the Sun from the Vienna Atomic Line Data Base (VALD) (Kupka et al. 1999). 6 Langford et al. A B Figure 3. Signal-to-noise ratio for mock spectra with Collier Cameron method. Contour plot of 1σ, 2σ and 3σ levels of detection (solid) for signal-to-noise ratio values in the range 102 -- 103, and ǫ values in the range 10−5 -- 2 × 10−4. Panel A - Results for the red wavelength range 560 -- 680 nm. Panel B - Results for the blue wavelength range 440 -- 515 nm. The dots represent the sampling of the parameter space. While the analysis methods developed by Charbonneau et al. (1999) and Collier Cameron et al. (1999) are required for a non- transiting system, they have not yielded a detection of the light scattered by a hot Jupiter planet. With the focus of many cur- rent extrasolar planet surveys on transiting objects, it is therefore timely to develop a method of extracting the planet-to-star flux spe- cific to transiting hot Jupiter planets, that does not require mod- elling the stellar contribution. In the next Section we describe a model-independent Fourier analysis method that requires minimal assumptions about the system and is suited to transiting planets where the parameters are constrained by other space-based, survey observations. 5 FOURIER ANALYSIS METHOD The Fourier analysis method for extracting the scattered signal from the combined planet and star flux described in this Section is based on a method developed to deal with the effect of the fixed-pattern response function of a detector on a moving spec- trum (Jenkins 2002). In the case of a transiting planet, the stellar spectrum is considered to be stationary (analogous to the response function), while the planet's scattered spectrum moves with each exposure. Separating out a moving spectrum via a Fourier trans- form is a useful technique for binary systems (see e.g. Hadrava (1995)). The observed spectrum recorded by the ´echelle spectrom- eter T(λ), consists of the direct starlight contribution S(λ), plus the starlight scattered by the orbiting planet P(λ). The planet spectrum is shifted with respect to the stationary stellar spectrum with a pre- dictable Doppler shift, dependent on the velocity of the planet at time of the exposure. There is also a contribution from statistical noise that is not frequency dependent. For the total spectrum T(λ) we define: tx(ω) = Re{F [T(λ)]}, ty(ω) = Im{F [T(λ)]}, where F is the Fourier transform operator. Similarly px(ω), py(ω), sx(ω) and sy(ω) are defined for the planet and stellar spectra. As each spectrum has a different Doppler velocity displacement for P(λ), depending on the observed velocity of the planet, the zero shift complex function px(ω) + ipy(ω) is multiplied by (8) (9) exp(−2πiωδk), where δk is the magnitude of the shift of P(λ) in the kth exposure compared to the zero velocity case. The minimum of the sum of the squared real and imaginary residuals (the minimum of Q2) is solved to determine the unknown planet and stellar spectrum coefficients px(ω) and py(ω), sx(ω) and sy(ω). This is done by setting the partial differentials of Q2 defined below with respect to the four unknowns px, py, sx and sy at each ω, to be zero; n Q2 = (px cos ∆k + py sin ∆k + sx − tx,k)2 Xk=1 Xk=1 (−px sin ∆k + py cos ∆k + sy − ty,k)2, n + (10) (11) where ∆k is an abbreviation for 2πωδk. Therefore, the best solutions for px(ω), py(ω), sx(ω) and sy(ω) are found by solving a system of four linear equations at each ω. The linear equations can be summarised as; C(ω) · u(ω) = b(ω), where u(ω) is a vector of px, py, sx and sy at each ω and the coef- ficients Ci,j and bi are listed in Table 1. (12) Singlular value decomposition is used to solve the system of linear equations. It can be verified that the solution is a minimum in Q2 by plotting nearby solutions. After solving for each vector u(ω) over all ω, the inverse Fourier transforms of the two pairs of terms u1(ω) + iu2(ω) and u3(ω) + iu4(ω) recover the best representations of P(λ) and S(λ). The planet spectrum, P(λ), is divided by the stellar spectrum, S(λ), and the mean found over all wavelengths, yielding the corresponding planet-to-star flux. The solution can be verified to be the planet contribution by repeating the measurement while shuffling the order of the Doppler shifts with respect to their exposures, which will yield a null result. If the velocity displacements of the observations in a set are equally spaced, the the matrices become singular and the solu- tions can blow up. Therefore, the timing of the observations should be such that the velocity shifts are noncommensurate. This is not unique to the Fourier analysis method, and should be considered when planning an observing program. Solving for the planet and stellar spectra in this way has the following benefits; (i) it does not require a model of the star, (ii) it is still effectively using the signal of all of the absorption lines, Detecting starlight scattered by transiting hot Jupiters 7 j=1 j=2 j=3 j=4 bi Ci,j i=1 i=2 n 0 0 n i=3 P cos ∆k P sin ∆k i=4 −P sin ∆k P cos ∆k P cos ∆k −P sin ∆k P tx,k cos ∆k −P ty,k sin ∆k P sin ∆k P cos ∆k P tx,k sin ∆k +P ty,k cos ∆k n 0 0 n P tx,k P ty,k Table 1. Coefficients for the matrix C(ω) and vector b(ω). All summations are from k = 1 to n, where n is the number of spectra. although they are not summed over to increase the signal-to-noise ratio, meaning that the planet-to-star flux solution can be wave- length dependent, and (iii) there is no assumption about the form of the planet spectra, only that it is moving compared to the stationary stellar spectrum with specific Doppler shifts. For example, forcing the planet spectra, px(ω) and py(ω) to be a scaled copy of the stel- lar spectra, ǫ(α)sx(ω) and ǫ(α)sy(ω) respectively, would require solving three highly non-linear coupled equations, and remove the freedom of solving for a wavelength dependent albedo. 5.1 Fourier Transform The Fourier transform is computed using FFTW, a C subroutine library (Frigo & Johnson 2005)2. Before transforming the input spectra into Fourier space, the spectra are extended above and be- low the wavelength range by adding the value of the continuum, and then damped to zero at the ends gradually, using a Hanning window which has the form; A B C (13) f (λ) = 1 2 (cid:16)1 − cos(cid:16) 2πλ N (cid:17)(cid:17) , for N pixels. FFTW assumes that the spectrum is infinitely periodic, and the Hanning window suppresses discontinuities as it loops over the wavelength range when transforming into Fourier space. The low frequency results of px(ω) and py(ω), sx(ω) and sy(ω) need to be dealt with carefully. At these frequencies the phase shifts induced by the changes in the planet's velocity be- come so small that they are difficult to detect. In essence, the planet and star signals are no longer easily separable. Mathematically, this problem manifests itself by making the matrix C singular or nearly so. As a consequence, the solutions start to be dominated by noise or low-frequency systematic errors. By contrast, narrow stellar fea- tures that have a presence in both S(λ) and P(λ) can create measur- able amplitudes and phase shifts at high frequencies, and thus they contain most of the planet's detectable signal. In order to extract this important information, the spectra are effectively put through a high-pass filter, and the low frequencies are not solved for. So that the result can be inverse Fourier transformed, the low frequency so- lutions in Fourier space are replaced with a predicted solution. This requires the simple assumption that the planet spectrum is a scaled copy of the stellar spectrum. The average of the Fourier amplitudes of the input spectra is taken to be an estimate of the Fourier ampli- tudes of the stellar spectra. The planet signal will be averaged out, due to different Doppler shifts in each exposure. This is multiplied 2 See htp://www.fftw.org/ Figure 4. Damping spurious signals at low frequencies. Panel A - Exam- ple of Fourier amplitudes of input blue mock spectra. Panel B - Example of Fourier amplitudes of the solution blue planet (upper subpanel) and star (lower subpanel) for mock spectra with ǫ = 0.1 and low frequencies replaced with the artificial signal to damp spurious small frequency signals. For the case where the planet-to-star flux is large, the planet and stellar signals ap- pear to be the same shape. Panel C - As for Panel B, with ǫ = 0.0001. For the smaller ǫ value the moving signal solution is of the order of the statistical noise, and hence the planet signal appears to be flat out to large frequencies. The central, small amplitude regions in Panels B and C, are due to the need for a smooth transition between the solution signal and the artificial signal. In this case a Hanning window was used to damp the solution and artificial signals to zero where they meet. The frequency range lost in doing this is the same for the planet and the star, therefore the resulting planet-to-star flux will not be affected. This can be verified with the mock spectra gener- ated with particular ǫ values. The Fourier amplitude of the zero frequency extends above the scale plotted, but has a finite value. Note that only every 5th point is plotted here. 8 Langford et al. by the predicted amplitude of the planet-to-star flux, and this test amplitude is iterated until it equals the solution from the Fourier analysis method - the real planet-to-star flux. A null result occurs when this iteration does not converge. The frequency region of the solution which is artificially re- placed is determined for when the matrix C is singular (ω = 0) or nearly singular. The gradual transition at the edge of this region uses a Hanning window to damp the artificial and solution Fourier amplitudes to zero where they meet (see Fig. 4). This is not the only possible method to combine the artificial and real solutions, how- ever it must be gradual enough not to cause ringing in the planet and star spectra when they are transformed back to real space. Due to this transition, sections of frequency space are lost. However, as it occurs equally in the stellar and planet spectra (due to the matching transition range), the resulting planet-to-star flux value is conserved. If the Fourier components are used to compare the stel- lar and planet solutions, this artificial replacement is not required, however the low frequency result is lost. The central frequency cannot just be set to zero to remove the low frequency solutions if the inverse Fourier transform is required, as it fixes the overall continuum average to also be zero. 5.2 Measuring the Planet-to-Star Flux The planet and the stellar spectra are solved for at each wavelength, by taking the inverse Fourier transform of the resulting Fourier components px(ω) and py(ω), sx(ω) and sy(ω). The spectra are retrieved, including the Hanning-windowed continuum-added ranges either side of the pixel range corresponding to the solution planet and stellar spectra; P(λ) and S(λ). The planet spectrum so- lution is aligned with the stellar spectrum, as it is shifted to the zero velocity case. The zero frequency value of the spectra is the power in the longest wavelength Fourier component, F(0). This corresponds to the continuum level, however we cannot explicitly solve for the zero shift solution. Therefore, it is artificially constructed before the inverse Fourier transform, and is used to iterate to the correct solution. The planet spectrum is divided by the stellar spectrum, and the mean ratio value taken for the entire range to be the value of the planet-to-star flux. This is then tested against the test ampli- tude for the artificial low frequency solutions, until the two values converge on the solution. The planet-to-star flux of the observation is determined with a small range of orbital phases, so that the amount of light the planet reflects does not vary rapidly. In order to scale this to the maxi- mum reflectance for comparison with the Collier Cameron result, a planet phase model must be assumed. For the mock spectra, the orbital phase is in the range 0.55 -- 0.60. This corresponds to an aver- age planet phase of 85 per cent illumination, for a Venus-like phase function. The planet phase at each exposure cannot be accounted for in solving for the planet spectrum in Fourier space, requiring a small region of the orbit to be observed. The actual planet-to-star flux ǫ(α) is therefore the model-independent measurement, which will be less then the maximum reflectance upper limit. 5.3 Non-Grey Albedo So far in this paper, a grey albedo has been assumed in constructing mock spectra. This means that the planet reflects an equal fraction of the starlight across the entire wavelength range. In the Collier Cameron method, the line weights can be attenuated according to a particular atmosphere model to account for a non-grey albedo. For exanple, in order to test for different classes of planets, such as the Class V roaster or Isolated Class IV models, a wavelength dependent planet-to-star flux cold be used for the artifical solution. Any increase in the likelihood of the measured planet-to-star flux can be used as an indication of the accuracy of the model. The Fourier analysis method provides an opportunity to solve for the planet signal at each wavelength, and therefore determine any slope of the albedo function with wavelength. However, as the zero shift solution cannot be solved, the F(0) values are estimated based on the assumption that the stellar spectrum can be repre- sented by the average input spectrum, and the planet as a scaled copy. Therefore, varying the shape of the continuum of the result- ing planet spectrum requires modelling possible wavelength depen- dent albedos as an iterative parameter to find the best solution. It is also possible to divide the input spectrum into short wavelength re- gions over which a non-grey albedo would vary slowly. This was the process adopted by Collier Cameron et al. (1999) for τ Bootis b over ranges of 40 nm, and their results suggested a wavelength de- pendence of the albedo at around 500 nm. Alternatively, solving for the planet's scattered spectrum allows the chance for strong absorp- tion lines to be explicitly revealed in high signal-to-noise ratio data. Typically these lines are expected to be below the level required to detect single absorption lines, but an advantage of the Fourier method is that it preserves the details of the planet's spectrum. 5.4 Monte Carlo Analysis Monte Carlo simulations were run for the Fourier analysis method with the mock spectra in order to determine the observing strat- egy most sensitive to small planet-to-star fluxes, and improvements possible with minimal assumptions and typical transit data. 5.4.1 Signal-to-Noise Ratio We constructed spectra with a range of signal-to-noise ratio and planet-to-star flux ratios, as per the Monte-Carlo analysis discussed in Section 4.1. The results of the Monte Carlo simulations with varying signal-to-noise ratios are presented in Fig. 5, correspond- ing to the red wavelength region of 560 -- 608 nm and the blue wave- length region of 440 -- 515 nm respectively. The planet-to-star flux is theoretically predicted to be ǫ ≤ 10−4 (Collier Cameron et al. 1999); the range of ǫ values is centered on this value. Fig. 2 shows the probability distribution and sensitivity level determined from an example Monte Carlo simulation, centred on the expected planet- to-star flux at maximum reflectance for the system. The resulting planet-to-star fluxes are scaled to the maximum reflectance for the contour plots presented in Fig. 5. The method is more sensitive to smaller planet-to-star fluxes in the blue wavelength range than the red wavelength range. This is most likely due to the increased number of absorption lines that have a larger signal in the Fourier domain, and in the higher fre- quency ranges, not as affected by small shifts in the planet spectra. For a typical signal-to-noise ratio of around 350, a 1σ detection can be made of scattered light with ǫ ≥ 4 × 10−5 in the blue wavelength region, as compared with ǫ ≥ 1 × 10−4 for the red wavelength region of the spectrum. In Fig. 5 the contours of levels of detection have a very steep gradient as ǫ decreases to 10−5, sug- gesting that very dim planets will require very high signal-to-noise ratios to be detectable via this method and ground-based capabil- ities. The theoretically predicted planet-to-star flux of ǫ ≤ 10−4 Detecting starlight scattered by transiting hot Jupiters 9 (Collier Cameron et al. 1999) is detectable at the 3σ level with rea- sonable signal-to-noise values in the blue wavelength region, and at the 1σ level in the red wavelength region. Similar to the Col- lier Cameron result, matching the 1σ upper limit of HD 209458b measured with the space-based satellite MOST (Rowe et al. 2008) requires the blue wavelength range Subaru HDS spectra with a signal-to-noise ratio greater than ∼ 900. 5.4.2 Spectral Resolution The spectral resolution of the test input spectra was also varied for Monte Carlo simulations with the Fourier analysis method. The signal-to-noise ratio was kept constant per pixel, but the oversam- pling varied by degrading the spectrum with Gaussian smoothing over an increasing radius. The spectral resolution range of R = 32,000 -- 90,000 corresponds to slit widths of 1.2 -- 0.4" and an over- sampling of 4 -- 12 pixels per resolution element. The critical sam- pling is 2 pixels per resolution element for Subaru HDS. Within this range there is no improvement in the sensitivity of the Fourier analysis method. However, increased spectral resolution may be beneficial to the data reduction process and correcting for instrumental effects and wavelength shifts not incorporated into the ideal test spectra. On the other hand, the increase in signal-to-noise ratio in ´echelle data gained by a decrease in spectral resolution may yield a measurement of a smaller planet-to-star flux. 5.4.3 Wavelength Range and Spectrum Length The Fourier analysis method was applied to two separate wave- length ranges. The two regions of spectra, 560.0 -- 608.0 nm (red wavelength range) and 440.0 -- 515.0 nm (blue wavelength range), are similar in total wavelength span, however, due to the con- stant bin width in logarithmic space, the blue region has many more pixels in total. A large number of lines are required to in- crease the planet signal above the level of the noise (Liu et al. 2007; Collier Cameron et al. 1999). For the Fourier analysis method, the blue wavelength range is more sensitive to smaller planet signals (see Fig. 5) due to the increased density of sharp absorption lines that have a larger signal in the Fourier domain. For the applica- tion of the Collier Cameron method to the mock spectra, with ideal stellar spectra removal, this is also true. However, for real data, where the stellar spectrum is not exactly known, a high density of absorption lines in the stellar spectrum is expected to increase the difficulty in fitting the continuum and removing the stellar spec- trum without affecting the planet signal. For many methods, spec- tra with wavelengths towards the red are better suited to detecting the planet signal, as absorption lines are scattered into regions of the spectrum without strong stellar lines (Liu et al. 2007). The con- tinuum regions in the red wavelength range also aid the removal of the stellar spectrum in the Collier Cameron method, as the template spectrum can be more accurately scaled. Using a shorter wavelength range is beneficial for decreas- ing the effect of assuming a wavelength-independent albedo. The Fourier analysis method shows gradual improvement in sensitivity to smaller planet-to-star fluxes for spectra covering a larger wave- length span, due to the increased number of absorption lines con- straining the moving signal. In practice this is limited by the regions of bad pixels on a detector, as the Fourier analysis method requires continuous spectra. 5.4.4 Number of Exposures For alternative methods of measuring the scattered light contri- bution of a non-transiting planetary system such as the Collier Cameron method, large numbers of spectra are required to fully cover the orbital phase range of the planet and determine the planet's velocity (Collier Cameron et al. 2002). This is not required for transiting systems, where the velocity of the planet is known for the full orbit. As the scattered light signal from the planet is intrin- sically dim, taking spectra at maximum planetary phase increases the chances of measuring a planet flux. Thirty exposures were used to measure the planet-to-star flux of the ideal system in testing the Fourier analysis and the Col- lier Cameron method. This allows for a range of planet velocities, with differing Doppler shifts, but limits the processing time and the spread of planet phases. The velocity must be large enough to shift the planet signal clear of the corresponding stellar features. Using a shorter exposure would yield more images within the possible orbital phase range, but reduce the signal in each exposure. For up to 50 exposures in a phase range of 0.55 -- 0.60 there is an improvement in the sensitivity to smaller planet-to-star fluxes of the Fourier analysis method, due to the stronger constraints on the moving signal. This results in exposures where the planet velocity differs by a couple of m/s in each exposure. The number of exposures with sufficiently different veloci- ties is limited by the range within which the planet phase does not rapidly vary, and the length of the exposure required for an ade- quate signal-to-noise ratio. It is important that the separations in the planet's velocity from one observation to the next are not pre- cisely uniform, as the solutions can blow up when the wavelength of the Fourier component is some multiple of the frequency spac- ing. The number of spectra used for the Monte Carlo simulations is consistent with the number of spectra available for the Subaru HDS data of HD 209458, which we discuss in the following Section. 5.5 Summary The Fourier analysis method is generally more sensitive to smaller planet-to-star fluxes than the Collier Cameron method for the ideal test spectra presented here, due to the ability to extract the planet spectrum without having to remove a modelled version of the stel- lar spectrum. At the 1σ level, the Fourier analysis and the Col- lier Cameron method are similarly sensitive to small planet-to-star fluxes in the blue wavelength range. In the red wavelength range, the Fourier analysis method is around three times more sensitive than the Collier Cameron method at the 1σ level. At the 2 and 3σ levels, the Fourier analysis method is twice as sensitive as the Col- lier Cameron method in the blue wavelength range. The Fourier analysis method utilises the known parameters in the transiting planet case, and allows a more sensitive measurement from a small portion of the orbit that can be observed over a short amount of time. This is useful for observations with ground-based telescopes for follow up observations to space-based discoveries. The Fourier analysis method developed here is best suited to studying hot Jupiter planets with orbital periods of a few days, and spectra taken of bright host stars in the blue wavelength range, just before or after a secondary eclipse. Observations of transit- ing hot Jupiter planet's should be planned for when the planet is furthest from the observer and therefore appears close to full phase and maximum reflectance. The planet should have a velocity large enough to shift the scattered spectrum clear of the corresponding stellar absorption lines. When the wavelength of the Fourier com- 10 Langford et al. A B Figure 5. Signal-to-noise ratio for mock spectra with the Fourier analysis method. Contour plot of 1σ, 2σ and 3σ levels of detection (solid) for signal-to-noise ratio values in the range 102 -- 103, and ǫ values in the range 10−5 -- 2 × 10−4. Panel A - Results for the red wavelength range 560 -- 608 nm. Panel B - Results for the blue wavelength range 440 -- 515 nm. The dots represent the sampling of the parameter space. The dot-dashed line shows the signal-to-noise ratio value of 350 for the HD 209458 data. A 1σ level of detection for this data requires a planet-to-star flux greater than ǫ = 1 × 10−4 in the red wavelength range, and ǫ = 4 × 10−5 in the blue wavelength range. The dashed line shows the upper limit on the planet-to-star flux of HD 209458b set by Rowe et al. (2008) of ǫ < 1.6 × 10−5. ponent is some multiple of a precisely uniform velocity spacing of the observations, the matrices become singular and the solutions can blow up. Therefore, observations should be planned with some intentional irregularity in velocity differences. Observations taken over a short velocity range are beneficial to reduce the variation due to the unknown planet phase function. The relatively small number of exposures required to obtain a measure- ment of the planet-to-star flux with the Fourier analysis method is therefore beneficial for rapid ground-based follow up to detections by space-based surveys. The Fourier analysis method is also suited to measuring the planet-to-star flux over short wavelength ranges and therefore determining the likelihood of hot Jupiter planets hav- ing wavelength dependent albedos in the optical. dimensional spectra need to be normalised for variation in the ef- ficiency of the CCD between the peak at the centre and the low count level close to the edge. An 11-knot cubic-spline was used to normalise the large scale variation of the response function of the red and blue CCDs, without removing the weak planet signal and the absorption lines. The simplest method of combining the or- ders of each CCD by averaging the overlapping regions of the nor- malised spectra can magnify the noise and was therefore avoided. Instead, the object spectra and the response functions were sep- arately summed. The combined spectra were then divided by the combined response function to remove large scale profiles. This preserves the signal in the overlapping regions. Bad pixels in the centre of the red CCD and at the edges of the blue CCD limit the full wavelength range possible for a continuous spectrum. 6 SUBARU HDS DATA In this Section we use Spectra of HD 209458 and HD 189733 to test the Fourier analysis method for detecting light scattered by a transiting extrasolar planet with ground-based capabilities. This data set includes 32 ´echelle spectra from observations taken 2002 October 26 of HD 209458 with Subaru HDS with an 0.8′′ slit (Noguchi et al. 2002) and 47 ´echelle spectra from observations taken 2008 August 26 of HD 189733 with Subaru HDS with a 0.4′′ slit. The total wavelength ranges of the CCDs are 554 -- 680 nm for the red, and 415 -- 550 nm for the blue when the orders are com- bined. The HD 209458 data has a spectral resolution of R = λ/∆λ = 45, 000, and a signal-to-noise ratio per pixel of ∼350 for a 500 second exposure. For details of the observations of HD 209458b see Winn et al. (2004) and Narita et al. (2005). The HD 189733 data has a spectral resolution of R = λ/∆λ = 90, 000 and a typical signal-to-noise ratio per pixel of ∼300 for a 500 sec- ond exposure. During the observations HD 209458b had just left a secondary eclipse and had velocities in the range 30 -- 80 km s−1 and HD 189733b had velocities of 60 -- 130 km s−1 just prior to a secondary eclipse. The ´echelle spectra taken were processed using standard IRAF procedures, following a correction for the non-linearity of the HDS CCDs using the function measured by Tajitsu2010. The one- 6.1 Correction for Instrumental Variations Before the combination of the dispersed orders, the variability of the spectra with time due to instrumental effects can be re- moved via a CCD response and wavelength correction as outlined in Winn et al. (2004). The mean ratio of two spectra taken at dif- ferent times should be equal to unity. This is not the case for the Subaru HDS data, as the ratio changes with wavelength bin. The spectrum format is affected by instrumental effects such as flexure of the spectrograph, changes in the temperature, changes in the in- clination of the ´echelle and cross-dispersion gratings and the flux calibration (Aoki 2002). The temperature can change by around 0.1 degrees per hour, causing a 0.14 CCD pixel shift of the spectrum. Changes in the collimator mirror or the inclinations of the gratings can cause up to 1 pixel difference in the repeatability of the spectrograph set up. The largest variation within a night of observations comes from the flux calibration, which can vary the profile of the orders of the spectra in different exposures by up to 10 per cent (Suzuki et al. 2003). The variation between exposures increases with time, such that the largest variation is seen between the first and last exposure. Following the process outlined in Winn et al. (2004), the pattern of the variation of the ratio with wavelength bin can be used to correct Detecting starlight scattered by transiting hot Jupiters 11 adjacent orders to scale the flux calibration to match the first expo- sure. The ratio of the nth order of two spectra in the jth wavelength bin is given by; R(λn,j) = S1(λn,j) S2(λn,j) , (14) where R signifies the ratio, S1 is the first exposure and S2 is a later exposure (Winn et al. 2004). The ratios of the exposures for orders adjacent to the one being corrected are boxcar smoothed over j, with a width of 100 pixels. The flux corrected nth order of the second spectrum S2b(λn) is given by; S2b(λn,j ) = [cn+1R′(λn+1,j ) + cn−1R′(λn−1,j )]S2(λn,j), (15) where R′(λn) signifies the boxcar smoothed ratio of the nth order of the spectra (Winn et al. 2004). Linear regression is used to find the factors cn+1 and cn−1 such that the sum of squared residuals is a minimum between S1(λn) and the corrected second spectrum. A correction can also be made for small variations in wave- length. These variations are of the order of a pixel, corresponding to a velocity shift of less than 2 kms−1, compared to an average planet velocity of 70 kms−1. The nth order of the matched spec- trum S2m(λn) is given by; S2m(λn,j ) = coS2b(λn,j ) + ∆λ dS2b(λ) dλ , (16) where the derivative is determined via 3-point Lagrangian interpo- lation (Winn et al. 2004). Linear regression is used to find the fac- tors c0 and ∆λ such that the sum of squared residuals is a minimum between S1(λn) and the matched second spectrum. For the Subaru HDS data, orders at wavelengths shorter than the location of the bad pixels in the centre of the red CCD were corrected, corresponding to a range of 558.83 -- 599.56 nm. The blue wavelength range of 445.28 -- 508.11 nm was also corrected before normalisation and combination. 7 RESULTS In order to predict the wavelength shift of the scattered spectra with respect to the stationary stellar spectra, the velocity of the planet was determined from the physical parameters listed in Table 2 and the Julian date at the time of each exposure. During the observa- tions HD 209458b had just left a secondary eclipse and had veloci- ties toward the observer in the range 30 -- 80 km s−1 and an average planet phase of 0.88. Just prior to a secondary eclipse HD 189733b had velocities of 60 -- 130 km s−1 away from the observer and an average planet phase of 0.74. The expected level of sensitivity of the Fourier analysis method can be determined for the observational parameters and the Monte Carlo simulations presented in Section 5. Fig. 5 shows con- tours of the level of detection of planet-to-star fluxes in the range 10−5 -- 2× 10−4 and signal-to-noise ratio values of 102 -- 103 for the red and the blue wavelength ranges. The dot-dashed line indicates the observational signal-to-noise ratio for the Subaru HDS ´echelle spectra of HD 209458 of 350. The upper limit of ǫ < 1.6×10−5 set by Rowe et al. (2008) with the MOST satellite is indicated by the dashed line. Upper limits can be derived from non-detections based on these Monte Carlo results, however the previous upper limit for HD 209458b is below the 1σ level of detection for this data with the Fourier analysis method. The spectra were analysed in the red (558.83 -- 599.56 nm) and blue (445.28 -- 508.11 nm) wavelength ranges separately, as Figure 6. Probability curve for HD 189733 in the red wavelength range. P(ǫtrueǫspectra), normalised to an area of unity. The dashed lines show the 1σ, 2σ and 3σ upper limits on the planet-to-star flux given the results of the analysis of the HD 189733 spectra. These correspond to upper limits of ǫ < 4.6 × 10−4, ǫ < 6.1 × 10−4 and ǫ < 7.5 × 10−4 respectively in the wavelength range 558.83 -- 599.56 nm. the Fourier analysis method requires continuous spectra. For the red and blue wavelength ranges of the HD 209458 data, the solutions were ǫ = 7.45 × 10−4 and ǫ = -1.50 × 10−4 respectively. A negative planet-to-star flux arising from the noise, as seen in the blue wavelength range solution, is not physical but may be used to statistically constrain the upper limit on the albedo. The red result is larger than expected by theory and the upper limit on the planet-to-star flux of HD 209458b of ǫ < 1.6 × 10−5 set by Rowe et al. (2008). Similarly, for the red and blue wavelength ranges of the HD 189733 data, the solutions were ǫ = 2.45×10−4 and ǫ = -1.08×10−4 respectively. An upper limit can be set on the planet-to-star flux via proba- bility distributions from Monte Carlo simulations and the physical constraint that the actual planet-to-star flux must be positive. From Bayes' theorem we have: P(ǫtrueǫspectra) ∝ P(ǫspectraǫtrue)P(ǫtrue), where P(ǫtrueǫspectra) is the probability that the measured planet-to-star flux is the physically correct value given the data. P(ǫspectraǫtrue) is the likelihood that the data will result in the measured planet-to-star flux. The prior P(ǫtrue) constrains the physical planet-to-star flux to be positive. (17) Monte Carlo simulations were run for the simulated data with the exact observational parameters of the Subaru HDS data in or- der to determine P(ǫspectraǫtrue), the likelihood that the data will result in the measured planet-to-star flux. For the HD 209458 re- sults, 32 spectra were generated with a signal-to-noise ratio of 350, a spectral resolution of 45,000 and an orbital phase range of 0.55 -- 0.60. For the HD 189733 results, 25 spectra were generated with a signal-to-noise ratio of 300, a spectral resolution of 90,000 and an orbital phase range of 0.34 -- 0.44. the results of The likelihood of the HD 209458 and HD 189733 data in the blue wavelength range shows a very small probability of having a positive and therefore physical planet-to- star flux. This suggests that the data correction did not fully account for the time varying signal due to the high density of absorption lines in the blue wavelength range, and the results measured with the Fourier analysis method are spurious. Similarly, the results for the red wavelength range of the HD 209458 spectra are thought to be spurious as the likelihood shows a very small probability of having a physical planet-to-star flux below the previously defined upper limit of ǫ < 1.6 × 10−5 set by Rowe et al. (2008). Fig. 6 shows the results of the Bayesian analysis for the red wavelength range of the HD 189733 data. The measured planet-to- star flux was ǫ = 2.45×10−4. The likelihood from the Monte Carlo 12 Langford et al. Orbital Period Orbital Inclination Mass of Planet b Mass of Host Star Transit Epoch Semi-Major Axis Porb i Mp M⋆ T0 a HD 209458 system HD 189733 system 3.52474859 ± 0.00000038 days 86.929+0.009 −0.010 degrees 0.64 ± 0.06 MJ 1.101+0.066 −0.062 M⊙ 2, 452, 826.628521 ± 0.000087 days 0.04704±0.00048 0.00047 AU 2.218573 ± 0.00002 days 85.79 ± 0.24 degrees 1.154 ± 0.033 MJ 0.82 ± 0.03 M⊙ 2, 453, 988.80336 ± 0.00024 days 0.03099±0.0006 0.00063 AU Table 2. Physical parameters of the star HD 209458 with transiting planet HD 209458b (Knutson et al. 2007) and of the star HD 189733 with transiting planet HD 189733b (Bouchy et al. 2005; Bakos et al. 2006). simulations was fit with a Gaussian distribution and multiplied by the prior condition that the planet-to-star flux must be positive. The probability that the measured planet-to-star flux is the physically correct value given the data, P(ǫtrueǫspectra), is then normalised to have an area of unity. From this curve the 1σ, 2σ and 3σ upper limits on the planet-to-star flux of HD 189733b in the wavelength range 558.83 -- 599.56 nm are found to be ǫ < 4.6 × 10−4, ǫ < 6.1 × 10−4 and ǫ < 7.5 × 10−4 respectively. 8 ANALYSIS To investigate the level of spurious signals in the data we shuffled Doppler shifts with respect to the corresponding exposures. This was performed for both stars in both wavelength ranges. A distri- bution of planet-to-star fluxes was measured due to spurious time variable signals arising out of the instrumental noise. Monte Carlo simulations were then used to determine the distribution of false planet-to-star fluxes that are not correlated with the Doppler shifts of the planet signal. Random velocities were selected in the same range as the real data; in the range 30 -- 80 km s−1 for HD 209548 and 60 -- 130 km s−1 for HD 189733. This was also done for the mock data. The distribution of ǫmeasured values for randomly selected Doppler shifts are compared for both wavelength ranges of the HD 209458 and HD 189733 Subaru HDS spectra in Fig. 7. The dis- tributions for the corresponding simulated mock spectra are shown in Fig. 8. Although the data corrections for instrumental effects re- duce the spread of the spurious signals for the Subaru HDS spectra, it is apparent that the time variability of the spectrum has not been fully removed. This can be seen by the comparison of the spread of the false-planet signals for the mock data, compared to the cor- rected Subaru HDS data. Similar data corrections based on low order polynomial corrections to the CCD response and the cross-correlation of spectral lines have been adopted for spectra taken with Sub- aru HDS (Narita et al. 2005; Snellen et al. 2008), the High- Resolution Echelle Spectrograph (HIRES) on Keck (Suzuki et al. 2003) and the High Resolution Spectrograph (HRS) on the Hobby- Eberly Telescope (HET) (Redfield et al. 2008). Narita et al. (2005) found the correction in Winn et al. (2004) to be accurate to approximately the Poisson noise limit for the HD 209458 data analysed here, and that the main challenge in using ground-based spectra for high precision measurements is instrumental variation, not the removable telluric contribution. Suzuki et al. (2003) calibrate the flux of Keck HIRES spectra to within 1 per cent, correcting for vignetting and extraction using ref- erence spectra. This suggests that instruments with increased stabil- A B HD 189733 Blue HD 189733 Red HD 209458 Blue HD 209458 Red Figure 7. Probability curves. Histograms from Monte Carlo simulations showing the probability of measuring an ǫmeasured with shuffled planet velocities used to obtain the solutions. Panel A - Results for the HD 189733 Subaru data. Panel B - Results for the HD 209458 Subaru data. Upper sub- panels - Results for data in the blue wavelength range. Lower subpanels - Results for data in the red wavelength range. The dashed curves show the Gaussian distribution with the same mean and standard deviation. ity are required for more accurate ground-based follow up charac- terisation of extrasolar planets. The flux and wavelength calibration correction reduced the spread of the spurious signals seen in the distributions in Fig. 7 by an order of magnitude. In the mock data the blue wavelength ranges have narrower distributions of false planet signals, suggest- ing that the increased density of absorption lines in this region is more likely to constrain the moving signal to be correlated with the planet velocity. This is not seen in the HD 189733 corrected data. Detecting starlight scattered by transiting hot Jupiters 13 A B HD 189733 Blue HD 189733 Red HD 209458 Blue HD 209458 Red Figure 8. Probability curves. Histograms from Monte Carlo simulations showing the probability of measuring an ǫmeasured with shuffled planet velocities used to obtain the solutions. Panel A - Results for the HD 189733 mock data. Panel B - Results for the HD 209458 mock data. Upper sub- panels - Results for data in the blue wavelength range. Lower subpanels - Results for data in the red wavelength range. The dashed curves show the Gaussian distribution with the same mean and standard deviation. This may be partly due to the 11-knot cubic-spline chosen to re- move the shape of the response of the CCDs being less suitable for spectra with a high density of lines. Additionally, in practice, the wavelength scale and flux calibration correction may be harder to implement for the higher density of spectral lines in the blue wave- length range. The red wavelength range of the corrected HD 189733 spectra is the closest to the mock spectra. It can be seen in Fig. 7 and Fig. 8 that the two distributions of false-planets have a similar spread. Therefore the time variable signal due to instrumental effects has been satisfactorily removed, and the corrected data is comparable to the ideal simulated data. From this corrected data the 1σ up- per limit on the planet-to-star flux in the wavelength range 558.83 -- 599.56 nm is ǫ < 4.6 × 10−4. This upper limit is not sensitive enough to constrain atmospheric models of hot Jupiter planets or rule out a highly reflective upper cloud deck. These instrumental variations would also hinder the applica- tion of the Collier Cameron method to this Subaru HDS data. It is beyond the scope of this paper to apply the Collier Cameron method to the non-idealised spectra. Future searches for the light scattered by an extrasolar planet will continue to adopt the Collier Cameron method, and the method developed by Charbonneau et al. (1999). These methods are still required for non-transiting planets. For transiting planets, the application of these methods with con- strained parameters may yield more stringent upper limits. The red wavelength HD 189733 data set showed successful implementation of instrumental corrections to Subaru HDS spec- tra and the application of the Fourier analysis method. The archive data obtained to test the Fourier analysis method did not have suf- ficient signal-to-noise ratio to significantly constrain the planet-to- star flux of the systems. Improved stability of the instrument while observing bright systems with high signal-to-noise ratio will be re- quired for follow up characterisation of transiting systems detected by space-based missions via the Fourier analysis method. 9 CONCLUSION Measuring planet-to-star flux ratios will continue to guide theories on the composition of the upper atmosphere of hot Jupiter planets and methods of radiation transfer. This is complementary to the de- tections of thermal emission from transiting very hot Jupiter plan- ets in the optical and infrared. However, previous upper limits set by non-detections of the light scattered by systems such as τ Bootis (Charbonneau et al. 1999; Collier Cameron et al. 1999; Leigh et al. 2003a) and HD 75289A (Rodler et al. 2008; Leigh et al. 2003b) support the need for model independent approaches that exploit the benefits of the numerous transiting systems to minimse the un- known orbital parameters. It is also advantageous to measure phase independent planet-to-star fluxes without assuming a grey albedo. In this paper we have introduced a new model-independent method for isolating the scattered starlight signal from the host star flux in high resolution spectra, that is suited to typical transit data. Using mock spectra this Fourier analysis method was shown to be more appropriate for typical observations of a well constrained transiting system, and therefore more sensitive to smaller planet-to- star fluxes, than the currently used Collier Cameron method. Us- ing Monte Carlo simulations the sensitivity to dim planets of the Fourier analysis method was shown to increase with higher signal- to-noise ratio data, and to be better suited to the blue region of the visible spectrum due to the increased signal in Fourier space of the numerous deep absorption lines. The Fourier analysis method is more likely to detect the planet-to-star flux of a dim planet such as HD 209458b, with Subaru HDS and ideal high signal-to-noise ratio spectra, than the Collier Cameron method. The Fourier analysis method for extracting the light scattered by transiting hot Jupiter planets from high resolution spectra was applied to ´echelle spectra of HD 209458 and HD 189733. Due to instrumental variations that could not be fully corrected for, there was no improvement on the measurement of the upper limit of the planet-to-star flux of HD 209458 compared to the current limit set by Rowe et al. (2008). A 1σ upper limit on the planet-to-star flux of HD 189733b was measured in the wavelength range of 558.83 -- 599.56 nm of ǫ < 4.5× 10−4. This result does not constrain atmo- spheric models. A deeper measurement of the upper limit of planet- to-star flux of this system with ground-based capabilities requires data with a higher signal-to-noise ratio, and increased stability of the telescope. Radial velocity surveys observe stars in a brighter mag- nitude range than transiting photometry surveys (Borucki et al. 2003; Tinney et al. 2001; Udry et al. 2000). For example Kepler is surveying stars of magnitude 14 -- 9, as compared to the Anglo- Australian Planet Search (AAPS) on the Anglo-Australian Tele- scope (AAT) which targets stars brighter than V< 7.5. Therefore measuring the light scattered by the upper atmosphere of plan- 14 Langford et al. ets will most likely require candidates selected via radial velocity methods and followed up with transit photometry, to ensure that adequate signal-to-noise ratio is obtained. The next generation of extremely large telescopes (ELTs) in- cluding the European Extremely Large Telescope (E-ELT), the Gi- ant Magellen Telescope (GMT) and the Thirty Meter Telescope (TMT), will be the largest optical and infrared telescopes ever built, dramatically increasing the sensitivity of current ground- based telescopes (Carlberg 2005). The reflecting mirror diameters range from 25 -- 42 m. With a diameter of 40 m, the collecting area and hence the signal-to-noise will be increased by a factor of ∼5 compared to similar observations with current 8 m ground-based telescopes. This will enable higher signal-to-noise ratios to be ob- tained (Gilmozzi & Spyromillo 2007). Therefore, smaller planet- to-star fluxes of transiting hot Jupiter planets will be measurable with ELTs and the Fourier analysis method. ACKNOWLEDGMENTS NSO/Kitt Peak FTS data used here were produced by National Op- tical Astronomy Observatory/Association of Universities for Re- search in Astronomy/National Science Foundation This paper is partly based on data collected at Subaru Tele- scope, which is operated by the National Astronomical Observatory of Japan. We thank the anonymous referee for their detailed comments, helping us to improve the paper. S.V.L aknowledges the support of an Australian Postgradu- ate Award, an Albert Shimmins Postgraduate Writing-Up Award, a Postgraduate Overseas Research Experience Scholarship, a Mel- bourne Abroad Traveling Scholarship and the hospitality of Prince- ton University Observatory during two extended visits in 2007 and 2008. N.N. is supported by a Japan Society for Promotion of Science (JSPS) Fellowship for Research (PD: 20-8141). Y.S. acknowledges the support from the Global Scholars Pro- gram of Princeton University. This work is partially supported by Grant-in-Aid for Scientific research of Japanese Ministry of Educa- tion, Culture, Sports, Science and Technology (No.20340041), and by JSPS Core-to-Core Program "International Research Network for Dark Energy". REFERENCES Alonso R., Guillot T., Mazeh T., Aigrain S., Alapini A., Barge P., Hatzes A., Pont F., 2009b, Astron. Astrophys., 501, L23 Alonso R. et al., 2009a, Astron. Astrophys., 506, 353 Aoki W., 2002, http://www/subarutelescope.org/Observing/ In- struments/HDS/index.html Bakos G. et al., 2006, Astrophys. J., 650, 1160 Berdyugina S., Berdyugin A., Fluri D., Piirola V., 2008, Astro- phys. J., 673, L83 Berdyugina S., Berdyugin A., Fluri D., Piirola V., 2011, Astro- phys. J., 728, L6 Borucki W. et al., 2003, ASPC, 294, 427 Bouchy F. et al., 2005, Astron. Astrophys., 444, L15 Burrows A., Ibgui L., Hubeny I., 2008, Astrophys. J., 682, 1277 Carlberg R., 2005, Proceedings of the International Astronomical Union. Vol 1. Charbonneau D., Noyes R., Korzennik S., Nisenson P., Jha S., Vogt S., Kibrick R., 1999, Astrophys. J., 522, L145 Christiansen J. et al., 2010, Astrophys. J., 710, 97 Collier Cameron A., Horne K., James D., Penny A., Semel M., 2000, in Penny Artymowicz Lagrange Russell eds, Planetary systems in the Universe. Proc. IAU Symp. Vol. 202, τ boo b: Not so bright, but just as heavy. Collier Cameron A., Horne K., Penny A., James D., 1999, Nature, 402, 751 Collier Cameron A., Horne K., Penny A., Leigh C., 2002, Mon. Not. R. Astron. Soc., 330, 187 Frigo M., Johnson S., 2005, Proc. IEEE, 93, 216 Gilmozzi R., Spyromillo J., 2007, The Messenger, 127, 11 Green D., Matthews J., Seager S., Kuschnig R., 2003, Astrophys. J., 597, 590 Hadrava P., 1995, Astron. Astrophys. , 389, 257 Hood B., Wood K., Seager S., Collier Cameron A., 2008, Mon. Not. R. Astron. Soc., 389, 257 Jenkins E., 2002, Astrophys. J., 580, 938 Kalas P. et al., 2008, Science, 322, 1345 Kipping D., Bakos G., 2011, Astrophys. J., 730, 50 Knutson H., Charbonneau D., Noyes R., Brown T., Gilliland R., 2007, Astrophys. J., 655, 564 Kupka F., Piskunov N., Ryabchikova T., Stempels H., Weiss W., 1999, Astron. Astrophys., 138, 119 Leigh C., Collier Cameron A., Guillot T., 2003c, Mon. Not. R. Astron. Soc., 346, 890 Leigh C., Collier Cameron A., Horne K., Penny A., James D., 2003a, Mon. Not. R. Astron. Soc., 344, 1271 Leigh C., Collier Cameron A., Udry S., Donati J.-F., Horne K., James D., Penny A., 2003b, Mon. Not. R. Astron. Soc., 346, L16 Liu X. et al., 2007, preprint (astro-ph/0711.2304) Marley M., Gelino C., Stephens D., Lunine J., Freedman R., 1999, Astrophys. J., 513, 879 Narita N. et al., 2005, PASJ, 57, 471 Noguchi K. et al., 2002, PASJ, 54, 855 Redfield S., Endl M., Cochran W., Koesterke L., 2008, Astrophys. J., 673, L87 Rodler F., Kurster M., Henning T., 2008, Astron. Astrophys., 485, 859 Rodler F., Kurster M., Henning T., 2010, Astron. Astrophys., 514, 23 Rogers J., Apai D., L´opez-Morales M., Sing D., Burrows A., 2009, Astrophys. J., 707, 1707 Rowe J. et al., 2006, Astrophys. J., 646, 1241 Rowe J. et al., 2008, Astrophys. J., 689, 1345 Seager S., Whitney B., Sasselov D., 2000, Astrophys. J., 540, 504 Sing D., L´opez-Morales M., 2009, Astron. Astrophys., 493, L31 Snellen I., Albrecht S., de Mooij E., Le Poole R., 2008, Astron. Astrophys., 487, 357 Sudarsky D., Burrows A., Hubeny I., 2003, Astrophys. J., 588, 1121 Sudarsky D., Burrows A., Pinto P., 2000, Astrophys. J., 538, 885 Suzuki N., Tytler D., Kirkman D., O'Meara J., Lubin D., 2003, PASP, 115, 1050 Tajitsu A. et al., 2010, PNAOJ In Press. Tinney C., Butler R., Marcy G., Jones H., Penny A., Vogt S., Apps K., Henry G., 2001, Astrophys. J., 551, 507 Udry S. et al., 2000, Astron. Astrophys., 356, 590 Wiktorowicz S., 2009, Astrophys. J., 696, 1116 Winn J., Suto Y., Turner E., Narita N., Frye B., Aoki W., Sato B., Yamada T., 2004, PASJ, 56, 655 This paper has been typeset from a TEX/ LATEX file prepared by the author. Detecting starlight scattered by transiting hot Jupiters 15
1809.02548
2
1809
2019-02-09T18:27:12
The Influence of H2O Pressure Broadening in High Metallicity Exoplanet Atmospheres
[ "astro-ph.EP" ]
Planet formation models suggest broad compositional diversity in the sub-Neptune/super-Earth regime, with a high likelihood for large atmospheric metal content (> 100 x Solar). With this comes the prevalence of numerous plausible bulk atmospheric constituents including N2, CO2, H2O, CO, and CH4. Given this compositional diversity there is a critical need to investigate the influence of the background gas on the broadening of the molecular absorption cross-sections and the subsequent influence on observed spectra. This broadening can become significant and the common H2/He or "air" broadening assumptions are no longer appropriate. In this work we investigate the role of water self-broadening on the emission and transmission spectra as well as on the vertical energy balance in representative sub-Neptune/super-Earth atmospheres. We find that the choice of the broadener species can result in a 10 -- 100 parts-per-million difference in the observed transmission and emission spectra and can significantly alter the 1-dimensional vertical temperature structure of the atmosphere. Choosing the correct background broadener is critical to the proper modeling and interpretation of transit spectra observations in high metallicity regimes, especially in the era of higher precision telescopes such as JWST.
astro-ph.EP
astro-ph
Draft version February 12, 2019 Typeset using LATEX twocolumn style in AASTeX62 9 1 0 2 b e F 9 . ] P E h p - o r t s a [ 2 v 8 4 5 2 0 . 9 0 8 1 : v i X r a The Influence of H2O Pressure Broadening in High Metallicity Exoplanet Atmospheres Ehsan Gharib-Nezhad1 and Michael R. Line2 1School of Molecular Sciences, Arizona State University, Tempe, AZ 85287, USA. 2School of Earth and Space Exploration, Arizona State University, Tempe, AZ 85287, USA. (Accepted December 21, 2018) Submitted to The Astrophysical Journal ABSTRACT Planet formation models suggest broad compositional diversity in the sub-Neptune/super-Earth regime, with a high likelihood for large atmospheric metal content (≥ 100 × Solar). With this comes the prevalence of numerous plausible bulk atmospheric constituents including N2, CO2, H2O, CO, and CH4. Given this compositional diversity there is a critical need to investigate the influence of the background gas on the broadening of the molecular absorption cross-sections and the subsequent influence on observed spectra. This broadening can become significant and the common H2/He or "air" broadening assumptions are no longer appropriate. In this work we investigate the role of water self-broadening on the emission and transmission spectra as well as on the vertical energy balance in representative sub-Neptune/super-Earth atmospheres. We find that the choice of the broadener species can result in a 10 -- 100 parts-per-million difference in the observed transmission and emission spectra and can significantly alter the 1-dimensional vertical temperature structure of the atmosphere. Choosing the correct background broadener is critical to the proper modeling and interpretation of transit spectra observations in high metallicity regimes, especially in the era of higher precision telescopes such as JWST. Keywords: planets and satellites: atmospheres , planets and satellites: composition, molecular data 1. INTRODUCTION A primary goal of exoplanet science is the determina- tion of basic planetary conditions. Transit spectropho- tometry observations of planetary atmospheres offer a window into fundamental quantities such as climate and composition e.g Madhusudhan et al. (2016)). Determin- ing atmospheric composition is a necessary requirement for assessing the relative importance of various chemi- cal processes (Moses 2014) and greatly assists in under- standing planet formation by linking volatile inventory to proto-planetary disk processes ( Oberg et al. 2011; Mordasini et al. 2016). One of the key findings of the Kepler Mission (Borucki et al. 1997) is that a majority of exoplanets fall within this "warm sub-Neptune" regime (∼2 -- 4 Earth radii, T<1000 K ) (Fressin et al. 2013). These planets have been an intense area of focus for transit spectra obser- vations with the Hubble Space Telescope (HST) (Krei- dberg et al. 2014a; Fraine et al. 2014; Knutson et al. 2014) and will be over the next decade as they serve as the link between jovian worlds and terrestrial planets as well as being the most prolific population of planets to be found by the Transiting Exoplanet Explorer Satellite (TESS, (Sullivan et al. 2015; Louie et al. 2018; Barclay et al. 2018; Kempton et al. 2018)). Planet formation, interior structure, and atmospheric chemistry modeling (Fortney et al. 2013; Moses et al. 2013; Lopez & Fortney 2014) suggest extreme composi- tional diversity within this sub-population, with a high likelihood for large atmospheric metallicities (¿300× So- lar). Given this potential for compositional diversity, the assumption of "jovian-like" H2/He-dominated atmo- spheres may not always be appropriate. Instead, with currently measured atmospheric metallicities reaching as high as ∼300-1000× solar (Line et al. 2014; Fraine et al. 2014; Kreidberg et al. 2014b; Knutson et al. 2014; Morley et al. 2017), molecules such as H2O and CO2 will become the dominant bulk constituents (Moses et al. 2013; Hu & Seager 2014). Along with this diversity in composition, comes with it numerous challenges in atmospheric modeling ranging from chemical modeling (Hu & Seager 2014) to cloud microphysics (Ohno & Okuzumi 2018) to 3D climate 2 Gharib-Nezhad & Line modeling (Kataria et al. 2014). Nearly all flavors of atmospheric modeling that aim to make observational predictions necessarily require radiative transfer compu- tations. A key necessary ingredient in radiative trans- fer computations is the opacities, which for planets, are dominated by the molecular absorption cross-sections (hereafter, ACS). The ACS of a given molecule typi- cally consist of billions of lines representing the abil- ity of a molecule to absorb or emit photons. Each line has its own line-width (or broadening) typically spec- ified through the degree of thermal/Doppler and pres- sure broadening (Goody & Yung 1995). Pressure broad- ening is the net cumulative effect of interactions be- tween the absorbing molecule in question (e.g., H2O) with its neighboring molecules (or bath gases, e.g., H2, He) or by self-broadening (H2O with itself). Much exo- atmospheric relevant ACS focus, specifically broadening, has been jovian-centric (e.g., H2/He dominated compo- sitions and broadening Freedman et al. (2008); Tennyson et al. (2016); Grimm & Heng (2015); Hedges & Mad- husudhan (2016)) being largely driven by the abundance of high fidelity "hot-Jupiter" observations and carry over from brown dwarf modeling. Exploration of pressure broadening assumptions in exo-atmospheres is not new (e.g., Grimm & Heng (2015); Hedges & Madhusudhan (2016)). Hedges & Madhusud- han (2016) provide a comprehensive overview of the various pressure broadening effects including resolution, line-wing cutoff, Doppler vs. pressure, and more rele- vant to our investigation, an initial look at the impact of a broadener choice. They too explore the impact of H2O vs H2 broadening on the H2O ACS, specifically over HST wavelengths, and find that the band-averaged ACS can change by up to an order-of-magnitude. In this letter we expand upon the work in Hedges & Madhusudhan (2016) to not only determine the influ- ence of H2O self-broadening on the H2O ACS, but also as a function of water fraction, and more importantly we quantitatively assess the integrated effect that the broadener choice has on the observable spectra as well as on the impact on atmospheric vertical energy balance. This work is crucial to the proper interpretation of tran- sit spectra observations in high metallicity regimes, ex- pected of the sub-Neptune/Super-Earth population. In §2 we describe our data sources and how we compute the ACS and the transmission/emission spectrum and self- consistent modeling approach. In §3 we compare the im- pact of H2O self-broadening with the standard H2/He broadening assumption. Finally, in §4 we discuss the implications and future prospects. We also make our newly computed water ACS grid for both broadeners publically available1. 2. METHODS In this initial investigation on the impact of non H2/He foreign broadening on transmission/emission spectra, we choose to focus on H2O because: 1) H2O is the most prominent absorber in exoplanet spectra due to its large abundance over a range of elemental compositions (Moses et al. 2013) and multiple strong absorption bands from the optical to far infrared wave- lengths and 2) it shows the largest sensitivity to choice of broadener when compared to other species (a factor of ∼7 increase in broadening when compared to H2/He, Table 1). Table 1. Lorentzian half-width coefficients γL [cm−1/bar]∗,† for relevant broadeners. The focus of this work is on influ- ence H2O self and H2/He broadening on the H2O absorption cross-sections (bold). Absorber Broadener γL relative to γH2/He L H2O CH4 CO2 CO Self ‡ H2/He ¶ CO2 air Self H2/He H2O CO2 air Self H2/He H2O air Self H2/He H2O CO2 air 0.3 -- 0.54 0.05 -- 0.08 0.15 -- 0.20 0.08 -- 0.1 0.06 -- 0.09 0.05 -- 0.08 0.06 -- 0.09 0.07 -- 0.09 0.02 -- 0.07 0.08 -- 0.12 0.09 -- 0.12 0.10 -- 0.14 0.05 -- 0.08 0.04 -- 0.09 0.04 -- 0.08 0.07 -- 0.1 0.09 -- 0.1 0.05 -- 0.07 7× 1× 3× 1.5× 1.5× 1× 1.5× 1.5× 1× 2× 2× 2.5× 1× 1× 1× 1.5× 1.5× 1× half-width ΓL = (T /296) −nT (cid:80) ∗ The pressure broadening/Lorentzian line profile is defined with a L is the Lorentzian coefficient for broadener, b, Pb is the partial pressure of broadener b, and nT is the thermal coefficient (typically 0.5 under kinetic theory). † Data extracted from Refs. in Table 3 of (Hartmann et al. 2018) and from (Gordon et al. 2017). ‡ Denoted by H2O@[self] in the text and figures. ¶ Denoted by H2O@[H2+He] in the text and figures. b γb L Pb where γb The approach is to compute the H2O ACS under different end-member compositional scenarios, with the first the standard "jovian-like" H2/He broadening (H2O@[H2+He]) and the second, pure H2O broadening 1 LINK:TBD UPON ACCEPTANCE The Influence of H2O Pressure-Broadening ... 3 (H2O@[self]), which is more appropriate for high metal- licity or all-steam atmospheres. We then determine the spectral differences between H2/He and self-broadening of H2O in representative atmospheres. 2.1. Computation of pressure-broadened H2 16O absorption cross-section Table 2. Grid and computational assumptions over which the H16 2 O cross sections are computed. There are 270 T-P combinations and two broadener choices (H2+He vs. H2O). A variable wavenumber resolution is chosen to properly sam- ple the Voigt-widths at each given T-P pair. Finer sampling results in negligible differences in the ACS. 1: 2: 85% H2 100% H2O 15% He ACS T(K) P(bar) Case Case 400 900 10−6 10−3 1 Resolution∗ 100 1000 425 475 1000 1100 3×10−6 10−5 3×10−3 10−2 3 10 -- -- P≥1 P<1 500 575 1200 1300 3×10−5 10−4 3×10−2 10−1 100 30 cm−1 cm−1 1000 30000 650 725 800 1400 1500 3×10−4 3×10−1 300 1/ΓV 2/ΓV : : cm−1 cm−1 bar: bar: Line wing cut-off† (cid:113) ∗ ΓV is the Voigt half-width approximated as 0.5346ΓL + G, with ΓG the Doppler width (Olivero & Long- bothum 1977). † The Lorentz wing shape may not be appropriate out at such distances (Freedman et al. 2008) 100 300 0.2166Γ2 L + Γ2 We utilize the freely-available EXOMOL (Tennyson et al. 2016) line-list data (e.g., BT2 line list (Barton et al. 2017)) and EXOCROSS2 routine (Yurchenko et al. 2017) to compute the pressure-broadened H2O ACS database (Table 2) for two broadening scenarios: 1) 85% H2 and 15% He (current standard assumption) using the J-dependent pressure coefficients from EXOMOL (Bar- ton et al. 2017), and 2) 100% H2O using the average value of available experimental self-broadening coeffi- cients (Ptashnik et al. 2016). 2.2. Modeling the Impact on Transmission/Emission Spectra of Transiting Exoplanets To assess the signifigance of the type-of-broadener as- sumption, we use the CHIMERA (Line et al. 2013, 2014; Stevenson et al. 2014; Kreidberg et al. 2015; Line & Par- mentier 2016; Kreidberg et al. 2018) code with our newly generated ACS (converted to λ/∆λ=100 correlated-K 2 https://github.com/Trovemaster/exocross coefficients (Amundsen et al. 2016)) to model tran- sit/eclipse spectra of a representative sub-Neptune like planet (GJ1214b (Harpsøe et al. 2013), Teq=500 -- 900K). We first generate forward model spectra using both sets of ACS (H2O@[self] and H2O@[H2+He]) given a fixed temperature-pressure profile (TP, Guillot (2010) Eqs. 24, 49 )3 and either 100% H2O or 500×Solar metallicity under thermochemical equilibrium4. Second, we com- pute a self-consistent radiative equilibrium atmosphere5 (Arcangeli et al. 2018; Mansfield et al. 2018; Kreidberg et al. 2018) to determine the impact of water broaden- ing on the vertical energy balance and, in turn, on the observed spectra. We discuss our findings in the next section. 3. RESULTS 3.1. Impact on Cross Sections Figure 1 illustrates the effect of temperature, pressure, and water abundance on the difference between @[self] and @[H2+He] broadened ACS near 6 µm. The top panel shows how broadening changes with temperature at a fixed pressure of 1 mbar. Differences are largest for cooler temperatures where pressure broadening becomes more important. The middle panel illustrates the im- pact of variable pressure at a fixed temperature (725K). Even at low pressures (1 µbar) pressure broadening dif- ferences are still present in the line wings. The bottom panel shows the effect of varying water abundance on the combined @[self]+@[H2] broadening at a fixed tem- perature and pressure (725K, 1 mbar). With pure self- broadening, differences in the line wings can approach an order of magnitude. For a ∼30% mole fraction of water, the ACS is about 3 -- 5× greater than pure hydro- gen broadening. While not shown, these differences be- come larger at longer wavelengths and smaller at shorter wavelengths due to the relative importance of Doppler- to-pressure broadening. 3.2. Direct Impact on Transmission/Emission Spectra More practically, Figure 2 summarizes the key impact of @[H2+He] versus @[self] broadening on the emission (top row) and transmission (bottom row) spectra of a 3 With κth = 3 × 10−2 cm2/g, γ = 0.1, Teq=500,700,900K, Tint=0K 4 NASA CEA2 (Gordon & Mcbride 1994) with scaled (Lodders et al. 2009) abundances. We include as opacities in this scenario H2/He broadened H2O, CH4, CO, CO2, NH3, H2S, Na, K, HCN, C2H2, TiO, VO, PH3, and H2 H2/He CIA (Freedman et al. 2014) 5 Zero internal heat flux, PHOENIX stellar model for GJ1214, and an equilibrium temperature of 550 K so as to keep tempera- tures at all layers within the valid cross-section temperature range of 400 -- 1500K 4 Gharib-Nezhad & Line In the all-steam atmospheres, emission differences (Figure 2a) are largest in the window regions (∼ 4µm, ∼ 10µm ). The increased flux for the @[H2+He] broad- ened ACS is because of the lower opacity, permiting flux from deeper, hotter layers to emerge (for a fixed TP). The increased opacity due to the @[self] broadening obscures the deeper/hotter layers, resulting in lowered fluxes at those wavelengths. These differences are, of course, strongly dependent upon the temperature struc- ture within in the atmosphere. As these spectra assume a fixed TP there is a difference in net radiated flux, which will most certainly have an influence on the radia- tive balance and thermal structure in the atmosphere, as discussed in §3.3. Transmission spectra tell a similar, albeit less dra- matic story with relative differences of ∼60 ppm across shown wavelength range. The "linear-like" slope in the differences (∆) with wavelength is due to the frequency dependence of Doppler-to-Pressure broadening. The effects at high metallicity (500×solar, Figure 2, right column) are less extreme (10 of ppm) due to the reduced abundance of H2O (10 -- 20%) and the signif- icant abundances of additional opacity sources (mainly CO2, CO, CH4, and H2/He). Furthermore, due to the reduced impact of H2O@[self] broadening (Figure 1), we expect an approximate (comparing 1 mbar line wings) reduction of 3 -- 5× to ∼< 10ppm in the transmission spectra. 3.3. Impact on Self-Consistent 1D Atmosphere Figure 3 shows the impact of self-broadening on the 1D radiative balance (and subsequent observational ef- fects) of a ∼550K planet under the all steam and 500×Solar scenarios. The @[self] broadening results in ∼100-180K hotter temperatures below the ∼1 mbar level and ∼60Kcooler above for the all steam scenario (Figure 3a). More intuitively, the increased @[self] mean opacity "shifts" the averaged thermal "τ =1" level to a ∼ 3× lower pressure in the all steam scenario. This shift is readily seen in the band averaged contribution functions (Figure 3a). A similar, but lesser, effect is seen in the 500×Solar metallicity scenario (up to ∼ 70 K) because the water abundance is lower by a factor of ∼ 5 (Figure 3b). The radiative response of the TP to the integrated flux differences (up to 40% for steam and 10% for 500×solar, green vs. red curves in Figure 3c,d ) between the @[self] vs. @[H2+He] acts to reduce the emission spectrum differences, however, to a still de- tectable 10s of ppm (Figure 3c,d). The transmission spectra (Figure 3e,f) show compara- ble differences (30 -- 40 ppm) to the 500 K scenario from Figure 2c,d. However, there are now two effects tak- Figure 1. Illustration of the impact of @[self] (blue) ver- sus @[H2+He] (red) on the absorption cross sections near 6µm. The top panel shows the influence of temperature on the broadening difference at a fixed represantative pressure of 1 mbar. At 1200K (1 mbar) the lines are purely Doppler broadened resulting in little effect. The middle panel shows the influence of pressure at a fixed temperature. The Doppler cores are negligible by 1 bar. The bottom panel shows the impact of the relative weighting of self versus H2 broaden- ing (e.g., composition dependence) at a fixed temperature and pressure. Cross-section differences are largest in the pressure-broadened line wings, with pure @[self] typically 1 order of magnitude larger. A factor of 5 in broadening difference occurs by the time the relative abundance of wa- ter reaches ∼30%. In general, @[self] broadening becomes more important at higher pressures, cooler temperatures, and longer wavelengths due to the increased prominence of pressure broadening over Doppler broadening. typical sub-Neptune under the assumption of a pure steam atmosphere (left column) and a 500×Solar metal- licity6 scenario (right column). Overall, we find that the differences are quite large, 10s to 100s of ppm, well within the detectable range of both HST (Kreidberg et al. 2014a), and certainly the James Webb Space Tele- scope (JWST, e.g., Greene et al. (2016); Bean et al. (2018)), especially for the anticipated windfall of such planets around bright stars (Sullivan et al. 2015). 6 While the water mixing ratio is only ∼10 -- 20% for these con- ditions, we still use the pure @[self]-broadened water ACS as it is still a more accurate approximation than pure @[H2+He] broad- ening Pressure EffectWater Partial Pressure EffectTemperature Effect The Influence of H2O Pressure-Broadening ... 5 Figure 2. Effect of water self-broadening (@[self], blue) compared to the standard H2/He broadening (@[H2+He], red) on pure steam (left column: a,c) and high metallicity (500× solar-right column: b,d) atmospheres with equilibrium temperatures of 500, 700, and 900K. The top row (a,b) compares emission spectra and the bottom row shows relative transmission spectrum differences (c,d). The bottom panel in each shows the relative spectral difference (∆). Differences range anywhere from a few 10s to a few 100s of ppm and show a strong wavelength dependence. ing place that create the transmission differences. The first is the scale height effect due to the differences in the TP (@[H2+He]-@[self], H2O TP). The second, as before, is the broadening differences. Both effects con- tribute equally to the overall differences in the transmis- sion spectra. Despite the self-consistent adjustment of the TP, differences in both emission and transmission are still above detectable levels (10s of ppm) 4. CONCLUSIONS The aim of this work was to determine the observable impact of broadener composition on observed transit- ing planet spectra, with application to < 1000K high- metallicity and all steam atmospheres, likely represen- tative of the sub-Neptune/Super-Earth population of planets. As a specific example, we focused on the differ- ence between H2/He broadening and self broadening on the water absorption cross-sections as water is typically the most prevalent species and absorber in planetary at- mospheres. From our analysis we arrive at the following key points: • Absorption cross section differences between water self and the standard assumed H2/He broadening are up to an order of magnitude in the pressure broadened line wings (similar to Hedges & Mad- husudhan (2016)), and is noticeable over a range of applicable temperatures and pressures. • The influence of self-broadening is composition de- pendent and non-linear, with ∼half of the differ- ence achieved by water mole fractions of∼30% for a representative temperature and pressure. abcdPure Steam500 x Solar 6 Gharib-Nezhad & Line Figure 3. Comparison of the @[self] (blue) vs. @[H2+He] (red) broadening in self-consistent 1D thermochemical-radiatve- equilibrium atmospheres for all steam (left column: a,c,e) and 500×solar (right column: b,d,f) composition. The top row (a,b) shows the derived radiative-equilibrium TP under each scenario. Thermal emission contribution functions averaged over represenative bands (Cont. Func. 5 -- 8, and 3.5 -- 4.3 µm) for each broadening scenario are shown in (a). Subplot (b) shows the thermochemical equilibrium mixing ratios along the @[self] TP for select species. Temperature differences can be up to 175K (20%) in the pure steam scenario and up to 70K in the 500×Solar scenario. The second row (c,d) shows the resultant secondary eclipse spectra and their differences below (∆). An additional emission spectrum (@[H2+He], @[self] TP-green), is shown in (c) and (d) assuming the same TP as the @[self] scenario in order to decouple the effects of the radiatively adjusted TP from the broadening differences. The last row (e,f) shows the resulting cloud free transmission spectra and relative differences. An additional transmission spectrum (@[H2+He], @[self] TP-green), is shown in (e) and (f) assuming the same TP as the @[self] scenario in order to decouple the effects of the broadening and scale height change due to TP variation. Spectral differences are on the order of 30-40 ppm in transmission but are much less in emission (∼60 ppm) when compared to Figure 2a,b due to the radiative adjustment of the TP. • Transmission and emission spectra differences for representative sub-Neptune atmospheres range be- tween a few 10s of ppm up to 100s of ppm, de- Pure Steam500 x Solarabcdef The Influence of H2O Pressure-Broadening ... 7 pending upon wavelength, temperature, and wa- ter abundance. These differences are not negligi- ble considering currently achieved HST precisions of ∼15 ppm and possible precisions as low as a few ppm for JWST. Differences will vary depend- ing upon additional parameters like temperature gradient (for emission), planet-to-star radius ra- tio, and scale height. • The assumption of water self-broadening (or lack thereof) can have a significant impact on the 1D vertical energy balance, with temperature differ- ences of up to 180K in pure steam atmospheres (or a half-a-decade lower pressure shift in the emis- sion levels) and 10s of K in high metallicity atmo- spheres. This work is certainly not an exhaustive exploration of all possible broadening (Table 1) or planetary atmo- sphere conditions. However, it serves to illustrate that the broadener assumption can have a non-negligible im- pact on the observables and continues to illustrate the importance and key role of laboratory data on planetary atmosphere modeling. (Fortney et al. 2016) 5. ACKNOWLEDGEMENTS EGN and MRL thank J. Lyons, A. Heays, R. Freed- man, M. Marley, J. Fortney, P. Molli`ere, and L. Pino for many useful discussions. We especially thank S. Yurchenko for invaluable assistance with the EX- OCROSS code, and ASU Research Computing center for the kind support on computational side. MRL ac- knowledges summer support from the NASA Exoplanet Research Program award NNX17AB56G. This work benefited from numerous conversations at the 2018 Ex- oplanet Summer Program in the Other Worlds Labora- tory (OWL) at the University of California, Santa Cruz, a program funded by the Heising-Simons Foundation. REFERENCES Amundsen, D. S., Mayne, N. J., Baraffe, I., et al. 2016, A & Gordon, I., Rothman, L., Hill, C., et al. 2017, JQSRT, 203, A, 595, A36 3 Arcangeli, J., D´esert, J.-M., Line, M. R., et al. 2018, ApJ, 855, L30 Gordon, S., & Mcbride, B. J. 1994, Technical Report Greene, T. P., Line, M. R., Montero, C., et al. 2016, ApJ, Barclay, T., Pepper, J., & Quintana, E. V. 2018, ArXiv 817, 17 e-prints, arXiv:1804.05050 Barton, E. J., Hill, C., Yurchenko, S. N., et al. 2017, JQSRT, 187, 453 Grimm, S. L., & Heng, K. 2015, ApJ, 808, 182 Guillot, T. 2010, A&A, 520, A27 Harpsøe, K. B. W., Hardis, S., Hinse, T. C., et al. 2013, Bean, J. L., Stevenson, K. B., Batalha, N. M., et al. 2018, A&A, 549, A10 arXiv:1803.04985 Hartmann, J.-M., Tran, H., Armante, R., et al. 2018, Borucki, W. J., Koch, D. G., Dunham, E. W., & Jenkins, JQSRT, 213, 178 J. M. 1997, in ASP Conf. Ser., Vol. 119, Planets Beyond the Solar System and the Next Generation of Space Missions, ed. D. Soderblom, 153 Hedges, C., & Madhusudhan, N. 2016, MNRAS, 458, 1427 Hu, R., & Seager, S. 2014, ApJ, 784, 63 Kataria, T., Showman, A. P., Fortney, J. J., Marley, M. S., Fortney, J. J., Mordasini, C., Nettelmann, N., et al. 2013, & Freedman, R. S. 2014, ApJ, 785, 92 ApJ, 775, 80 Kempton, E. M.-R., Bean, J. L., Louie, D. R., et al. 2018, Fortney, J. J., Robinson, T. D., Domagal-Goldman, S., ArXiv e-prints, arXiv:1805.03671 et al. 2016, arXiv:1602.06305 Knutson, H. A., Benneke, B., Deming, D., & Homeier, D. Fraine, J., Deming, D., Benneke, B., et al. 2014, Nature, 2014, Nature, 505, 66 513, 526 Kreidberg, L., Line, M. R., Thorngren, D., Morley, C. V., & Freedman, R. S., Lustig-Yaeger, J., Fortney, J. J., et al. Stevenson, K. B. 2018, ApJ, 858, L6 2014, ApJS, 214, 25 Kreidberg, L., Bean, J. L., D´esert, J.-M., et al. 2014a, Freedman, R. S., Marley, M. S., & Lodders, K. 2008, ApJS, Nature, 505, 69 174, 504 Kreidberg, L., Bean, J. L., D´esert, J. M., et al. 2014b, Fressin, F., Torres, G., Charbonneau, D., et al. 2013, ApJ, ApJL, 793, 2 766, 81 Goody, R. M., & Yung, Y. L. 1995, Atmospheric Radiation : Theoretical Basis., 2nd edn. (Oxford University Press), 536 Kreidberg, L., Line, M. R., Bean, J. L., et al. 2015, ApJ, 814, 66 Line, M. R., Knutson, H., Deming, D., Wilkins, A., & Desert, J.-M. 2013, ApJ, 778, 183 8 Gharib-Nezhad & Line Line, M. R., Knutson, H., Wolf, A. S., & Yung, Y. L. 2014, Oberg, K. I., Murray-Clay, R., & Bergin, E. A. 2011, ApJL, ApJ, 783, 70 Line, M. R., & Parmentier, V. 2016, ApJ, 820, 78 Lodders, K., Palme, H., & Gail, H.-P. 2009, 712 -- 770 Lopez, E. D., & Fortney, J. J. 2014, ApJ, 792, 1 Louie, D. R., Deming, D., Albert, L., et al. 2018, PASP, 130, 044401 Madhusudhan, N., Ag´undez, M., Moses, J. I., & Hu, Y. 2016, Space Sci. Rev., 205, 285 Mansfield, M., Bean, J. L., Line, M. R., et al. 2018, AJ, 743, L16 Ohno, K., & Okuzumi, S. 2018, ApJ, 859, 34 Olivero, J., & Longbothum, R. 1977, JQSRT, 17, 233 Ptashnik, I. V., McPheat, R., Polyansky, O. L., Shine, K. P., & Smith, K. M. 2016, JQSRT, 177, 92 Stevenson, K. B., Bean, J. L., Seifahrt, A., et al. 2014, AJ, 147, 161 156, 10 Sullivan, P. W., Winn, J. N., Berta-Thompson, Z. K., et al. Mordasini, C., van Boekel, R., Molli`ere, P., Henning, T., & 2015, ApJ, 809, 77 Benneke, B. 2016, ApJ, 832, 1 Morley, C. V., Knutson, H., Line, M., et al. 2017, AJ, 153, 86 Tennyson, J., Yurchenko, S. N., Al-Refaie, A. F., et al. 2016, JMS, 327, 73 Moses, J. I. 2014, Phil. Trans. Roy. Soc. A, 372, 20130073 Moses, J. I., Line, M. R., Visscher, C., et al. 2013, ApJ, Yurchenko, S. N., Tennyson, J., & Barton, E. J. 2017, J. Phys.: Conf. Ser., 810, 012010 777, 34
0901.2886
2
0901
2009-04-13T04:20:03
Electric charging of dust aggregates and its effect on dust coagulation in protoplanetary disks
[ "astro-ph.EP", "astro-ph.SR" ]
Mutual sticking of dust aggregates is the first step toward planetesimal formation in protoplanetary disks. In spite that the electric charging of dust particles is well recognized in some contexts, it has been largely ignored in the current modeling of dust coagulation. In this study, we present a general analysis of the dust charge state in protoplanetary disks, and then demonstrate how the electric charging could dramatically change the currently accepted scenario of dust coagulation. First, we describe a new semianalytical method to calculate the dust charge state and gas ionization state self-consistently. This method is far more efficient than previous numerical methods, and provides a general and clear description of the charge state of gas-dust mixture. Second, we apply this analysis to early evolutionary stages where the dust has been thought to grow into fractal ($D \sim 2$) aggregates with a quasi-monodisperse (i.e., narrow) size distribution. We find that, for a wide range of model parameters, the fractal growth is strongly inhibited by the electric repulsion between colliding aggregates and eventually "freezes out" on its way to the subsequent growth stage involving collisional compression. Strong disk turbulence would help the aggregates to overcome this growth barrier, but then it would cause catastrophic collisional fragmentation in later growth stages. These facts suggest that the combination of electric repulsion and collisional fragmentation would impose a serious limitation on dust growth in protoplanetary disks. We propose a possible scenario of dust evolution after the freeze-out. Finally, we point out that the fractal growth of dust aggregates tends to maintain a low ionization degree and, as a result, a large magnetorotationally stable region in the disk.
astro-ph.EP
astro-ph
SUMITTED TO APJ 2009 JANUARY 20; ACCEPTED 2009 APRIL 13 Preprint typeset using LATEX style emulateapj v. 08/13/06 ELECTRIC CHARGING OF DUST AGGREGATES AND ITS EFFECT ON DUST COAGULATION IN PROTOPLANETARY DISKS Graduate School of Human and Environmental Studies, Kyoto University, Yoshida-nihonmatsu-cho, Sakyo-ku, Kyoto 606-8501, Japan Sumitted to ApJ 2009 January 20; accepted 2009 April 13 SATOSHI OKUZUMI ABSTRACT Mutual sticking of dust aggregates is the first step toward planetesimal formation in protoplanetary disks. In spite that the electric charging of dust particles is well recognized in some contexts, it has been largely ignored in the current modeling of dust coagulation. In this study, we present a general analysis of the dust charge state in protoplanetary disks, and then demonstrate how the electric charging could dramatically change the currently accepted scenario of dust coagulation. First, we describe a new semianalytical method to calculate the dust charge state and gas ionization state self-consistently. This method is far more efficient than previous numerical methods, and provides a general and clear description of the charge state of gas-dust mixture. Second, we apply this analysis to compute the collisional cross section of growing aggregates taking their charging into account. As an illustrative example, we focus on early evolutionary stages where the dust has been thought to grow into fractal (D ∼ 2) aggregates with a quasi-monodisperse (i.e., narrow) size distribution. We find that, for a wide range of model parameters, the fractal growth is strongly inhibited by the electric repulsion between colliding aggregates and eventually "freezes out" on its way to the subsequent growth stage involving collisional compression. Strong disk turbulence would help the aggregates to overcome this growth barrier, but then it would cause catastrophic collisional fragmentation in later growth stages. These facts suggest that the combination of electric repulsion and collisional fragmentation would impose a serious limitation on dust growth in protoplanetary disks. We propose a possible scenario of dust evolution after the freeze-out. Finally, we point out that the fractal growth of dust aggregates tends to maintain a low ionization degree and, as a result, a large magnetorotationally stable region in the disk. Subject headings: dust, extinction -- methods: analytical -- planetary systems: formation -- planetary sys- tems: protoplanetary disks -- plasmas 1. INTRODUCTION The initial step toward planetesimal formation in protoplan- etary disks is the collisional growth of submicron dust grains into macroscopic aggregates. A standard scenario is that dust aggregates grow by mutual sticking, gradually settle to the midplane of the disk, and finally form a dense dust layer. It is still an open issue whether subsequent growth is established by the gravitational instability of the layer or the direct growth of the aggregates. To address this issue, further understanding on earlier evolutionary stages is needed. It has been recognized that the internal structure of ag- gregates is a key factor for their growth and settling. Early studies on dust coagulation modeled the aggregates as a compact, nonporous object (e.g., Weidenschilling 1980; Nakagawa et al. 1981). Both numerical simulations and lab- oratory experiments have revealed, however, that aggregates are not at all compact, but has an open, fluffy structure (for a review, see Meakin 1991; Blum 2004; Dominik et al. 2007). This is particularly true for aggregates formed at an early growth stage where the collisional velocity is too low for col- liding aggregates to compress each other. It has been ob- served in numerical (Kempf et al. 1999) as well as experi- mental (Wurm & Blum 1998; Blum et al. 1998, 2000) studies that the outcome is an ensemble of fractal aggregates with the fractal dimension D . 2 and with a quasi-monodisperse (i.e., narrow) mass distribution. This fractal growth typically lasts until the aggregates become centimeter-sized (Suyama et al. 2008). A remarkable dynamical property of these fluffy ag- gregates is that they keep a strong coupling to ambient gas and Electronic address: [email protected] thus a low drift velocity relative to the gas throughout the evo- lution. This could be crucial to the formation of very thin dust layer where planetesimals may be formed by gravitational in- stability. Dust grains and aggregates are not only the building block of planetesimals but also powerful absorbers of charged par- ticles in the gas disks. It is now widely accepted that tur- bulence in the disks is attributed to magnetorotational insta- bility (MRI; Balbus & Hawley 1991). For this mechanism to work, at least a part of the disk needs to be sufficiently ionized for the gas to couple to magnetic fields. Many au- thors have examined whether protoplanetary disks can be ionized enough to sustain MHD turbulence (Gammie 1996; Glassgold et al. 1997; Sano et al. 2000; Igea & Glassgold 1999; Ilgner & Nelson 2006a,b,c; Wardle 2007). One of the important findings is that the turbulent region is strongly con- trolled by the concentration of dust materials since they ef- ficiently remove away ionized particles from ambient gas (Sano et al. 2000; Ilgner & Nelson 2006a; Wardle 2007). Although the importance of dust charging is well recog- nized in the above context, its effect on dust coagulation in protoplanetary disks has been hardly examined. Charg- ing of aggregates causes electrostatic interaction between them, which may significantly increase or decrease the co- agulation rates. Recently, a series of studies have suggested that charge-induced dipole interaction might trigger runaway growth of dust aggregates (Ivlev et al. 2002; Konopka et al. 2005). However, these studies considered a situation where the ambient gas is not ionized and the net charge of dust ag- gregates vanishes identically. In protoplanetary disks, on the contrary, the net charge of dust aggregates does not vanish 2 OKUZUMI due to the presence of weakly ionized ambient gas, and there- fore both dust charging and gas ionization must be taken into account. This study explores how the electrostatic charging of dust aggregates could be crucial to their coagulation in protoplan- etary disks. For this purpose, we have to know in advance how the charge state of aggregates evolves with their growth. This is a complicated problem, since we also have to solve the ionization state of ambient gases self-consistently. Previ- ous studies (Sano et al. 2000; Ilgner & Nelson 2006a; Wardle 2007) have handled this problem with direct numerical cal- culations in which dust particles with different charges and sizes are treated as different charged species as well as many species of ions. However, this approach becomes inefficient when one tries to solve this problem and dust growth simul- taneously, since the dispersion of charge and size increases as the dust growth. The central strategy taken in this study is to solve this problem as analytically as possible. This approach does not only reduce the computational expense but also pro- vides general insight into the charge state of gas-dust mixture. As a result, we show that all the conditions for ionization- recombination equilibrium are reduced to a single algebraic equation. Just by solving this equation numerically, we can obtain both of the dust charge state and the gas ionization state analytically. We also confirm that the semianalytical calcula- tions agree very well with direct numerical calculations using the original equations. This semianalytical method will be a powerful tool for the simulations of charged dust coagulation and MRI turbulence. As an illustrative example, we calculate the collisional cross section of dust aggregates growing in a protoplane- tary disk taking into account their electric charging. We fo- cus on early stages of dust evolution where the aggregates has been thought to experience fractal, quasi-monodisperse growth (e.g., Blum 2004; Dominik et al. 2007). For a wide range of model parameters, we find that the effective cross section is quickly suppressed as the fractal growth proceeds and finally vanishes at a surprisingly early stage. This means that the fractal growth "freezes out" on its way to the sub- sequent growth stage where collisional compression of ag- gregates occurs. This is because the electrostatic repulsion between aggregates becomes strong enough to prevent their mutual collision. Strong turbulence in the disk will help the aggregates to overcome this electric barrier, but then it will cause catastrophic disruption of collided aggregates at later stages. Therefore, if the freeze-out of the fractal growth truly means the end of dust evolution, the combination of the elec- tric charging and the collisional disruption imposes a very strict limitation on dust coagulation and subsequent planetes- imal formation in protoplanetary disks. Our findings strongly suggest that the dust charing effect should be seriously taking into account in the modeling of dust evolution. This paper is organized as follows. In §2, we present a set of equations that describes the reactions of charged particles (ions, electrons, and dust aggregates), and derive the equa- tion that determines the equilibrium state. In §3, we calculate the electrostatic repulsion energy between two colliding ag- gregates to show that the quasi-monodisperse fractal growth is strongly inhibited for a wide range of disk parameters. In §4, we discuss the validity of some important assumptions and point out a possible scenario of dust evolution after the "freeze-out" of the fractal growth. A summary is presented in §5. gas (g) ... X = H2, He, (k) g cosmic rays radionuclides X-rays . . . electron (e) ion (i) X (k) i +X (k,l) g X (l) i ... +i +e Z= 1 +i +e +i +e Z=0 dust (d) Z=+1 +i +e ... ionization gas-phase recombination ion-neutral reaction absorption by dust FIG. 1. -- Schematic diagram of the ionization-recombination reactions in a gas-dust mixture. Ions and free electrons are created (solid arrows) by some ionizing sources (e.g., cosmic rays) and are removed through the gas-phase recombination (dotted arrows) or the adsorption to dust (double line arrows). Some species of ions may react with neutral gas particles to create different species of ions (dashed arrows). The equilibrium charge distribution nd(Z) of dust aggregates are determined by the balance between all these reactions. 2. EQUILIBRIUM CHARGE DISTRIBUTION 2.1. Kinetic equations for ionization-recombination reactions g We model the ionization-recombination reactions in a gas- dust mixture as follows (see also fig.1). Some ionizing sources (e.g. cosmic rays) create ions X(k) and free electrons i from neutral particles X(k) g at a rate ζ(k) (here k(= 1,2,···) la- bels each species of ions and associated neutrals). We assume that the ions and electrons are quickly thermalized and have thermal velocities u(k) and ue, respectively. We neglect the i possibility that the free electrons might be much more ener- getic in a MRI-active region (Inutsuka & Sano 2005). The ions may react with neutrals X(k,l) to produce another species of ions X(l) i (l 6= k), or may recombine with free electrons in the gas phase. We denote the rate coefficient for the ion-neutral ′(k,l) and β(k), reaction and the gas-phase recombination by β respectively. Also, the ions and free electrons may collide with dust aggregates to adsorb onto their surfaces. We write the collisional cross section for an aggregate and an ion (elec- tron) as σdi(de). These cross sections generally include the ef- fect of electrostatic interaction as well as the sticking proba- bility (see §2.2). Each dust aggregate may have different in- ternal structure and charge Ze from the other. We represent a set of parameters describing the structure (e.g., mass, radius) as I = {I1,I2,···}. In this section, we assume that the above reactions proceeds faster than the mutual collision of dust ag- gregates and treat I as constant parameters. The validity of this assumption is discussed in §4.1. The above charge reactions are described by a set of kinetic equations. Let us denote the number densities of ions X(k) , i free electrons, and dust aggregates as n(k) , ne, and nd(I,Z), i respectively. The rate equations for n(k) , ne, and nd(I,Z) are i given by n(k) i = ζ(k)n(k) g - u(k) i n(k) i Z dIXZ σdi(I,Z)nd(I,Z) ELECTRIC CHARGING OF DUST IN PROTOPLANETARY DISKS 3 ne =Xk - Xk nd(I,Z) =Xk and - β(k)n(k) i ne - Xl hβ ′(k,l)n(k) i n(k,l) g - β ′(l,k)n(l) i n(l,k) g i , (1) ζ(k)n(k) g - ueneZ dIXZ β(k)n(k) i ne, σde(I,Z)nd(I,Z) (2) u(k) i n(k) i [σdi(I,Z - 1)nd(I,Z - 1) - σdi(I,Z)nd(I,Z)] +uene[σde(I,Z + 1)nd(I,Z + 1) - σde(I,Z)nd(I,Z)], (3) respectively. In equations (1) and (2), n(k) denote the number densities of neutrals X(k) . We assume that neutral particles are much more abundant than charged particles and regard n(k) as constant parameters. In addition, we impose the charge neutrality condition g and X(k,l) g and n(k,l) g and n(k,l) g g g ζ. It is evident that equation (7) is much simpler than the orig- inal equation (1). This is mainly attributed to the cancellation of the last term in the original equation, i.e., the term describ- ing the ion-neutral reactions. Of course, this also means that we have lost the chance to know the composition of ions in detail. This fact does not bother us since our primary interest is the charge state of dust aggregates, not the composition of ions. Now we try to solve the equations (2) -- (4) and (7) under the equilibrium condition (5) as analytically as possible. First, equations (7) and (2) are written as ζng - uihσdiindni - βnine = 0, ζng - uehσdeindne - βnine = 0, respectively. Here we have defined the averages of an arbi- trary function F = F(I,Z) over Z and over I as (11) (12) (13) (14) (15) (16) hFi(I) ≡ 1 F(I,Z)nd(I,Z) nd(I)XZ nd(Z)Z F(I,Z)nd(I,Z)dI 1 tively. We have also defined the total number density nd of with nd(I) ≡ PZ nd(I,Z) and nd(Z) ≡ R nd(I,Z)dI, respec- dust aggregates by nd =PZR nd(I,Z)dI. Equations (11) and (12) can be easily solved in terms of ni and ne as ni = ne = 2β s1 + uehσdeind 2β s1 + uihσdiind 4βζng uiuehσdiihσdein2 d 4βζng uiuehσdiihσdein2 d - 1! , - 1!. Next, equation (3) is reduced to uiniσdi(I,Z - 1)nd(I,Z - 1) - ueneσde(I,Z)nd(I,Z) = uiniσdi(I,Z)nd(I,Z) - ueneσde(I,Z + 1)nd(I,Z + 1). (17) This equation means that the "flux" of the distribution nd(I,Z) from the state Z - 1 to Z must be balanced to that from Z to Z + 1. If the flux did not vanish, there would exist a steady "flow" of the charge state distribution nd(I,Z) streaming from one direction to the other in Z-space. However, since nd(I,Z) must be vanish at Z → ±∞, such a steady flow must not ex- ist. Therefore, both of the left and right-hand sides of equa- tion (17) must be zero. Hence we have uiniσdi(I,Z)nd(I,Z) = ueneσde(I,Z + 1)nd(I,Z + 1). (18) This is just the condition of detailed balance between charge states Z and Z + 1. For the sake of later convenience, we here rewrite the charge neutrality condition (4) using the definitions (6), (13), and (14) as ni - ne +hZind = 0. (19) As we will see later, this is the final equation that determines the equilibrium solution. The next step is to solve the detailed balance equation (18). To carry out the calculation, we need to specify the forms of the effective collision cross sections, σdi and σde. For simplic- ity, we model a fractal aggregate as a spherical, porous body with radius a and projected cross section σ. Also, we ne- glect the electric polarization of aggregates (Draine & Sutin n(k) i Xk - ne +Z dIXZ Znd(I,Z) = 0. (4) F(Z) ≡ Equations (1) -- (4) form the closed set of basic equations for ionization-recombination reactions in gas-dust mixture. 2.2. The equilibrium solution The equilibrium solution is obtained by imposing the con- ditions n(k) i = ne = nd(I,Z) = 0 (5) for all k, I, and Z. Usually, the equilibrium solutions are calculated with a re- action scheme involving many ion species and ion-neutral re- actions (e.g., Umebayashi & Nakano 1980; Ilgner & Nelson 2006a). Because of the complexity of the ion-neutral reac- tions, it is generally impossible to solve this problem analyti- cally without any approximation or simplification. This problem, however, becomes analytically tractable if we dot not require to distinguish ion species. This is just achieved by introducing the total ion density n(k) i . (6) ni =Xk Taking the sum of equations (1) over all k, we obtain the rate equation for ni, ni = ζng - uiniZ dIXZ where ng =Pk n(k) g σdi(I,Z)nd(I,Z) - βnine, (7) is the total number density of the gas, and ui = β = ζ = 1 1 niXk niXk ngXk 1 n(k) i u(k) i , n(k) i β(k), n(k) g ζ(k), (8) (9) (10) are the average ion velocity, gas-phase recombination rate co- efficient, and ionization rate, respectively. Similarly, equa- tions (2) -- (4) can be also written down in terms of ni, ui, β, and 4 OKUZUMI 1987) since a fractal aggregate is likely to have low dielec- tricity. Draine & Sutin (1987) found that the strength of the electric polarization relative to the charge-charge (Coulomb) interaction is determined by a factor (ǫ - 1)/(ǫ + 2), where ǫ is the dielectric constant of the material. Given that a porous aggregate is well approximated by a dilute medium, the fac- tor (ǫ - 1)/(ǫ + 2) is inversely proportional to its mean internal density (Clausius-Mossotti relation; see, e.g., Jackson 1998). For example, the mean internal density of a fractal aggre- gate with D ≈ 2 is about inversely proportional to its radius a. Therefore, we expect that the polarization effect of such a low-density aggregate is negligible. Neglecting the polariza- tion effect, the effective cross sections are simply written as (Spitzer 1941) (20) (21) (22) (23) (24) σdi(I,Z) = σsi(Z)×  σde(I,Z) = σse(Z)×  Z < 0, (cid:18)1 - λZ a (cid:19) , exp(cid:18)- λZ a (cid:19) , Z ≥ 0, exp(cid:18) λZ a (cid:19) , (cid:18)1 + λZ a (cid:19) , Z ≥ 0, Z < 0, where si(e)(Z) is the probability that a colliding ion (electron) sticks to one of constituent monomers, and λ = e2/kBT . In this study, we assume that si(Z) and se(Z) are independent of the net charge Z carried by an aggregate, i.e., si(e)(Z) ≡ si(e). As shown in Appendix, the solution nd(I,Z) to equation (18) with equations (20) and (21) is well approximated by a Gaussian distribution where and nd(I,Z) = hZia ≡ (Z - hZia)2 2h∆Z2ia (cid:21) , 1 nd(I) exp(cid:20)- Znd(I,Z) ≡ - Γa p2πh∆Z2ia nd(I)XZ (Z - hZia)2nd(I,Z) = λ 1 + Γ 2 + Γ a λ h∆Z2ia ≡ 1 nd(I)XZ are the mean and dispersion of the charge distribution for fixed a, respectively. This solution is valid when the radius a is much larger than λ, as is for aggregates much larger than constituent monomers (see Appendix). The nondimensional parameter Γ ≡ - hZiλ/a = (- hZie2/a)/kBT measures the elec- trostatic attraction (repulsion) energy between a charged ag- gregate and an incident ion (electron) relative to the thermal kinetic energy. Since the charge distribution is parametrized by Γ only, ni and ne can be written as a function of a single parameter Γ. Using equations (22) -- (24), hσdii and hσdei are evaluated as hσdei = σse exp(cid:20)- Γ + λ(1 + Γ) respectively. Here we have used in equation (26) that a/λ ≫ 1. Substituting equations (25) and (26) into equations (15) and (16), we have 2a(2 + Γ)(cid:21) ≈ σse exp(- Γ), hσdii = σsi(1 + Γ), (25) (26) (30) (31) (32) (33) (34) where ne = ζng seueσnd √1 + 2g(Γ)- 1 exp(- Γ)g(Γ) , g(Γ) = 2βζng siuiseue(σnd)2 exp Γ 1 + Γ . (28) (29) Finally, the neutrality condition (19) with equations (23), (27), and (28) leads to the equation for Γ, Γg(Γ) 1 1 + Γ - (cid:20) siui seue exp Γ + 1 Θ √1 + 2g(Γ)- 1(cid:21) = 0, where we have defined a nondimensional parameter ζngλ siuiσ an2 d = ζnge2 siuiσ an2 dkBT . Θ ≡ We have consequently arrived at the conclusion that all the conditions for ionization equilibrium, equations (11), (12), and (18), are reduced to a single equation (30) for a single parameter Γ. To summarize the above analysis, we have considered the charge state of a dust-gas mixture in the presence of ionization sources. We have found that the self-consistent equilibrium solutions are written as analytical functions of a single param- eter Γ (eqs. [22] -- [24], [27], and [28]), and have obtained the equation for this master parameter (eq. [30]). This equation can be easily solved numerically, and thus gives semianalyti- cal solutions to ni, ne, and nd(I,Z). The resultant equations can be further simplified when the gas-phase recombination rate coefficient β is so small that the factor g defined in equation (29) is much less than unity. In this limit, equations (27), (28), and (30) are simply rewritten as ni = ne = ζng siuiσnd ζng seueσnd 1 1 + Γ , exp Γ, and 1 1 + Γ - (cid:20) siui seue exp Γ + Γ Θ(cid:21) = 0, respectively. As seen in §3.1.3, this approximation is valid in typical protoplanetary disks unless the dust is significantly depleted (e.g., by vertical sedimentation). 2.3. Limiting cases Equation (30), or (34), provides clear insight into the charge state of a gas-dust mixture. The first and second terms in the bracket in this equation originate from ne and - hZind in the charge neutrality condition (4), respectively. This means that the parameter Θ determines which of free electrons and dust aggregates are the dominant carriers of negative charge. In the following, we categorize the charge state from limiting cases of equation (34). 2.3.1. Θ → ∞: ion-electron plasma limit In the limit Θ → ∞, the contribution from charged dust becomes negligibly small. Hence, equation (34) is well ap- proximated by 1 1 + Γ ≈ siui seue exp Γ. (35) ni = ζng siuiσnd √1 + 2g(Γ)- 1 (1 + Γ)g(Γ) , (27) This is a well-known charge-equilibrium condition for a dust particle immersed in an ordinary ion-electron plasma (Spitzer ELECTRIC CHARGING OF DUST IN PROTOPLANETARY DISKS 5 1941; Shukla & Mamun 2002). We denote the solution Γ to equation (35) by Γmax, since this is the maximum value of Γ obtained from equation (34). Typically, Γmax takes a value of order unity. 2.3.2. Θ → 0: ion-dust plasma limit In the limit Θ → 0, on the other hand, the contribution from free electrons becomes negligible. This means that the dom- inant carriers of negative charges are dust particles, not free electrons. We shall refer to this limit as the ion-dust plasma limit. In this limit, equation (34) can be approximated by 1 1 + Γ ≈ The solution is easily obtained as √1 + 4Θ2 - 1 Θ Γ . Γ ≈ 2 ≈ Θ. Since Γ is now negligibly small, the effective cross sections hσdi(e)i are approximately equal to σsi(e). Equations (32) and (33) lead to the ratio of ne to ni, ne ni ≈ siui seue ≈ si ser me mi . (38) where we have used that ui(e) =p8kBT /πmi(e). We note that this value of ne/ni is larger than that of Umebayashi (1983) by a factor of 1/√se. This is because we have assumed the value of se as independent of Z while Umebayashi (1983) consid- ered se(Z > 0) = 1. As far as the author knows, there is no ex- perimental data that validates either of the assumptions. How- ever, this difference is practically unimportant unless the ratio pse(Z < 0)/se(Z ≥ 0) is much less than unity. The value of Θ at the transition from one plasma regime to the other can be roughly estimated by equating the asymptotic solutions for both limits, i.e., Θ ≈ Γmax. We will use this estimation in the next section. 3. APPLICATION: ELECTRIC BARRIER AGAINST DUST GROWTH Electrostatic interaction between charged aggregates affect their collisional cross section. Let us consider two dust aggre- gates with mass m j, radius a j and charge Z je, where j(= 1,2) labels the aggregates. The kinetic energy for relative motion of two aggregates 1 and 2 is written as m(∆u)2, Ekin = (39) where m = m1m2/(m1 + m2) and ∆u are the reduced mass and the relative speed for the aggregate pair. The electrostatic en- ergy between the aggregates just before contact is 1 2 Eel = Z1Z2e2 a1 + a2 . (40) Neglecting the polarization effect as done in §2, the effective collision cross section σdd for the aggregates is expressed as (Spitzer 1941) σdd =  π(a1 + a2)2(cid:18)1 - Eel 0, Ekin(cid:19) , Ekin > Eel, Ekin ≤ Eel, (41) Therefore, the condition for the aggregates to collide with each other is (42) In this section, we examine whether this condition is satisfied in an early stage of dust evolution in a protoplanetary disk. Ekin > Eel. and (36) (37) 3.1. Model setup 3.1.1. Protoplanetary disk model We assume that the gas surface density Σg and the temper- ature T of the disk obey power laws Σg(r) = 1.7× 103 fΣ(cid:16) r 1 AU(cid:17) T (r) = 280(cid:16) r 1 AU(cid:17) - 1/2 - 3/2 g/cm2, (43) K, (44) where r is the distance from the central star and fΣ is a nondi- mensional scaling parameter. The model with fΣ = 1 is known as the minimum-mass solar nebula (MMSN) model (Hayashi 1981). We adopt fΣ = 1 unless otherwise noted. The tem- perature profile (44) is valid only for optically thin regions. We nevertheless employ this profile throughout the disk since the main result is insensitive to the detail of the temperature profile (see eq. [65] below). The hydrostatic equilibrium in the vertical direction gives the gas density distribution ρg(r,z) = Σg √2πH(r) exp(cid:20)- z2 2H(r)2(cid:21) , (45) where z is the height from the disk midplane and H(r) = cs(r)/ΩK(r) is the gas scale height. The isothermal sound ve- locity cs and the Kepler rotational frequency ΩK(r) are given by cs(r) =pkBT (r)/µmH and ΩK(r) =pGM∗/r3, where µ is the mean molecular weight, mH is the hydrogen mass, G is the gravitational constant, and M∗ is the mass of the central star. We adopt µ = 2.34 and M∗ = 1M⊙ in the following calcula- tion. The total number density ng of gas particles is given by ng = ρg/µmH. We assume that dust material is well mixed in the disk and that the dust density ρd is simply related to the gas density ρg by ρd(r,z) = fdgρg(r,z), (46) where fdg is the dust-to-gas ratio in the disk. The solar sys- tem abundance of condensates including water ice estimated by Pollack et al. (1994) leads to the dust-to-gas ratio fdg = 0.014 as well as the monomer bulk density ρ0 = 1.4g/cm3 (Tanaka et al. 2005). We do not consider the sublimation of water ice in inner disk regions for simplicity. 3.1.2. Ionization rate we study, consider Galactic this (Umebayashi & Nakano (Igea & Glassgold In rays rays 1999), (Umebayashi & Nakano 2009) as Thus, we decompose the ionization rate ζ as 1981), and cosmic stellar X- radionuclides the ionizing sources. ζ ≈ ζCR + ζXR + ζRA (47) where ζCR, ζXR, and ζRA denote the rate of ionization by cos- mic rays, X-rays, and radionuclides, respectively. We do not consider thermal ionization. This is negligible for T ≪ 103K, or for r ≫ 0.1AU (Umebayashi 1983). Charged particles are created primarily by ionization of H2 and He. The ioniza- tion rate for He is related to that for H2 by ζ(He) = 0.84ζ(H2) (Umebayashi & Nakano 1990, 2009), so it is sufficient to know ζ(H2) only. The total ionization rate ζ is given by ζ = + ζ(He)xHe, where xH2 = nH2/ng and xHe = nHe/ng are ζ(H2)xH2 the fractional abundances of H2 and He. We calculate xH2 and 6 OKUZUMI ζ(H2) CR (r,z)≈ 2 (exp(cid:18)- χ + exp(cid:18)- χ g(r,z) xHe from the solar system abundance by Anders & Grevesse (1989). The cosmic-ray ionization rate ζ(H2) CR for H2 is given by a fitting formula (Umebayashi & Nakano 2009) + g(r,z) + g(r,z) ζ(H2) CR,0 - 4/3 χCR (cid:19)(cid:20)1 +(cid:16)χ χCR (cid:17)3/4(cid:21) g(r,z) + g(r,z) χCR (cid:17)3/4(cid:21) χCR (cid:19)(cid:20)1 +(cid:16) χ - 4/3) ,(48) CR,0 ≈ 1.0× 10- 17/s is the cosmic-ray ionization rate where ζ(H2) for H2 in the interstellar space, χCR ≈ 96 g/cm2 is the attenua- z ρg(r,z′)dz′ tion length of the ionization rate, and χ + and χ g(r,z) are the vertical gas column densities measured from the upper and lower infinities, re- spectively. For the radionuclide ionization rate, we assume RA ≈ 7 × 10- 19/s, which corresponds to the ionization ζ(H2) rate by 26Al with an abundance ratio 26Al/27Al = 5 × 10- 5 g(r,z) =R ∞ g(r,z) = Σg(r) - χ (Umebayashi & Nakano 2009). The stellar X-ray ionization rate has been calculated by Igea & Glassgold (1999) using the Monte Carlo radiative transfer code including Compton scattering. A useful fitting formula is given by Turner & Sano (2008), ζXR(r,z)≈ ζXR,0(cid:16) r 1AU(cid:17) ×(cid:26)exp(cid:18)- χ LXR - 2(cid:16) 2× 1030erg/s(cid:17) χXR (cid:19) + exp(cid:18)- χ + g(r,z) χXR (cid:19)(cid:27) ,(49) where LXR is the X-ray luminosity, and ζXR,0 = 2.6× 10- 15/s and χXR = 8.0g/cm2 are the fitting parameters. This fitting formula approximately reproduces the kBTXR = 5keV result of Igea & Glassgold (1999) within the column density range χg & 1g/cm2, where scattered hard (& 5keV) X-rays are re- sponsible for the ionization. We use equation (49) in the following calculation since the typical value of χg is within the above range. We take LXR = 2× 1030erg/s in accordance with the median characteristic X-ray luminosity observed by Chandra for young solar-mass stars in the Orion Neb- ula Cluster (Wolk et al. 2005). Although the characteristic X-ray temperature kBTXR ≈ 2.4keV observed by Wolk et al. (2005) is lower than the assumed value of kBTXR ≈ 5keV, the choice of the temperature does not significantly affect the resulting ionization rate (Igea & Glassgold 1999). We do not consider temporary increase in the X-ray luminosity due to stellar flaring (Wolk et al. 2005), since our analytical method assumes stationary ionization processes. As pointed out by Ilgner & Nelson (2006c), the flaring could quantita- tively change the ionization state of the disk. We will examine the effect of the time-dependent flaring on dust growth in the future work. Figure 2 shows the total ionization rate ζ as well as its three components (ζCR, ζXR, and ζRA) as a function of r and z. X- ray ionization is dominant at outer radii (r & 4AU), while radionuclide ionization dominates at inner radii (r . 1AU). Cosmic-ray ionization is important in outer (r & 2AU) and low-altitude (z . H) regions. 3.1.3. Dust growth model Based on the results of recent laboratory and computer sim- ulations mentioned in §1, we model the dust growth in proto- planetary disks as follows. We start with monodisperse, non- aggregated dust grains (monomers) with radius a0. The mass m0 and number density nd0 of monomers are then written as m0 = (4π/3)ρ0a3 0 and nd0 = ρd/m0. The dust is assumed to grow into an ensemble of quasi-monodisperse, fractal aggre- gates with typical monomer number N and fractal dimension D ∼ 2. Under this assumption, we may regard N as the la- bel of dust evolutionary stages. Theoretically, this type of growth is the best modeled by the so-called ballistic cluster- cluster aggregation (BCCA; e.g., Meakin 1991). For this rea- son, we shall refer to the quasi-monodisperse fractal growth as the "BCCA growth."1 Note that the number density nd of aggregates is inversely proportional to N, since nd0 = Nnd is conserved. The fractal growth continues until colliding ag- gregates become energetic enough to compress each other. According to the microscopic model of Dominik & Tielens (1997), the critical kinetic energy for the onset of collisional compression is given by Ekin ∼ Eroll, where Eroll = 6π2γ ξcrit a0 2 ≈ 5.9× 10- 10(cid:16) γ 100erg/cm2(cid:17)(cid:16) ξcrit 2 A(cid:17)(cid:16) a0 0.1µm(cid:17)erg (50) is the energy needed to roll a monomer on another monomer in contact by 90◦, γ is the surface adhesion energy for the two monomers, and ξcrit is the critical tangential displace- ment for starting the rolling. For icy monomers, γ is es- timated as γ ≈ 100g/cm2 (Israelachvili 1992) but a realis- tic value of ξcrit is unknown. For a conservative estimation, we assume the minimum displacement ξcrit = 2 A anticipated by the theory (Dominik & Tielens 1997), which makes our aggregates the most easily compressed. The assumed value of Eroll is not much different from the experimental value for rocky (SiO2) monomers, Eroll ≈ 1.3× 10- 9(a0/0.1µm)erg (Heim et al. 1999), so the duration of the fractal growth stage is insensitive to our choice of dust material. We restrict our calculation to an early growth stage where the relative kinetic energy Ekin does not exceed the critical rolling energy Eroll. The radius a of a fractal aggregate is approximately given by a ≈ a0N1/D. (51) A classical, compact aggregate has D = 3, while a BCCA cluster has D ≈ 1.9 (Meakin 1991). We adopt D = 2 in the following calculation. The projected cross section σ of an aggregate is simply set to σ ≈ πa2 ≈ σ0N2/D ≈ σ0N, where σ0 = πa2 0 is the geometrical cross section of a monomer. This assumption is consistent with the fact that σ ∝ N for D . 2 (Meakin & Donn 1988; Meakin et al. 1989; Minato et al. 2006). Note that the quantity σnd is independent of N, i.e., conserved for the BCCA growth. Assuming a quasi-monodisperse size distribution, the ki- netic energy (39) is written as 1 4 Ekin ≈ m(∆u)2, (52) 1 Exactly speaking, the BCCA (i.e., collision between identical aggre- gates; see Meakin 1991) can only occur when the relative velocity of ag- gregates is induced by Brownian motion. When the relative velocity is in- duced by differential sedimentation, any identical aggregates cannot collide with each other, and therefore a BCCA cluster in its original sense cannot be created. On the other hand, a laboratory experiment by Blum et al. (1998) shows that the outcome of sedimentation-driven coagulation is an ensemble of quasi-monodisperse, fractal aggregates with D ≈ 1.7, as is for Brownian- motion-driven growth (e.g., Kempf et al. 1999). This may be explained by the fact that the dominant growth mode in differential sedimentation is the col- lision between similar aggregates (Tanaka et al. 2005). For this reason, this study treats the growth by differential sedimentation as the "BCCA growth." - - - - ELECTRIC CHARGING OF DUST IN PROTOPLANETARY DISKS 7 a z=H total X-ray  s   Ζ 10-17 10-18 y a r s m i c o c radionuclide 10-16 b r =5AU total  s   Ζ 10-17 X-ray 10-18 1 10 0.0 0.5 r AU cosmic ray radionuclide 1.5 2.0 1.0 zH FIG. 2. -- Total ionization rate ζ (thick gray curves) at different disk radii (a) and and altitudes (b). Here H denotes the scale height of the disk. The solid, dashed, and dotted black curves represent the contribution from Galactic cosmic rays, stellar X-rays, and radionuclides, respectively. where we have used m ≈ m/2. In a protoplanetary disk, rel- ative motion of aggregates is induced by Brownian motion, sedimentation toward the midplane of the disk, and turbu- lence. We therefore write the relative velocity ∆u as ∆u ≈p(∆uBrown)2 + (∆used)2 + (∆uturb)2, (53) where ∆uBrown, ∆used, and ∆uturb are the relative speed in- duced by the Brownian motion, differential sedimentation, and turbulence, respectively. The mean relative speed of the Brownian motion is given by ∆uBrown =r 8kBT π m ≈r 16kBT πm . (54) is In fact, the relative speed of the Brownian motion fluc- tuates according to the Maxwell distribution, and aggre- gates have a chance to get a relative kinetic energy E much larger than the thermal energy ∼ kBT with a probability ∝ E1/2 exp(- E/kBT ). However, as we see later, the effect of the thermal velocity fluctuation is insignificant, since the electro- static energy can go up to 105kBT . The relative speed induced by the differential sedimentation ∆used = Ω2 (55) where tstop is the stopping time of an aggregate. For aggre- gates smaller than the mean free path ℓg of gas particles, tstop is given by Epstein's law Kz∆tstop, tstop = 3 4ρgug m σ , (56) where ug =p8kBT /πµmH is the mean thermal speed of gas particles. The mean free path in our disk model is calculated to be ℓg ∼ 1(r/1 AU)11/4 cm, which is much larger than a typ- ical size of aggregate which we are interested in. For this reason, we always use Epstein's law (56) in the following cal- culation. Also, we replace ∆tstop with its maximum value tstop, which leads to the most conservative evaluation of dust charg- ing effect. The relative velocity induced by turbulence is given by ∆uturb ≈ usmall tsmall ∆tstop, (57) where usmall and tsmall are the characteristic velocity and turnover time of the smallest turbulent eddies, respectively (Weidenschilling 1984; Ormel & Cuzzi 2007). This ex- pression is valid for aggregates with stopping times tstop much smaller than tsmall. The velocity and turnover time of the largest eddies, ularge and tlarge, are related to usmall and tsmall by ularge = Re1/4usmall and tlarge ≈ Re1/2tsmall, where Re = νturb/νmol is the Reynolds number. The molecular is written as νmol = 0.5ug/ngσmol, where viscosity νmol σmol = 2 × 10- 15cm2 is the molecular collision cross section (Chapman & Cowling 1970). We express the turbulence vis- - 1 cosity νturb = u2 K , where αturb is the α-parameter. The turnover time of the largest eddies is - 1 taken to be tlarge ≈ Ω K , and thus their velocity is given by ularge ≈ √αturbcs. We consider αturb = 0, 10- 4,10- 3, and 10- 2 in this study. Again, we replace ∆tstop in equation (57) with tstop. If both of colliding aggregates have the mean charge hZi, the electrostatic energy (40) is written as largetlarge as νturb = αturbc2 s Ω Eel ≈ hZi2e2 2a = Γ2a 2λ2 . (58) where Γ is the master parameter defined in equation (23). Equation (58) overestimates a true repulsion energy when one (or both) of the aggregates has a positive charge Z > 0, or a negative charge - Z > 0 smaller than the average value - hZi. To account for the dispersion of the charge distribution, we introduce the "three-sigma" electrostatic energy Eel,3σ ≡ hZi(hZi + 3h∆Z2i1/2)e2 2a , (59) ability that a collision involves the electrostatic energy larger where the dispersion h∆Z2i1/2 is calculated from equation (24). Eel,3σ represents the electrostatic energy between two aggregates with charges hZi and hZi + 3h∆Z2i1/2. The prob- than Eel,3σ is as small as "three sigma" (∼ 10- 3). Note that Eel,3σ is always smaller than Eel since hZi is always nega- tive. Also, Eel,3σ becomes negative if - hZi is smaller than 3h∆Z2i1/2. This actually happens when the aggregate size is sufficiently small (see fig. 3a below). We compute the charge state at each evolutionary stage from equation (34) with a ≈ a(N), σ ≈ σ(N). The sticking coefficients are estimated by phonon theory to be si ≈ 1 and se = 0.1 . . . 1 (Umebayashi 1983). We adopt si = 1 and se = 0.3 in this study. The results obtained in this section are insensi- tive to the choice of se as long as 0.1 . se . 1. We do not use the original equation (30) because the gas-phase ionization is negligible in the present case. The gas-phase recombination rate is typically β ∼ 10- 12 ...- 7 cm3/s. Using this value, we can 8 OKUZUMI estimate g(Γ) as βζng g(Γ)∼ uiseue(σnd)2 ∼ 10- 11 ...- 6(cid:16) fdg 10- 2(cid:17) - 2(cid:16) r 5AU(cid:17)3(cid:16) T 130K(cid:17) - 1/2(cid:16) ζ 10- 17/s(cid:17), (60) independently of N. Therefore, equation (34) is valid unless dust is depleted and fdg decreases by many orders of magni- tude. ,H+ 2,H+ 3,He+ To confirm that our semianalytical calculation does work well, we have also performed fully numerical calculations including multi-component ions. In the numerical calcu- lations, a simple reaction model by Umebayashi & Nakano (1980) is adopted. This reaction model involves five light ,C+), heavy molecular ions (m+), metal ions (H+ ions (M+), free electrons, and charged dust aggregates. Heavy molecular ions and metal ions are represented by HCO+ and Mg+, respectively. We adopt the same values of the rate co- efficients β(k), β g /ng as those used by Sano et al. (2000). As seen below, the dom- inant ions are metal ions, which is essentially due to the fast charge transfer from heavy molecules to metal atoms. With this fact, we set the average ion mass mi to be the mass of Mg+ (mi = 24mH) in semianalytical calculations. The numeri- cal solutions are obtained from equations (1) -- (5). We remark that the full numerical calculation is far more time-consuming than the semianalytical one. ′(k,l) and the neutral gas abundances n(k,l) 3.2. Results 3.2.1. The fiducial case: αturb = 0, a0 = 0.1µm Here we show the result for the case αturb = 0 (i.e., laminar disk) and a0 = 0.1µm as a fiducial example. The results for different values of a0 and αturb are examined in §3.2.2 and §3.2.3, respectively. Figure 3 shows the evolution of dust charge state and gas ionization state at r = 5 AU, z = H. Here each evolutionary stage is labeled by the number of constituent monomers in an aggregate, N. The evolution of the mean hZi and dis- persion h∆Z2i of the dust charge distribution as well as the master parameter Γ is shown in figure 3a. We find that Γ in- creases with N and reaches to the maximum value Γmax = 2.8 at N ≈ 107. This means that the gas-dust mixture is an ion- dust plasma (Θ ≪ Γmax) at the initial stage, and evolves into an ion-electron plasma (Θ ≫ Γmax) as the dust grows. This is expected from the analysis in the last section: in the BCCA d) is proportional to N1/2, and thus in- growth, Θ ∝ 1/(σan2 creases with the growth. We have confirmed that the above value of Γmax is consistent with the solution to equation (35). It is useful to introduce the critical monomer number N ≡ Nmax at which the transition from the ion-dust plasma regime to the ion-electron plasma regime occurs. As explained in §2.3, this value can be estimated by setting Θ ≈ Γmax, or σan2 d ≈ ζngλ uiΓmax . (61) Substituting nd ≈ nd0/N, σ ≈ σ0N, and a ≈ a0N1/2 into this equation, we obtain the critical size for the transition Nmax ≈(cid:18) σ0a0uin2 d0Γmax ζnge2 (cid:19) 1/2 Σ(cid:16) fdg 0.014(cid:17)4(cid:16) r 5 AU(cid:17) ≈ 106 f 2 - 6(cid:16) ×(cid:16) a0 0.1µm(cid:17) 1.4g/cm3(cid:17) or equivalently, amax ≈ a0N1/2 max ρ0 ≈ 10- 2 fΣ(cid:16) fdg 0.014(cid:17)2(cid:16) r 5 AU(cid:17) - 2(cid:16) ×(cid:16) a0 0.1µm(cid:17) 1.4g/cm3(cid:17) ρ0 - 6(cid:16) T 130 K(cid:17)2(cid:16) - 4 ζ 10- 17/s(cid:17) - 2 , (62) - 3(cid:16) T 130 K(cid:17)(cid:16) - 2 cm, ζ 10- 17/s(cid:17) - 1/2 (63) where we have explicitly expressed the dependence on fΣ. The mean (negative) charge - hZi is proportional to N in the ion-dust regime (N ≪ Nmax), and is proportional to N1/2 in the ion-electron regime (N ≫ Nmax). This is easily under- stood if one recalls that - hZi ∝ Γa ∝ ΓN1/2. For N ≪ Nmax, Γ ≈ Θ is proportional to N1/2, so - hZi ∝ N. For N ≫ Nmax, Γ approaches a constant, and thus - hZi ∝ N1/2. On the other hand, the dispersion h∆Z2i1/2 is found to be nearly propor- tional to N1/4, which is because h∆Z2i1/2 ∼ (a/λ)1/2 and a ∝ N1/2. It is important to notice here that the relative width h∆Z2i1/2/hZi of the charge distribution becomes sharper and sharper as the dust grows. The transition of the plasma state is better illustrated by fig- ure 3b. This figure shows the number densities of ions and electrons, ni and ne, as well as the net dust charge density hZind. For N ≪ Nmax, dust is the dominant carrier of nega- tive charges, as expected for the ion-dust plasma state. The free electron density ne is smaller than that of ions by a factor of (1/se)(me/mi)1/2 ∼ 10- 2 (see eq. [38]). As the aggregates grow and N reaches the critical number Nmax, the dust charge density - hZind begins to decrease and ne begins to increase. Finally, at N ≈ 107, free electrons become the dominant nega- tive charge carrier, and the ion-electron plasma state (ni ≈ ne) is established. Interestingly, the abundances of charged species, ni and ne, are nearly constant for both limits of ion-dust and ion-electron plasma regimes. This result is in contrast to that of previous studies based on the classical, compact (D = 3) growth model (e.g., Sano et al. 2000; Wardle 2007) in which ni and ne in- crease as the dust grows. This difference is attributed to the fact that the net projected area σnd of dust aggregates is kept nearly constant for D . 2, while it decreases for D = 3. Us- ing the constancy σnd = σ0nd0, equations (32) and (33) can be rewritten as ni = ni0/(1 + Γ) and ne = ni0(seue/ui) exp Γ, where ni0 = ζng uiσ0nd0 - 1(cid:16) r ≈ 10- 14(cid:16) fdg 5AU(cid:17)3(cid:16) 0.014(cid:17) - 3(cid:16) ×(cid:16) a0 0.1µm(cid:17) 1.4g/cm3(cid:17) ρ0 ζ 10- 17/s(cid:17) - 1 ng (64) is the abundance of ions for N ≪ Nmax. For both plasma lim- its, the factors 1 + Γ and exp Γ are approximately constant, so both ni and ne approach constant values. Physically speaking, the constancy of σnd for D . 2 means that all the monomers in a fractal aggregate with D . 2 are exposed to outer space, and are thus capable to capture free electrons and ions. Figure 3 also shows the result of full numerical calculations using the simplified ion-reaction scheme. We find an excel- lent agreement between the semianalytical and full numerical ELECTRIC CHARGING OF DUST IN PROTOPLANETARY DISKS 9 6 4 2 0 -2 G g o l , Z g o l -5 -4 a à ae à ae 0 à ae 2 log acm -3 -2 -1 0 ae -Z ae ae ae à ae à ae à DZ2 12  à à à G ae à à ae semianalytical full numerical ae à 4 6 8 10 log N g n  n g o l -14 -15 -16 -17 log acm -3 -2 -1 0 -5 -4 b ì ì ì ì ae ae ae ae ì ae ì ae ì ae à ni »nM+ ì ì ì ì à à ae à à - ne à ò à ò à ò à ò à ò ò ò ò à ae  Z  n d ae ò ò nm+ ò ae semianalytical ae à ì ò full numerical 0 2 4 6 8 10 log N FIG. 3. -- (a) The evolution of the dust charge distribution for the "BCCA" (i.e., quasi-monodisperse, D ≈ 2) dust growth. The solid, dashed, and dotted curves represent the average hZi, dispersion h∆Z2i1/2, and master parameter Γ of the charge distribution calculated using the semianalytical method (eqs.[23], [24], and [34]). The filled circles and squares denote hZi and h∆Z2i1/2 obtained from full-numerical calculations using the Umebayashi & Nakano model (see §3.1.3). (b) The evolution of the gas ionization state and the dust charge state for the BCCA growth. ni and ne are the number densities of ions and electrons in the gas phase, and hZind is the net dust charge density. nM+ and nm+ are the number densities of metal ions and molecular ions. log acm -2 -4 -1 r =5AU, z=H, Αturb=0, a0=0.1Μm -3 Eroll 0 øøøø -5 6 4 T B k  E g o l 2 0 0 Ekin Eel 2 ‰ Eel,3 Σ 4 6 log N 8 10 FIG. 4. -- The relative kinetic energy Ekin (eq. [52]; solid black curve) versus the electrostatic energy Eel (eq. [(58)]; solid gray curve) for the BCCA growth as a function of the monomer number N. The disk position is (r, z) = (5AU, H), and the turbulence parameter and the monomer size are set to αturb = 0 (i.e., laminar) and a0 = 0.1µm. The collision cross section for an aggregate pair with Eel and Ekin vanishes when Eel > Ekin (see eq. [42]), meaning that the BCCA growth "freezes out" at the size indicated by the cross (×) symbol. The dashed gray curve shows the "three-sigma" electrostatic en- ergy Eel,3σ (eq. [59]), representing the effect of charge fluctuation. The star (⋆) symbol indicates the size at which Ekin reaches the critical rolling-friction energy Eroll (eq. [50]; dotted black curve). Above this critical energy, colli- sional compression of aggregates becomes effective. calculations. It is clear that our semianalytical approach is not only efficient but also accurate. The most abundant ions are metal ions for all stages of dust evolution. Molecular ions, the second most abundant ones, are an order of magnitude fewer than metal ions. Now we examine the growth condition. Figure 4 shows the kinetic energy Ekin and electrostatic energy Eel for collid- ing aggregates at each evolutionary stages. For N . 105 (a . 30µm), the thermal (Brownian) motion dominates the rela- tive velocity of the aggregates, so Ekin is kept constant ≈ kBT . Fore N & 105, the vertical sedimentation dominates the aggre- gate motion and Ekin increases with N. On the other hand, Eel always grows with N, in proportional to N3/2 for N . Nmax and to N1/2 for N & Nmax. This is explained from the fact that hZi ∝ N for the ion-dust plasma regime (N . Nmax) and hZi ∝ a ∝ N1/2 for the ion-electron plasma regime (N & Nmax). As a result, the growth condition (42) breaks down at N ≈ 103.5 (a ≈ 6µm). This means that the collision between aggregates with average charge hZi becomes impos- sible at this stage. Note that the repulsion energy Eel reaches 10 times the thermal energy ∼ kBT as early as N ≈ 104, and finally goes up to 105kBT at the onset of collisional compres- sion. It is evident that the thermal fluctuation of the kinetic energy cannot help the aggregates to grow beyond N ≫ 104. One may expect that the fluctuation of aggregate charge could help the growth. To see this effect, we overplot in figure 4 the "three-sigma" electrostatic energy Eel,3σ defined in equation (59). We see that Eel,3σ quickly converges to Eel and finally exceeds Ekin at N ≈ 104. This is expected from figure 3a: the relative width hZi/h∆Z2i1/2 of the the charge distribution becomes narrower and narrower as the dust grows. Therefore, the result that the aggregates cannot grow beyond N ≫ 104 is preserved even if the charge fluctua- tion is taken into account.2 All the above facts suggest that the BCCA growth of dust aggregates at this disk position "freezes out" at size N ∼ 104(a ∼ 10µm). It is interesting to compare this critical size for the freeze- out with that for the first compression, i.e., the size at which the growth mode changes from the BCCA to the growth in- volving collisional compression. We overplot the critical 2 It is also found that Eel,3σ is negative for smaller sizes, N . 103.7. This means that colliding aggregates can possess opposite charges with a proba- bility larger than "three-sigma". This is because the average negative charge - hZi in this stage is smaller than 3h∆Z2i1/2, as seen in figure 3a. However, the attraction energy - Eel,3σ is insignificant: at most half of the kinetic energy Ekin. 10 OKUZUMI rolling energy Eroll (eq. [50]) in figure 4. Comparing this crit- ical energy with Ekin, we find that the collisional compression begins at size N ≈ 1010(a ≈ 1cm), i.e., many orders of mag- nitude smaller than the above critical freeze-out size. This illustrates how fast the charging of aggregates begins to affect their collisional growth. Figure 5 compares the energy ratio Eel/Ekin for different disk positions. We find that the barrier against the BCCA growth appears irrespectively of the disk radius r. By con- trast, the growth barrier vanishes at z & 3H because the sed- imentation velocity at the high altitudes is large enough for aggregates to overcome the barrier. To summarize, the freeze-out of the BCCA growth is very likely to occur in this fiducial model, except at high altitudes over the midplane. As seen in figure 4, the energy ratio Eel/Ekin takes its max- imum at N ≈ Nmax. With this fact, we can roughly estimate the maximum value (Eel/Ekin)max of the energy ratio at z ≈ H as follows. We assume that the relative motion of aggregates by N ≈ Nmax is dominated by vertical sedimentation, as is for r = 5AU. Then, substituting ∆u ≈ ∆used into equation (52) and using equations (58) and (62), we obtain - 2(cid:16) Ekin(cid:17)max ≈ 30 fΣ(cid:16) fdg (cid:16) Eel 0.014(cid:17) 10- 17/s(cid:17) - 1(cid:16) ×(cid:16) a0 0.1µm(cid:17) 1.4g/cm3(cid:17) ρ0 ζ - 1 (65) at z ≈ H. Notably, (Eel/Ekin)max is explicitly independent of both r and T . This means that the "height" of the growth barrier is insensitive to the temperature profile. Equation (65) does not hold beyond r ≈ 5 AU, since the maximum energy ratio appears in the Brownian motion regime. This equation is nevertheless useful because it allows us a rough estimation on how the growth barrier depend on the model parameters. For example, for fixed ζ, a0, and ρ0, the the growth barrier is more serious if the disk is more massive ( fΣ > 1) or more depleted of dust ( fdg < 10- 2). 3.2.2. Effect of monomer size The actual size of dust monomers in protoplanetary disks is unknown. Infrared observations of interstellar medium sug- gest the size distribution of interstellar grains ranges from ≈ 0.005µm to ≈ 0.25µm (MRN distribution; Mathis et al. 1977). If interstellar grains are not aggregates but monomers, the typical monomer size in protoplanetary disks will fall within the range 0.001µm . a . 1µm. Figure 6 compares the energy ratio Eel/Ekin for different monomer sizes a0 = 0.01, 0.1, and 1µm. We find that Nmax ∝ a- 6 (amax ∝ a- 2 0 ) and (Eel/Ekin)max ∝ a- 1 0 , as expected from equations (62) and (65). This delay (measured in the "degree of growth" N) is attributed to the larger cross section of BCCA clusters com- posed of smaller monomers. Such a larger cross section causes a quick depletion of free electrons in the gas phase, resulting in the delay of transition from the ion-dust plasma regime to the ion-electron regime. In addition, the larger cross section produces a stronger coupling to the gas, and in turn a slower increase of the kinetic energy. We find that the growth condition (42) breaks down much before the on- set of collisional compression for all a0 . 1µm. Therefore, the "freeze-out" of the BCCA growth is not prevented even if the monomer size in protoplanetary disks is assumed to the maximum value inferred by the MRN distribution. 0 3.2.3. Effect of turbulence Here we examine how strong turbulence is needed to re- move the growth barrier. Figure 7 shows the evolution of en- ergy ratio Eel/Ekin for three turbulent cases αturb = 10- 4,10- 3, and 10- 2. We see that turbulence of αturb . 10- 4 does not affect the energy ratio for any size N. This is not surpris- ing because both ∆uturb and ∆used scale with ∆tstop. The relative velocity ∆uturb induced by turbulence is estimated as ∆uturb ≈ ∆tstopRe1/4ularge/tlarge ∼ ∆tstopRe1/4α KH. At z ∼ H, the Reynolds number Re is of order ∼ αturbΣgσmol/mg, so ∆uturb is written as 1/2 turb Ω2 α KH 3/4 turb (66) ∆tstopΩ2 mg (cid:17)1/4 ∆uturb ∼(cid:16) Σgσmol On the other hand, the relative velocity by differential sedi- mentation is ∆used ∼ ∆tstopΩ2 ∆uturb ∆used ∼(cid:16) Σgσmol 10- 4(cid:17)3/4(cid:16) r 5 AU(cid:17) turb ∼(cid:16) αturb mg (cid:17)1/4 KH, so we find Therefore, the effect of turbulence on the aggregate collision is negligible for all N as long as αturb . 10- 4. For αturb & 10- 2, we find that the growth barrier is entirely removed. This suggests that relatively strong (αturb & 10- 2) turbulence is a key ingredient for early stages of dust coagu- lation. In §4.3, we discuss this topic in more detail. . (67) - 3/8 3/4 α 4. DISCUSSION 4.1. Validity of the charge equilibrium In this study, we have assumed that the charge reactions are much faster than the dust-dust collisions. Now we show that this assumption is actually valid for evolutionary stages which we are interested in. The typical time scale for the system to relax to an ionization-recombination equilibrium can be measured by the average time tcoll,d needed for an aggregate to collide with an ion, t- 1 coll,i ≈ uiσdini ≈ ζng nd , (68) where we have used that ζng ≈ uiσdinind since the gas- phase recombination is negligible in the presence of dust (see §3.1.3). On the other hand, the mean collision time tcoll,d be- tween aggregates is written as t- 1 coll,d ≈ σddnd∆u ∼ σnd∆u. Therefore, the ratio of these time scales is estimated as (69) (70) mg σΣg ζ ΩK . If the dust velocity is dominated by the Brownian motion, as σn2 d tcoll,d tcoll,i ∼ ζng mg(cid:17)2 cs dg(cid:16) m ∆u ∼ f - 2 is for small aggregates, ∆u ∼ cspmg/m and ∆u tcoll,d mg(cid:17)5/2 mg dg(cid:16) m tcoll,i(cid:12)(cid:12)(cid:12)(cid:12)Brown ∼ f - 2 - 2(cid:16) ∼ N3/2(cid:16) fdg 10- 2(cid:17) σΣg ζ ΩK ζ 10- 17/s(cid:17)(cid:16) r 2 AU(cid:17)3 , (71) where we have used a0 ∼ 0.1µm and ρ0 ∼ 1g/cm3. We find that dust coagulation can be safely neglected if r & 2AU ELECTRIC CHARGING OF DUST IN PROTOPLANETARY DISKS 11 log acm log acm -5 -4 -3 -2 -1 0 1 -5 -4 -3 -2 -1 0 1 a z=H, Αturb=0, a0=0.1Μm r =15AU 5AU ‰ ‰ 1AU ø ø ‰ 2 1 0 i n k E l  e E g o i n k E l  e E g o l l -1 -2 0 ø 2 4 6 log N 8 10 12 2 1 0 -1 -2 0 b r =5AU, Αturb=0, a0=0.1Μm ‰‰ z=1H 2H ø ø 3H 2 4 6 log N 8 10 12 FIG. 5. -- The ratios Eel/Ekin between the electrostatic and kinetic energies (solid curves) at various disk radii r (a) and altitudes z (b) as functions of N. The α-parameter and the monomer radius are set to αturb = 0 and a0 = 0.1µm. The cross (×) symbols indicate the size at which the growth condition (42) breaks down for aggregates with mean charge hZi. The dashed curves show the ratio Eel,3σ/Ekin for the "three-sigma" energy, which represents the effect of charge dispersion. The star (⋆) symbols indicate the sizes at which Ekin reaches the critical rolling-friction energy Eroll. 3 2 1 ‰ 0 i n k E l  e E g o l a0=1Μm r =5AU, z=H, Αturb=0 0.01Μm 0.1Μm ø ‰ ø ‰ -1 -2 0 2 4 ø 6 8 log N 10 12 14 FIG. 6. -- The energy ratios Eel/Ekin (solid curves) at r = 5AU and z = H for different monomer sizes a0. The cross (×) symbols indicate the "freeze- out" sizes at which the growth condition (42) breaks down. The dashed curves show Eel,3σ/Ekin. The star (⋆) symbols indicate the sizes at which Ekin reaches the critical rolling-friction energy Eroll. or N ≫ 1. This is true even if the motion of dust aggre- gates is dominated by vertical sedimentation, since ∆used ∼ cs(m/Σgσ), and thus m mg ζ ΩK dg tcoll,d tcoll,i(cid:12)(cid:12)(cid:12)(cid:12)sed ∼ f - 2 ∼ 103N(cid:16) fdg 10- 2(cid:17) - 2(cid:16) Thus, it is concluded that the growth of dust aggregates can be neglected if r & 2AU or N ≫ 1. . (72) 10- 17/s(cid:17)(cid:16) r 2 AU(cid:17)3/2 ζ i n k E l  e E g o l log acm -5 -4 -3 -2 -1 0 1 2 1 0 -1 -2 0 r =5AU, z=H, a0=0.1Μm Αturbd10-4 10-3 10-2 ø ø ø ‰‰ 2 4 6 log N 8 10 12 FIG. 7. -- The energy ratios Eel/Ekin (solid curves) at r = 5AU and z = H for different turbulence strengths αturb. The cross (×) symbols indicate the "freeze-out" sizes at which the growth condition (42) breaks down. The dashed curves show Eel,3σ/Ekin. The star (⋆) symbols indicate the sizes at which Ekin reaches the critical rolling-friction energy Eroll. 4.2. Internal electrostatic force In the last section, we have implicitly assumed that charged aggregates can stick to each other as long as the collision con- dition (42) is satisfied. One might wonder if the collided ag- gregates are pulled off from each other by the electrostatic repulsion. In fact, this repulsion is much weaker than the at- traction (due to van der Waals force) between two monomers in contact. The electrostatic repulsion force Fel,int acting be- 12 OKUZUMI tween two collided aggregates is estimated as Γ2e2 Fel,int ∼ (Ze)2 a2 ∼ λ2 . 10- 8(cid:16) T (73) where we have used that Γ < Γmax ∼ 3. Note that max- imum value of Fel,int is independent of the aggregate size a. On the other hand, the critical force needed to sep- arate two monomers in contact is (Johnson et al. 1971; Dominik & Tielens 1997) 130 K(cid:17)2 dyn, Fcrit = 3πγ a0 2 ∼ 10- 3(cid:16) γ 102g/cm2(cid:17)(cid:16) a0 0.1µm(cid:17) dyn, (74) where γ is the surface adhesive energy mentioned in the last section. Thus, the electrostatic force inside an aggregate is negligibly weak compared to the contact force between two constituent monomers. 4.3. Dust growth in strong turbulence As seen in §3.2.3, the electrostatic barrier against the fractal growth will be removed if considerably strong (αturb & 10- 2) turbulence is present. We here discuss whether such turbu- lence is likely to occur in protoplanetary disks, and what would happen after the dust overcome the electrostatic growth barrier. The most robust mechanism for disk turbulence is MRI 3. MRI-driven turbulence will achieve αturb ∼ 10- 2 in its satu- rated state (e.g. Sano et al. 1998). Therefore, fractal aggre- gates will be able to overcome the electric barrier in MRI- active regions. Sano et al. (2000) calculated the active region using the MMSN model and found that MRI will be active only at outer (r & 20AU) disk radii or high (z & 2H) altitudes if the dust size is 0.1µm. The size of the active region does not vary even if the dust grows since, as seen in §3.2.1, the ionization fraction is kept nearly constant as long as the dust undergoes the fractal growth. Therefore, the region in which the fractal aggregates can overcome the electrostatic barrier is limited to outer disk radii and high altitudes. A more serious problem is that such strong turbulence causes another kind of growth barrier, i.e., catastrophic frag- mentation of colliding aggregates. In turbulent regions, large aggregates with tstop ∼ 1/ΩK have the maximum collisional velocity of order ularge ∼ √αturbcs, which amounts to more than 100m/s for αturb & 10- 2. On the other hand, as shown by recent N-body simulations (Wada et al. 2008), catastrophic fragmentation will take place for relative velocity ∆u & 30m/s. Therefore, it is very likely that strong turbulence de- stroys the aggregates and consequently prevents further dust growth. This idea is supported by a recent statistical study (Brauer et al. 2008). Thus, the combination of electric repulsion and collisional fragmentation might strictly limit the dust growth and sub- sequent planetesimal formation in protoplanetary disks. It is important to think of a possibility that dust evolution will continue in some way even if the turbulence is weak and the quasi-monodisperse fractal growth does freeze out. This is the topic of the next subsection. 4.4. A possible scenario to overcome the electric growth barrier Eel,ft Ekin,ft ≈ Eel,ff Ekin,ff(cid:16) ∆uff ∆utf(cid:17)2 . (77) In §3, we have ignored the size distribution of aggre- gates. In fact, there may exist some aggregates considerably larger than average-sized ones. In the following, we consider whether such large aggregates can continue to grow even if the growth of average-sized ones has frozen out. Let us consider a small population of irregularly large ag- gregates (referred to as "test aggregates") growing with a large population of standard (D ∼ 2) fractal aggregates ("field ag- gregates"). Under this assumption, the kinetic energy of rela- tive motion between test and field aggregates is written as 1 1 2 1 2 mtmf mt + mf 2(cid:16) ∆utf ∆uff(cid:17)2 Ekin,tf = mf(∆utf)2 = (∆utf)2 ≈ Ekin,ff, (75) where the subscripts 't' and 'f' respectively represent the test and field aggregates, and we have used the assumption mt ≫ mf. Ekin,ff and ∆uff are the kinetic energy of relative motion and the relative velocity between two field aggregates, and are thus equivalent to Ekin and ∆u in §3. On the other hand, the electrostatic energy between test and field aggregates is Eel,tf = atafΓ2e2 at + af ≈ Γ2af λ2 = 1 2 Eel,ff, (76) where Eel,ff is equivalent to Eel in §3, and we have used that at ≫ af. Thus, the energy ratio Eel,tf/Ekin,tf is written as Now we assume that the growth of field aggregates has frozen out due to the electric barrier, i.e., Ekin,ff = Eel,ff. At this stage, the condition for the collision between test and field aggregates, Ekin,tf > Eel,tf, reduces to a simple inequality ∆utf > ∆uff. (78) We readily notice that Brownian motion does not sat- isfy this condition since ∆utf ≈ p8kBT /πmf and ∆uff ≈ p16kBT /πmf. The remained possibilities are the differen- tial sedimentation and turbulent-driven motion. In both cases, the relative velocity ∆u is proportional to ∆tstop, or ∆(m/σ). It is very important to notice here that the condition (78) is not safely satisfied as long as the test aggregate is as fluffy as field aggregates, i.e., mt/σt is comparable to mf/σf. Hence, the condition (78) will be safely satisfied only if the test aggre- gate is more compact and has larger m/σ than the field aggre- gates. Moreover, collision with a smaller aggregate generally tends to increase mt/σt, allowing the test aggregate the next collision. Therefore, if there exists an aggregate that is large and compact, it will be able to continue growing by sweeping up smaller "frozen" aggregates. The above consideration suggests that the freeze-out of the quasi-monodisperse fractal growth may not mean the termi- nation of dust evolution. Rather, it may be the beginning of bimodal growth in which only a small fraction of aggregates can grow larger and larger while the rest remain frozen. We plan to examine this possibility in more detail in the future studies. In any case, we expect that the effect of dust charg- ing should qualitatively modify the current scenario of dust growth in protoplanetary disks. 3 It is unknown whether any mechanism other than MRI can drive and sustain turbulence with αturb & 10- 2 in the early stage of dust evolution. For example, convective instability may operate in this stage (Lin & Papaloizou 1980), but it is unlikely to sustain such strong turbulence (Stone & Balbus 1996). 5. SUMMARY In this study, we have investigated the electric charging of dust aggregates and its effect on collisional dust growth in protoplanetary disks. We have found that the conditions for ELECTRIC CHARGING OF DUST IN PROTOPLANETARY DISKS 13 ionization-recombination equilibrium are reduced to a single equation (eq.[30]). Just by solving this equation numerically, the dust charge state and gas ionization state can be analyti- cally computed for an arbitrary ensemble of aggregates in a self-consistent way. It is also confirmed that our semianalyti- cal method reproduces the results of a previously used, more complicated numerical method (§3.2.1, fig.3). This formal- ism thus provides a fast charge-state solver that will allow a coupled simulation of MRI-driven turbulence and dust coag- ulation. As an application, we have explored the effect of electro- static charging on an early stage of dust coagulation in proto- planetary disks. We considered the quasi-monodisperse frac- tal growth with the fractal dimension D ∼ 2 as suggested by previous laboratory experiments and N-body simulations (Blum 2004; Dominik et al. 2007). Our findings are summa- rized as follows: 1. For a wide range of model parameters, the effective cross section for the mutual collision of aggregates is quickly sup- pressed as the fractal growth proceeds and finally vanishes at a certain aggregate size (§§3.2.1, 3.2.2). This is due to the strong electrostatic repulsion between aggregates charg- ing negatively on average, and happens much before the col- lisional compression of aggregates becomes effective. Both the charge fluctuation and the thermal velocity fluctuation do not help the aggregates to overcome the growth barrier. With- out strong turbulence, the quasi-monodisperse fractal growth is very likely to "freeze out" on its way to the subsequent growth stage. 2. Strong (αturb & 10- 2) turbulence will help the aggregates to overcome the above growth barrier (§3.2.3). However, such turbulence is likely to occur only in MRI-active regions, i.e., at outer disk radii or high altitudes (§4.3). Furthermore, it will cause another serious problem -- the catastrophic disrup- tion of collided aggregates -- in later stages. These facts sug- gest that the combination of electric repulsion and collisional disruption may strictly limit the collisional growth of dust ag- gregates in protoplanetary disks. 3. The freeze-out of the fractal growth might be followed by bimodal growth in which only a small fraction of large ag- gregates can continue growing while a large fraction of small fractal aggregates remains frozen (§4.4). This could qualita- tively change the current scenario of planetesimal formation in protoplanetary disks (Dominik et al. 2007). We will exam- ine this possibility in more detail in forthcoming papers. Finally, we point out that the fractal (D . 2) dust growth tends to keep the ionization degree of the disk small due to the open nature of aggregates (§3.2.1, fig. 3b). This means that the magnetorotationally unstable region hardly expands until the collisional compression of the aggregates begins to work. This conclusion is in contrast to that of previous studies (e.g., Sano et al. 2000; Wardle 2007) which claimed that the ionization degree increase as the aggregates grow. However, they assumed compact dust growth, which clearly contradicts recent laboratory experiments and N-body simulations. Thus, the magnetorotational stability of protoplanetary disks must be reexamined taking into account that the fractal nature of dust aggregates. The author thanks M. Sakagami, S. Inutsuka, and H. Tanaka for careful reading of the manuscript and for valuable com- ments. APPENDIX DERIVATION OF THE CHARGE DISTRIBUTION Since the velocity of free electrons is much greater than that of ions, dust aggregates charge up negatively on average. Thus, let us assume Z < 0 and use the expressions of σdi and σde valid for Z < 0. Then, equation (18) is written as where τ = a/λ. We make the following assumption: where h∆Z2i is the variance of the charge state distribution, and ε ≪ 1. We also assume that nd(I,Z) varies with a typical scale ∼ h∆Z2i1/2. Under these assumptions, nd(I,Z + 1) can be written as nd(I,Z + 1) = nd(I,Z) + ∂nd ∂Z - 1 = h∆Z2i (I,Z) + O(ε2). - 1/2 ·h∆Z2i1/2/τ ∼ ε2. Substituting them into equation Also, exp[(Z + 1)/τ] is written as exp(Z/τ) + O(ε2) since τ (A1), we obtain a first-order differential equation for nd(I,Z), (A3) where ∂nd ∂Z W (I,Z) ≡ 1 - (I,Z) +W(I,Z)nd(I,Z) ≈ 0, siuini(1 - Z/τ) seuene exp(Z/τ) . (A4) (A5) Equation (A4) is accurate to terms of the first order in ε. Let us denote the solution of W (I,Z) = 0 by Z0 and write Z = Z0 + δZ. Also, we make an additional approximation that δZ ≪ τ. Expanding W (I,Z) in powers of δZ/τ and using W (I,Z0) = 0, we have 2 - Z0/τ 1 - Z0/τ W (I,Z) ≈ δZ τ (A6) , siuini(cid:18)1 - Z τ(cid:19)nd(I,Z) = seuene exp(cid:18) Z + 1 τ (cid:19)nd(I,Z + 1), 1 h∆Z2i1/2 ∼ h∆Z2i1/2 τ ∼ ε, (A1) (A2) 14 OKUZUMI which is accurate to the first order in δZ/τ. Hence, equation (A4) is approximated by where ∂nd ∂Z (I,Z) + δZ h∆Z2i nd(I,Z) ≈ 0, h∆Z2i ≡ 1 - Z0/τ 2 - Z0/τ τ . (A7) (A8) It is an easy task to show that the solution to (A7) is a Gaussian distribution (22) with average hZi = Z0 and variance h∆Z2i. Rewriting Z0 and τ using Z0 = - Γτ and τ = a/λ, we obtain equations (23) and (24). We note that equation (A8) has been also obtained by Draine & Sutin (1987), but they did not show the derivation of this equation in their paper. - 1/2 ∼ h∆Z2i1/2/τ, which is equivalent to h∆Z2i ∼ Now we justify the assumption (A2). The first part of the assumption h∆Z2i τ, is always satisfied since τ /2 < h∆Z2i < τ (see eq.[A8]). The second part h∆Z2i1/2/τ ∼ ε ≪ 1 is also satisfied in typical protoplanetary disks if dust aggregates are not as small as constituent monomers (∼ 0.1µm), because 0.1µm(cid:19) It is noted that the approximation δZ ≪ τ used to obtain equation (A6) is rewritten as δZ ≪ τ 1/2h∆Zi1/2. Hence, this approximation is good as long as the region δZ . h∆Zi1/2 is considered. - 1/2 ∼(cid:18) a ∼ τ (A9) h∆Z2i1/2 τ . - 1/2(cid:18) T 130K(cid:19) - 1/2 REFERENCES 285 Anders, E., & Grevesse, N. 1989, Geochim. Cosmochim. Acta, 53, 197 Balbus, S. A., & Hawley, J. F. 1991, ApJ, 376, 214 Blum, J.,Wurm, G., Poppe, T., & Heim, L.-O. 1998, Earth Moon Planets, 80, Blum, J., et al. 2000, Phys. Rev. Lett., 85, 2426 Blum, J. 2004, in ASP Conf. Ser. 309, Astrophysics of Dust, ed. A. N. Witt, G.C. Clayton, & B. T. Draine (San Francisco: ASP), 369 Brauer, F., Dullemond, C. P., Henning, Th. 2008, A&A, 480, 859 Chapman, S., & Cowling, T. G. 1970, The Mathematical Theory of Nonuniform Gases (London: Cambridge Univ. Press) Dominik, C. P., Blum, J., Cuzzi, J. N., & Wurm, G. 2007, in Protostars and Planets V, ed. B. Reipurth, D. Jewitt, & K. Keil (Tucson: Univ. Arizona Press), 783 Dominik, C., & Tielens, A. G. G. M. 1997, ApJ, 480, 647 Draine, B. T., & Sutin, B. 1987, ApJ, 320, 803 Gammie, C. F. 1996, ApJ, 457, 355 Glassgold, A. E., Najita, J., & Igea, J. 1997 ApJ, 480, 344 Hayashi, C. 1981, Prog. Theor. Phys. Suppl., 70, 35 Heim, L.-O., Blum, J., Preuss, M., & Butt, H.-J. 1999, Phys. Rev. Lett., 83, Igea, J., & Glassgold, A. E. 1999, ApJ, 518, 848 Ilgner, M., & Nelson, R. P. 2006a, A&A, 445, 205 Ilgner, M., & Nelson, R. P. 2006b, A&A, 445, 223 Ilgner, M., & Nelson, R. P. 2006c, A&A, 445, 731 Inutsuka, S., & Sano, T. 2005, ApJ, 628, L155 Israelachvili, J. 1992, Intermolecular and Surface Forces (2nd ed.; London: Academic Press) Ivlev, A. V., Morfill, G. E., & Konopka, U. 2002, Phys. Rev. Lett., 89, 195502 Jackson, J. D. 1998, Classical Electrodynamics (3rd ed.; New York: Wiley) Johnson K. L., Kendall, K., & Roberts, A. D. 1971, Proc. R. Soc. London A, 3328 Kempf, S., Pfalzner, S., & Henning, T. K. 1999, Icarus, 141, 388 Konopka, U., Mokler F., Ivlev, A. V., Kretschmer, M., Morfill, G. E., et al. 2005, New Journal of Physics, 7, 227 Lin, D. N. C., & Papaloizou, J. C. B. 1980, MNRAS, 191, 37 324, 301 452, 701 Mathis, J.S., Rumpl, W., & Nordsieck, K. H. 1977, ApJ, 217, 425 Meakin, P., & Donn, B. 1988, ApJ, 329, L39 Meakin, P., Donn, B., & Mulholland, G. W. 1989, Langmuir, 5, 510 Meakin, P. 1991, Rev. Geophys., 29, 317 Minato, T., Köhler, M., Kimura, H., Mann, I., and Yamamoto, T. 2006, A&A, Nakagawa, Y., Nakazawa, K., & Hayashi, C. 1981, Icarus, 45, 517 Ormel, C. W., & Cuzzi, J. N. 2007, A&A, 466, 413 Pollack, J. B., Hollenbach, D., Beckwith, S., Simonelli, D. P., Roush, T., & Fong, W. 1994, ApJ, 421, 615 Sano, T., Inutsuka, S., & Miyama, S. M. 1998, ApJ, 506, L57 Sano, T., Miyama, S. M., Umebayashi, T. & Nakano, T. 2000, ApJ, 543, 486 Shukla, P. K., & Mamun, A. A. 2002, Introduction to Dusty Plasma Physics (Bristol: IoP) Spitzer, L. 1941, ApJ, 93, 369. Stone, J. M., & Balbus, S. A. 1996, ApJ, 464, 364 Suyama, T., Wada, K., & Tanaka, H. 2008, ApJ, 684, 1310 Tanaka, H., Himeno Y., & Ida, S. 2005, ApJ, 625, 414 Turner, N. J., & Sano, T. 2008, ApJ, 679, L131 Umebayashi, T. 1983, Prog. Theor. Phys., 69, 480 Umebayashi, T., & Nakano, T. 1980, PASJ, 32, 405 Umebayashi, T., & Nakano, T. 1981, PASJ, 33, 617 Umebayashi, T.. & Nakano, T. 1990, MNRAS, 243, 103 Umebayashi, T.. & Nakano, T. 2009, ApJ, 690, 69 Wada, K., Tanaka, H., Suyama, T., Kimura, H., & Yamamoto, T. 2008, ApJ, 677, 1296 Wardle, M. 2007, Ap&SS, 311, 35 Weidenschilling, S. J. 1980, Icarus, 44, 172 Weidenschilling, S. J. 1984, Icarus, 60, 553 Wolk, S. J., Harnden, F. R., Flaccomio, E., Micela, G., Favata, F., Shang, H., & Feigelson, E. D. 2005, ApJ, 160, 423 Wurm, G., & Blum, J. 1998, Icarus, 132, 125
1705.06304
1
1705
2017-05-17T18:29:59
The compositional diversity of non-Vesta basaltic asteroids
[ "astro-ph.EP" ]
We present near-infrared (0.78-2.45 {\mu}m) reflectance spectra for nine middle and outer main belt (a > 2.5 AU) basaltic asteroids. Three of these objects are spectrally distinct from all classifications in the Bus-DeMeo system and could represent spectral end members in the existing taxonomy or be representatives of a new spectral type. The remainder of the sample are classified as V- or R- type. All of these asteroids are dynamically detached from the Vesta collisional family, but are too small to be intact differentiated parent bodies, implying that they originated from differentiated planetesimals which have since been destroyed or ejected from the solar system. The 1- and 2-{\mu}m band centers of all objects, determined using the Modified Gaussian Model (MGM), were compared to those of 47 Vestoids and fifteen HED meteorites of known composition. The HEDs enabled us to determine formulas relating Band 1 and Band 2 centers to pyroxene ferrosilite (Fs) compositions. Using these formulas we present the most comprehensive compositional analysis to date of middle and outer belt basaltic asteroids. We also conduct a careful error analysis of the MGM-derived band centers for implementation in future analyses. The six outer belt V- and R-type asteroids show more dispersion in parameter space than the Vestoids, reflecting greater compositional diversity than Vesta and its associated bodies. The objects analyzed have Fs numbers which are, on average, between five and ten molar percent lower than those of the Vestoids; however, identification and compositional analysis of additional outer belt basaltic asteroids would help to confirm or refute this result. Given the gradient in oxidation state which existed within the solar nebula, these results tentatively suggest that these objects formed at either a different time or location than 4 Vesta.
astro-ph.EP
astro-ph
The compositional diversity of non-Vesta basaltic asteroids Thomas B. Leitha,b, Nicholas A. Moskovitza, Rhiannon G. Maynec, Francesca E. DeMeod, Driss Takire, Brian J. Burta,d, Richard P. Binzeld, Dimitra Pefkoud aLowell Observatory, Flagstaff, AZ, 86001, USA bHarvard-Smithsonian Center for Astrophysics, Cambridge, MA, 02138, USA cMonnig Meteorite Collection, Texas Christian University, Fort Worth, TX, 76129, USA dMassachusetts Institute of Technology, Cambridge, MA, 02139, USA eAstrogeology Science Center, United States Geological Survey, Flagstaff, AZ, 86001, USA Abstract We present near-infrared (0.78-2.45 µm) reflectance spectra for nine middle and outer main belt (a > 2.5 AU) basaltic asteroids. Three of these objects are spectrally distinct from all classifications in the Bus-DeMeo system and could represent spectral end members in the existing taxonomy or be representatives of a new spectral type. The remainder of the sample are classified as V- or R- type. All of these asteroids are dynamically detached from the Vesta collisional family, but are too small to be intact differentiated parent bodies, implying that they originated from differentiated planetesimals which have since been destroyed or ejected from the solar system. The 1- and 2-µm band centers of all objects, determined using the Modified Gaussian Model (MGM), were compared to those of 47 Vestoids and fifteen HED meteorites of known composition. The HEDs enabled us to determine formulas relating Band 1 and Band 2 centers to pyroxene ferrosilite (Fs) compositions. Using these formulas we present the most comprehensive compositional analysis to date of middle and outer belt basaltic asteroids. We also conduct a careful error analysis of the MGM-derived band centers for implementation in future analyses. The six outer belt V- and R-type asteroids show more dispersion in parameter space than the Vestoids, reflecting greater compositional diversity than Vesta and its associated bodies. The objects analyzed have Fs numbers which are, on average, between five and Preprint submitted to Icarus September 19, 2018 ten molar percent lower than those of the Vestoids; however, identification and compositional analysis of additional outer belt basaltic asteroids would help to confirm or refute this result. Given the gradient in oxidation state which existed within the solar nebula, these results tentatively suggest that these objects formed at either a different time or location than 4 Vesta. 1. Introduction Basaltic asteroids provide us with a rare window into the earliest periods in the formation of the solar system. These objects, which are a relative rar- ity in the main belt, are thought to form only in the mantle of planetesimals which have undergone the process of metal-silicate differentiation (Gaffey et al., 1993, 2002). Models suggest that these planetesimals were tens to many hun- dreds of km in size (Hevey and Sanders, 2006; Moskovitz and Gaidos, 2011; Moskovitz and Walker, 2011; Neumann et al., 2012). Compositional analysis of iron meteorites suggests that at least 60 such objects were once present in the inner solar system (Chabot and Haack, 2006; Burbine et al., 2002). Nearly all basaltic asteroids in the main belt are members of the Vesta family, and most are thought to be ejecta from the impact that created Rheasilvia, an impact basin which covers Vesta's southern hemisphere and is one of the largest such features in the solar system (Marazari et al., 1996; McSween et al., 2013). The HED (Howardite, Eucrite, Diogenite) meteorites are thought to also originate from within the Vestoid family, making them a rare example of meteorites which can be tied to a specific differentiated parent body. Eucrites are thought to be pieces of Vesta's crust, Diogenites are thought to originate within Vesta's upper mantle, and Howardites are a brecciated mix of both (Burbine, et al., 2001). A few basaltic asteroids which are not associated with Vesta's dynamical family have been discovered in recent years. These objects lie in the middle and outer main belt across Jupiter's 3:1 mean motion resonance from Vesta (see Figure 1). Roig et al. (2008) show that there is a less than 1 percent chance of a 5-km V-type asteroid being ejected from Vesta and crossing the 3:1 2 Kirkwood gap. Several outer belt basaltic asteroids, such as (1459) Magnya, are also simply too large to have plausibly been ejected from Vesta (Ieva et al., 2015). Furthermore, the Dawn spacecraft has provided sufficiently high spatial resolution of Vesta's surface spectral features to conclude that at least two V-types beyond the 3:1 Kirkwood gap could not have originated from the Rheasilvia impact basin (Ieva et al., 2015). Thus, they are most likely fragments of differentiated parent bodies which have since been destroyed or ejected from the solar system (Lazzaro et al., 2000; Hammergren et al., 2006; Roig et al., 2008; Huaman et al., 2014). To further understand these objects, along with how and where they may have formed, we have analyzed near-infrared reflectance spectra of nine basaltic asteroids located in the main belt between a = 2.5 AU and a = 3.3 AU. We then compared the spectra of these asteroids to those of 47 Vestoid asteroids and fifteen HED meteorites of known composition. Eight additional achondrite, olivine-rich meteorites from the Acapulcoite and Lodranite suites were analyzed to serve as references for objects that display 1- and 2-µm absorption bands but do not have HED-like compositions. While previous studies have examined several of these asteroids before, there has yet to be a comprehensive analysis of a majority of the known middle and outer belt basaltic asteroids. Moskovitz et al. (2008c), Hardersen et al. (2004), and Roig et al. (2008) performed focused studies of single objects, while Ieva et al. (2015) analyzed only two outer belt V-types -- (1459) Magnya and (21238) Panarea, both of which are discussed in this work. Our study examines a larger sample of basaltic bodies past the 3:1 Kirkwood gap in order to broadly assess these objects' composition and make inferences about where this population could have originated. The location of these objects' orbits within the main belt is shown in Figure 1. 3 Figure 1: Dynamical map of the main belt, showing Vesta (black triangle), the Vestoids (red dots), our V-type targets (black circles), and non-V-types discussed in this work (gray triangles). Other asteroids are shown as blue dots. The dashed lines show the location of major mean motion and secular resonances. It is dynamically unlikely that any of our targets are derived from Vesta. Throughout this work we use several terms. V-type asteroids, the category to which most of our target objects belong, are a spectral type within the Bus- DeMeo asteroid taxonomic system (Bus and Binzel, 2002b; DeMeo et al., 2009), distinguished by deep absorption bands near 1 and 2 µm which are characteris- tic of a pyroxene composition. One of our target asteroids, (349) Dembowska, is not a V-type, but rather is the only known R-type according to the Bus-DeMeo taxonomic system (Bus and Binzel, 2002b; DeMeo et al. 2009). The R-type spectrum also shows significant absorption features near 1 and 2 µm, but has a slightly flatter spectrum and a more complex 1-µm absorption feature, likely due to the presence of olivine (Gaffey et al., 1993). The Vestoid family is a dynami- 4 2.22.42.62.83.03.2Semi-major Axis (AU)05101520Inclination (deg)Vesta1459223082123810504140521349747210537143903:15:27:32:1ν6 cal and compositional grouping of V-type asteroids associated with 4 Vesta, the second-most massive body in the main belt and the only asteroid believed to be a largely intact differentiated planetesimal. Other types of planetary differen- tiation are known to occur, such as ice-silicate differentiation -- however, when used in this work, the term differentiation refers exclusively to metal-silicate differentiation. The process of metal-silicate differentiation was most likely driven by ra- dioactive decay of short-lived isotopes such as 26Al and 60Fe (Grimm and Mc- Sween, 1993; Goswami and Vanhala, 2000; Tachibana and Huss, 2003), which have half-lives of 0.73 and 1.5 Myr, respectively. This places constraints on both the time and location of these objects' formation, as they must have un- dergone differentiation within a few half-lives and could not have formed in the gas giant region, where they would have accreted too slowly to undergo melting (Moskovitz et al., 2008a). Bottke et al. (2006) propose that metal-silicate differ- entiation occurred in the inner solar system, interior to a = 2.0 AU. This would indicate that Vesta, the Vestoids, and the other main belt basaltic asteroids which are the focus of this study may have formed at a considerably smaller heliocentric distance than their current locations and were later scattered out- ward. In order to further explore the possibility of this scenario, we analyzed the pyroxene mineralogy of these asteroids. Pyroxenes have varying abundances of iron, magnesium, and calcium, the molar percentages of which are reflected in their respective Ferrosilite (Fs), Enstatite (En), and Wollastonite (Wo) numbers (e.g. Mayne et al., 2010). In this work, we focused on determining the Fs number of each of our target asteroids. In addition to scaling with the bulk iron concentration in which it formed, a given pyroxene's Fs number is strongly influenced by the oxidation state of its formation environment, since oxidizing conditions tend to produce more iron-rich minerals. Given that an oxidation gradient existed within the solar nebula, comparing two objects' Fs numbers can thus shed light on where they formed relative to one another (Rubin and Wasson, 1995). 5 To take advantage of this phenomenon, we derived formulas relating the cen- ters of pyroxene absorption bands from HED meteorite spectra taken from the RELAB (Reflectance Experiment Laboratory) database at Brown University. We then used them to calculate the Fs content of our target objects. In doing so we aim to provide observational constraints on the Bottke et al. (2006) model of scattering in the young solar system. The remainder of this paper is organized as follows: in Section 2 we discuss how and when our asteroid observations were obtained. In Section 3 we fit observational spectra of our target objects, Vestoids, and HED meteorites using the Modified Gaussian Model (Sunshine et al. 1990; Sunshine et al. 1993) and relate spectral properties to composition. In Section 4 we discuss the spectral classification of our objects, and interpret the results in light of the Bottke et al. (2006) model of scattering and the "Grand Tack" planetary migration model put forward by Walsh et al. (2012). 2. Observations and reduction 2.1. Asteroid observations Much of the data for this work were drawn from published, archival sources. For objects with little or no existing spectral data, we obtained new near-infrared (0.8-2.5 µm) spectra with two instruments: SpeX at NASA's Infrared Telescope Facility (IRTF) on Mauna Kea (Rayner et al., 2005) and the Folded-port In- frared Echellette (FIRE) spectrograph on the Magellan Baade 6.5-m at Las Campanas in Chile (Simcoe et al., 2008). The sources of visible (where applica- ble) and near-IR spectra for each of our target objects are shown in Table 1. In all cases except the asteroids (349) Dembowska and (1459) Magnya, our middle and outer belt targets were originally identified as candidate basaltic asteroids based on their photometric colors in the Sloan Digital Sky Survey Moving Ob- ject Catalog (SDSS MOC; Ivezic et al. 2001). The ugriz photometric band passes employed by the SDSS MOC enable coarse assignment of asteroid spec- tral types and are particularly well suited at distinguishing basaltic asteroids 6 with deep 1 µm absorption features (Rig and Gil-Hutton, 2006; Moskovitz et al. 2008b; Solontoi et al. 2012). The only two targets not selected for follow-up based on SDSS colors were (349) Dembowska, long known to have unusual spec- tral characteristics (Feierberg et al. 1980), and (1459) Magnya, discovered as an outer belt V-type as part of the larger S3OS2 survey (Lazzaro et al. 2000). Sources of observational data Object Visible data NIR data (349) Dembowska Bus et al., 2002 DeMeo et al. 2009 (1459) Magnya Lazzaro et al. 2000 Hardersen et al. 2004 (7472) Kumakiri No visible data IRTF, Sept 6, 2010 (10537) 1991 RY16 (14390) 1990 QP19 (21238) Panarea (1) Moskovitz et al. 2008b Moskovitz et al. 2008b No visible data IRTF, July 11, 2013 Roig et al. 2008 FIRE, Aug 24, 2010 (21238) Panarea (2) Roig et al. 2008 IRTF, Sept 8, 2015 (22308) 1990 UO4 (40521) 1999 RL95 (105041) 2000 KO41 No visible data FIRE, July 3, 2012 No visible data IRTF, Sept 14, 2015 Solontoi et al. 2012 FIRE, Oct 24, 2012 Table 1: Sources of all observational data for target asteroids. Spectra of these objects are shown in Figure 2. Acquisition of data from these two instruments followed analogous proce- dures. For both the slit mask was oriented along the parallactic angle at the start of each observation to minimize the effects of atmospheric dispersion. So- lar analogs were observed to correct for telluric absorption and to remove the solar spectrum from the measured reflectance. Objects and solar analogs were observed near the meridian and at similar airmass. Individual exposures for the asteroids were held between 120-180s to avoid saturation of telluric emission features and, for FIRE, to avoid saturated thermal emission from the instru- ment and telescope at wavelengths longer than about 2.2 µm. Exposures were obtained in standard ABBA nod sequences with the target offset by several arc seconds along the slit at the two nod positions. FIRE was operated in its high-throughput prism mode with a 0.8 × 50" slit. These settings produced single-order spectra at a resolution of approximately 400 from 0.8 to 2.45 µm. At the IRTF, SpeX was configured in its low resolution (R = 250) prism mode with a 0.8" slit for wavelength coverage from 0.8 to 2.5 µm. Reduction of the SpeX data followed DeMeo et al. (2009). Reduction 7 of the FIRE data employed an IDL package designed for the instrument and based on the Spextool pipeline (Cushing et al., 2004) using procedures detailed in Binzel et al. (2015). Spectra for all of our outer belt target objects are shown in Figure 2. In addition to these observations, we utilized spectral data on 4 Vesta from Gaffey (1997) and a large sample of Vestoid spectra previously compiled in Moskovitz et al. (2010). We calculated signal-to-noise for each band center independently by taking the inverse of the standard deviation of the residuals to our fits (see section 3.1) in the region of each band center. We defined this region for Band 1 as 0.85-1.05 µm, and for Band 2 as 1.85-2.05 µm. Observational circumstances for each of our targets are summarized in Table 2. 8 Figure 2: Normalized and vertically offset spectra and MGM fits to each of the middle and outer belt objects discussed in this work. Data points are plotted as solid circles, and the MGM fit as a solid black line. The numerical designation of each object is listed at right with its spectral type, where known. Objects with an ambiguous spectral classification -- i.e. not consistent with any known types -- are denoted with a "?". These three spectra are plotted with the closest available type to which they could be assigned (from top to bottom: V, O, and O). 9 1.01.52.02.53.0Wavelength (µm)0.00.51.01.52.0Normalized Reflectance7472 (?)14390 (?)10537 (?)349 (R)1459 (V)21238 (V)22308 (V)40521 (V)105041 (V) 2.2. Meteorite spectra We compared our asteroid data to reflectance spectra of HED meteorites taken using the near-IR bidirectional spectrometer at at the NASA/Keck Re- flectance Experiment Laboratory at Brown University (Pieters, 1983). In addi- tion, we analyzed reflectance spectra of Acapulcoite and Lodranite meteorites taken at the same facility. The HED meteorites provided spectra of Vestoid analogs with known, laboratory-determined compositions. The Acapulcoites and Lodranites, meanwhile, served as comparators with similar spectral fea- tures (i.e. 1- and 2-µm absorption bands). These specific groups were of par- ticular interest as they are known to be more olivine-rich than the Vestoids and HED meteorites (e.g. McCoy et al., 1996) and thus can provide insight on the limitations of our methodology for non-Vesta compositions. The process of compositional validation performed using these meteorite spectra is outlined in sections 3.4 and 3.5. For each sample we analyzed spectral data taken from observations of the smallest grain size available. Specific grain sizes for each meteorite measured are available in Table 3, along with Fs numbers drawn from either Mayne et al. (2011) or Burbine et al. (2009) as indicated. 10 Object (4) Vesta Feb 18-20 , 1981 UT Date r(AU ) Mag. texp (min) S/N (B1) S/N (B2) Vesta and Vestoids α(◦ ) - 8 16 32 24 87 40 12 32 32 28 48 32 24 8 28 32 28 32 40 48 32 48 28 64 33 54 93 27 40 120 64 40 48 48 40 32 24 24 8 11 40 14 56 30 24 48 523 560 637 534 296 531 260 371 325 300 152 504 279 354 302 437 354 344 456 380 317 295 594 356 385 117 327 861 190 146 359 114 263 323 316 123 304 473 182 - 320 197 97 168 77 152 89 241 223 301 442 168 209 226 198 154 42 146 197 203 72 281 358 423 38 312 225 129 182 72 114 226 186 256 140 55 96 212 31 40 124 51 40 82 81 50 63 240 80 - 389 167 32 14 35 23 25 143 50 107 (809) Lundia Aug 26 , 2008 (956) Elisa (1468) Zomba (1929) Kollaa (2045) Peking Jul 5 , 2008 Sep 30 , 2003 Feb 19 , 2001 Aug 26 , 2008 (2371) Dimitrov Aug 1 , 2009 (2442) Corbett Sep 15 , 2002 (2511) Patterson May 7 , 2004 (2566) Kirghizia May 8 , 2002 (2579) Spartacus Oct 10 , 2000 (2653) Principia Nov 26 , 2002 (2763) Jeans Jun 26 , 2004 (2795) Lepage Apr 9 , 2005 (2823) van der Laan Nov 22 , 2005 (2851) Harbin Aug 24 , 2001 (2912) Lapalma Feb 20 , 2001 (3155) Lee Jun 22 , 2001 (3344) Modena Sep 4 , 2005 (3657) Ermolova Aug 1 , 2009 (3703) Volkonskaya Jun 3 , 2006 (3782) Celle Nov 26 , 2002 (4038) Kristina Oct 28 , 2002 (4188) Kitezh (4215) Kamo (4796) Lewis (5481) Kiuchi Aug 14 , 2001 Nov 11 , 2002 Jan 9 , 2009 Sep 5 , 2005 (5498) Gustafsson Aug 26 , 2008 (7800) Zhongkeyuan Jan 9 , 2009 (9481) Menchu Aug 26 , 2008 (9553) Colas Jan 8 , 2009 (16416) 1987 SM3 Nov 23 , 2007 (26886) 1994 TJ2 Jul 5 , 2008 (27343) 2000 CT102 Aug 26 , 2008 (33881) 2000 JK66 Nov 23 , 2007 (36412) 2000 OP49 Nov 23 , 2007 (38070) 1999 GG2 Oct 5 , 2006 (50098) 2000 AG98 Sep 4 , 2005 (97276) 1999 XC143 Nov 23 , 2007 2.38 1.93 1.85 1.6 2.21 2.5 2.47 2.25 2.23 2.41 2.22 2.53 2.24 2.26 2.2 2.42 2.14 2.56 2.17 2.18 2.25 2.6 2.08 2.1 2.52 2.46 2.37 1.92 2.51 2.48 1.99 2.4 2.09 1.94 1.77 2.12 1.9 1.94 2.05 4 23.9 16.6 38.3 14.4 16.9 19.6 7.8 21.2 9.8 18.9 17.8 12.6 4.7 16.1 8.7 8.7 8.4 17.9 25.4 9.8 17.1 10.3 10.8 17.3 13.9 10.5 7.8 2.8 6.1 20.3 6 10.9 11.7 23.3 11.8 4.1 11.4 11.4 6.1 14.6 14.6 16.5 15.3 16.2 16.7 15.6 16.1 16 16.4 16.3 15.7 15.9 16.4 15.6 15.3 16.2 16.1 16.6 17.2 16.8 15.8 15.3 16.5 17 16.3 16.3 17.1 17.3 17.4 16.9 17.2 16.6 16.6 16.9 16.9 15.7 15.7 Outer Belt V- and R-Types (349) Dembowska (1) Jun 21 , 2001 2.956 17.9 (349) Dembowska (2) November 26 , 2001 2.71 (1459) Magnya Mar 23 , 2003 3.851 (21238) Panarea (1) Jul 20 , 2006 (21238) Panarea (2) Sept 8 , 2015 (22308) 1990 UO4 (40521) 1999 RL95 (105041) 2000 KO41 Jul 3 , 2012 Sept 14 , 2015 Oct 24 , 2012 2.8 1.84 2.55 1.41 2.89 1 17.3 5.5 21.6 15 3.1 7.3 10.5 10.5 16.3 17.1 17.5 17.8 17.6 18.6 Spectrally Anomalous Targets (See Section 4.1) (7472) Kumakiri Sep 6-7 , 2010 (10537) 1991 RY16 (14390) 1990 QP19 Jan 30 , 2008 Jul 11 , 2008 3.33 2.85 2.9 6 3.7 17 16.4 19.7 18 64 137 24 Table 2: Observational data for main belt basaltic asteroids. The columns in this table are: object designation, UT date of observation, heliocentric distance at the time of observation in AU (r), phase angle of the target in degrees (α), V-band magnitude of the target (V, retrieved from Moskovitz et al. 2010), the net exposure time in minutes, and S/N in the region of each band center. 11 3. MGM band fitting Figure 3: MGM fit to the spectrum of the outer belt V-type (1459) Magnya. The data are represented by gray dots, and the continuum by a red dashed line. The modeled bands are in blue, the fit in purple, and the residuals in orange. MGM allows for highly accurate fits to complex features like the overlapping bands at 0.9 and 1.2 µm and is effective at isolating the characteristics of individual bands. 3.1. Band analysis 1-µm and 2-µm band centers for all asteroid and meteorite spectra were determined using the Modified Gaussian Model (MGM) (Sunshine et al. 1990, Sunshine et al. 1993). MGM uses a flexible continuum and a series of modified Gaussians representing absorption bands to fit a spectrum. Parameters repre- senting the offset, slope, and curvature of the continuum, as well as the number, location, width, and depth of bands are supplied by the user, then adjusted by MGM within a user-defined range until it has achieved an optimal fit. MGM 12 is thus able to accurately fit spectra and determine individual band parameters even in the presence of overlapping bands, which is particularly important when analyzing pyroxene spectra featuring absorption bands centered near 1 µm and 1.2 µm (Mayne et al., 2010). Previous research using MGM to analyze basaltic asteroid and meteorite spectra has employed a physically-motivated fit in order to yield specific min- eralogy (Mayne et al., 2010; Ieva et al., 2015). However, our approach was a purely mathematically-motivated one, similar to that employed by Thomas & Binzel (2010) and Mayne et al. (2011). We aimed to determine a single set of input parameters which would yield the best possible fit when used with all of our HED meteorite and asteroid spectra while simultaneously minimiz- ing free parameters by using the fewest possible bands. While this approach would not be sufficient for extracting mineralogical information on any single object, applying it to spectra of HED meteorites with known mineralogy and basaltic asteroids yields relative mineralogical insight. These can then be used for comparison across populations of basaltic objects. To ensure that our MGM fits were well-calibrated to objects with known min- eralogy, we began by collecting spectra of meteorites with known bulk pyroxene Fs numbers from the online database curated by the NASA/Keck Reflectance Experiment Laboratory (RELAB) at Brown University (Pieters, 1983). These spectra were then truncated to the NIR region of 0.78-2.45 µm in order to match the minimum spectral range of our outer belt asteroid spectra. We then attempted MGM fits to our HED meteorite spectra using a variety of input parameters fitting 3, 4, and 5 bands with flat, sloped, and nonlinear contin- uums. The success of each set of parameters was judged by both the quality of the fit as measured by the overall RMS error output by MGM and by how closely the resulting band centers matched those found in Mayne et al. (2011). RELAB does not provide error bars on its spectra. However, we were able to estimate a wavelength-averaged signal-to-noise (S/N) of each spectrum as the mean reflectance value divided by the standard deviation of the RMS residuals of the fit, which we then mapped to band center errors using the methodology 13 outlined in section 3.2. We ultimately found that a set of parameters modeling four bands and a sloped continuum yielded the best fit while minimizing free parameters. How- ever, in the case of three Eucrite meteorites -- ALHA81001, BTN00300, and MAC02522 -- the output band centers varied too significantly from those mea- sured in Mayne et al. (2011), and so these objects were excluded from our sample. Our exact set of input parameters was further refined by using MGM to fit spectra of Vesta, numerous V-type asteroids, and our sample of basaltic outer belt asteroids. For several objects, MGM fit an unrealistically wide 1.2- µm band, which resulted in the 1-µm band falling well below where it could plausibly be located. As a result, the maximum width of the 1.2-µm band was restricted significantly more than that of other bands. Ultimately we were able to determine a single set of input parameters which could be utilized for MGM fits to all of our meteorite and asteroid spectra, with the lone exception of the outer belt asteroid (14390) 1990 QP19. This object, along with asteroids (7572) Kumakiri and (10537) 1991 RY16, may represent a new spectral type of asteroid and is discussed at greater length in section 4.1. Our final set of input parameters is shown in Table 4, and a sample MGM fit (to (1459) Magnya) is shown in figure 3. 14 Object RELAB Grain Size Measured Fs B1 Center B2 Center designation (µm) Number (µm) (µm) Diogenites EETA79002 GRO95555 Johnstown LAP91900 Tatahouine Howardites EET87503 QUE94200 Y791573 Eucrites Chervony Kut EET87520 GRA98098 Ibitira MET01081 MB-TXH-067-A MP-TXH-068-A MB-TXH-095-A MP-TXH-077-A MP-TXH-088-A 0-25 0-25 0-25 0-25 0-25 MB-TXH-068-AP MP-TXH-069-A MP-TXH-099-A 0-25 0-25 0-25 MT-HYM-035 MT-HYM-029 MT-HYM-034 MP-TXH-054-A MT-HYM-033 0-38 0-45 0-38 0-25 0-45 0-25 0-45 0-25 0-25 0-25 0-25 0-25 0-25 0-25 0-25 0-25 Moore County MP-TXH-086-A PCA91078 MT-HYM-031 Serra de Mage MP-TXH-092-A MT-CMP-001 TB-TJM-036 TB-TJM-039 TB-TJM-034 TB-TJM-037 MB-CMP-026 MB-TXH-037 MB-TXH-038 Acapulcoites Acapulco ALHA81187 ALHA81261 Lodranites Lodran GRA95209 MAC88177 Y74357 Y791491 aMayne et al. (2011) b Burbine et al. (2009) cMcCoy et al. (1996) d Floss (1991) eMcCoy et al. (1997) 22.0a 25.0a 23.5b 23.0a 23.0b 36.5a 30.5a - 41.9a 39.7a 51.2a 42.0a 44.6a 37.3a 40.2a 36.2a 8.8c 5.1c 7.5c 0.917 ± 0.003 0.921 ± 0.003 0.916 ± 0.002 0.919 ± 0.003 0.918 ± 0.002 1.889 ± 0.004 1.903 ± 0.004 1.881 ± 0.003 1.906 ± 0.004 1.893 ± 0.003 0.928 ± 0.003 0.921 ± 0.003 0.927 ± 0.004 1.954 ± 0.004 1.921 ± 0.003 1.963 ± 0.005 0.934 ± 0.003 0.946 ± 0.003 0.937 ± 0.003 0.938 ± 0.002 0.934 ± 0.003 0.935 ± 0.003 0.953 ± 0.005 0.927 ± 0.003 2.005 ± 0.004 2.016 ± 0.004 1.996 ± 0.004 1.990 ± 0.003 1.982 ± 0.004 1.982 ± 0.004 2.021 ± 0.005 1.963 ± 0.004 0.933 ± 0.005 0.907 ± 0.003 0.935 ± 0.005 1.809 ± 0.006 1.862 ± 0.004 1.845 ± 0.005 10.1e 2.4-4.7d 9.1e 10.1e 8.5e 0.918 ± 0.004 0.916 ± 0.004 0.948 ± 0.005 0.945 ± 0.003 0.932 ± 0.003 1.863 ± 0.004 1.877 ± 0.004 1.895 ± 0.005 1.881 ± 0.003 1.864 ± 0.004 Table 3: MGM fit results for HED meteorites and Acapulcoite and Lodranite meteorites modeled for comparison purposes. The columns in this table are: object designation, RELAB database designation, B1 center, B2 center, Fs number, grain size. Error on meteorite band centers is determined using the methodology outlined in section 3.2. 15 Continuum offset 5.000 ×10−1 ± 0.1 Band designation A Band 1 B Band 2 Continuum slope -3.986 ×10−5 ± 5.0 ×10−5 Band center (µm) 0.450 ± 0.075 0.935 ± 0.100 1.178 ± 0.100 2.010 ± 0.100 Band FWHM (µm) 0.397 ± 0.100 0.189 ± 0.200 0.250 ± 0.025 0.703 ± 0.400 Band strength -0.530 ± 0.10 -7.074 ± 0.50 -0.500 ± 0.75 -0.500 ± 0.50 Table 4: Input parameters for MGM fitting. The top section shows continuum parameters and the bottom shows band parameters. The value after the modulus represents the 1-sigma variation allowed for each parameter. Bands designated A and B are located below Band 1 and between Band 1 and 2, respectively. 3.2. Band Center Error Analysis Formal error bars on fitted band centers behave non-linearly as a function of signal-to-noise (S/N). These errors are non-trivial to estimate because typical observational data only contain uncertainty associated with reflectance values whereas the accuracy of band centers is an uncertainty associated with the spectral dispersion axis. Thus the S/N of spectral data cannot be used as a direct indicator of band center uncertainty. We performed a series of Monte Carlo experiments to assign formal error bars to MGM-fitted band centers as a function of S/N. We began with RELAB spectra of a Howardite (Y791573), a Eucrite (Ibitira), and a Diogenite (Tata- houine), and then fit their band centers using MGM. These particular samples were chosen because they have exceedingly high S/N (>> 100) and are well fit with our 4 band MGM model. We then added a prescribed level of random noise to the spectra and re-fit the band centers. This addition of noise and fitting of the spectra was repeated 1000 times. The standard deviation of the band centers from the 1000 trials was then used to assign 3σ uncertainties to each band center as a function of the prescribed noise level. This process was repeated for a range of typical S/N values from 5 to 100 (Figure 4). Each of the HED subtypes produced roughly equivalent results. 16 Figure 4: 3σ error in band center location for Band 1 (circles) and Band 2 (triangles) as a function of S/N, calculated according to the process outlined in section 3.2 for the Howardite Y791573. The relationship between percentage error in band center and the S/N of the spectrum is essential to properly assess uncertainties on the spectral parameters presented here, but can also be a useful tool for future observational planning. For example, if a particular science case requires that a band center be measured to a precision of ∼ 1%, then the requisite S/N in the region of the band center would be ∼ 15. To better enable the use of these Monte Carlo results for future observations we provide approximate analytical expressions to quantify the percent error in each band as a function of S/N: %ErrorB1 ∼ 16.5 ∗ (S/N )−1 %ErrorB2 ∼ 7.8 ∗ (S/N )−1 17 (1) (2) 020406080100S/N012345% Error in Band Center This uncertainty analysis is purely based on statistical (S/N) errors asso- ciated with spectral data. This analysis does not apply to systematic errors that could also influence the accuracy of band center fits. For example, outlying points in asteroid reflectance spectra due to incomplete removal of telluric bands around 0.95, 1.1, and 1.9 µm can skew band centers in ways that are difficult to quantify. Treatment of such systematic effects is beyond the scope of this work. These band center error estimates are specific to V-type and HED spectra. While these results likely provide a rough estimate on formal band center errors for other compositions (e.g. ordinary chondrites), a similar analysis would be required to properly assess the errors for different spectral types/compositions. 3.3. Temperature corrections to band centers Higher surface temperatures cause a shift in observed band centers to lower wavelengths. Thus, when comparing meteorite laboratory spectra to observa- tional data, it was important to determine the temperature of each asteroid and correct MGM-derived band centers accordingly. Surface temperatures were estimated using the following equation, from Burbine et al. (2009): T = [(1 − A)L0/16ησπr2]1/4 (3) Where A is the albedo of the asteroid, L0 is solar luminosity, η is the beaming factor (set to unity),  is the infrared emissivity of the asteroid (set to 0.9), σ is the Stefan-Boltzmann constant, and r is the heliocentric distance of the asteroid at the time of its observation. As discussed in Burbine et al. (2009), adjustments on the scale of 0.1 to η and σ alter the resulting temperature by only ± 10 K. Albedos for each object were retrieved from the WISE catalog (Mainzer et al. 2011) or IRAS (Tedesco et al., 2002). For objects whose albedo was unavailable, the albedo of 4 Vesta (0.4) was assumed, as it should be representative for V- type asteroids. Albedo errors for all objects had effects on the corrected band centers at or below the order of 0.001%, and thus were not accounted for. Having calculated the surface temperature of each object, we then corrected our MGM-derived band centers using the formulas determined by Burbine et al. 18 (2009). Burbine et al. analyzed two orthopyroxene samples and calculated two formulas for each band correction. We applied these equations to the temper- ature of each object calculated in Equation 3, and then averaged the resulting corrections for each pair of equations and applied them to our MGM-derived band centers. The temperature corrections to our asteroid band centers, as well as the temperature-corrected band centers themselves, are shown in Table 6. 3.4. Compositional validation Our next step was to address how well we could use MGM-derived band centers to constrain the bulk pyroxene Fs number of any given object. We began by calibrating meteorite spectra against their known Fs numbers. A clear correlation exists between Band 1 and Band 2 centers and Fs number among HED meteorites. Thus, after plotting each meteorite's band centers against its Fs number, we fit a linear relation between the two, yielding the following equations: and Fs# (molar %) = 664 × B1 − 583 Fs# (molar %) = 169 × B2 − 296 (4) (5) where B1 and B2 represent the MGM-derived Band 1 and Band 2 centers in microns. These equations are plotted over the Band 1 and Band 2 centers of our HED meteorite sample in Figure 5. The RMS error of these equations is ± 6 for Equation 4 and ± 4 for Equation 5. This compares favorably to the analogous equations derived by Gaffey et al. (2002) based on their analyses of terrestrial pyroxenes, which have RMS errors of ± 7 and ± 4, respectively. Gaffey et al. been used elsewhere to calculate Fs number for V-type asteroids (Burbine et (2002)'s equations have al., 2009; Ieva et al., 2015). The calibration of Fs number is better for Band 2 centers than for Band 1 centers, and so at infinite signal to noise Equation 5 is 19 more accurate. However, telescopic spectra generally display lower S/N around 2 µm, suggesting that uncertainty in compositional information derived from this band will eventually be dominated by statistical S/N errors as opposed to the systematic RMS errors associated with the absolute calibration of these equations. In circumstances of extremely low S/N in the 2-µm region, one would want to completely ignore Band 2 and rely solely on Equation 4. We adopt an intermediate approach of averaging the compositional information derived from Band 1 and Band 2. Fs numbers for each of the meteorites in our sample are shown in Table 5, along with the Fs numbers predicted by equations (4) and (5) and Gaffey et al. (2002)'s equations. 3.5. Comparison to non-Vesta-like meteorite spectra As these equations were calibrated to HED meteorites with compositions analogous to Vesta, it was important to gauge their accuracy for non-Vesta achondritic compositions. Of particular interest for the non-Vesta basaltic as- teroids (where we cannot a priori assume a pyroxene-dominated composition like that of Vesta and the HEDs) is the influence of olivine on our derived py- roxene mineralogy. Olivine can artificially increase the MGM-derived Band 1 center, and with it the Fs number derived using Equation 4. This effect pushes the object above the HED "Main Sequence" -- the region of band parameter space occupied by HED meteorites shown in Figure 6. Among our outer belt target objects, at least one -- (349) Dembowska -- is known to be an olivine-rich body, with a surface olivine:pyroxene ratio perhaps even as high as 1:1 (Gaffey et al., 1993). As such, it was important to test Equations 4 and 5 on olivine-rich meteorites in order to gauge their accuracy. Acapulcoites and Lodranites are achondrites likely from differentiated or partially differentiated parent bodies and are known to be more olivine-rich than the HED suite. We obtained reflectance spectra of three Acapulcoites and five Lodranites from the RELAB database, all with known compositions, and performed MGM fits on them using the same set of input paramters used for the HED meteorites (see Table 4). The resulting band centers for each meteorite 20 are shown alongside those of the HEDs in Table 3, and the mean band centers for the Acapulcoites and Lodranites are each shown in Figure 6. Acapulcoites and Lodranites were significantly lower in Fs content than their HED counterparts, with Fs numbers of 10.1 molar percent or less. As predicted, the outputs of Equations 4 and 5 do not accurately account for this. Equation 4 is particularly inaccurate, overestimating the Fs numbers of Acapulcoites and Lodranites by averages of 24 and 27, respectively. Equation 5 does a slightly better job, overestimating the Fs numbers of these objects by averages of 8 and 13, respectively. This is to be expected, as the complex olivine features around 1.2 µm have a significant effect on the apparent Band 1 center, but exert less influence on the apparent Band 2 center. Since the presence of olivine inflates Band 1 center far more than Band 2 center, the accuracy of our equations can be quantified by examining the "Fs discrepancy" between the Fs numbers predicted by each band center. The av- erage Fs discrepancy among our HED meteorite sample was just 2 ± 2 molar percent, while among our Vestoid sample the average Fs discrepancy was only slightly higher at 4 ± 3 molar percent. However, the Acapulcoites and Lo- dranites fell well outside this range, with an average Fs discrepancy of 15 ± 9. Notably, the Fs discrepancy values for the HEDs and Vestoids are smaller than the RMS error on Equations 4 and 5, but the Fs discrepancy values for other objects are much larger. Our equations are thus not suitable for analyzing the composition of these objects. However, Figure 6 demonstrates that our method of MGM fitting is nevertheless capable of clearly differentiating these objects in band space from HED meteorites and Vestoids. Furthermore, our derived pyroxene Fs numbers almost certainly establish an upper limit for these cases. Two of the outer belt objects listed in Table 7 also have Fs discrepancies above the range set by the HED and Vestoid samples: (349) Dembowska (8.5) and (22308) 1990 UO4 (17.7). Our method of compositional analysis thus cannot be used to uniquely determine the Fs content of these objects. However, because they fall above the HED main sequence we are still able to use Equation 4 to set a confident upper limit on their Fs content. This is in keeping with the results 21 of our analysis of Acapulcoites and Lodranites, for which Equations 4 and 5 universally overestimated Fs content. Figure 5: Fs number of HED meteorites plotted against Band 1 and Band 2 centers. Diogenites are represented by diamonds, Howardites by triangles, and Eucrites by squares. Each was fitted using the IDL REGRESS function. A clear correlation is evident between band centers and Fs number. 22 0.920.930.940.95Band 1 Center (µm)2025303540455055Fs Number1.901.952.00Band 2 Center (µm)2025303540455055 Object Diogenites EETA79002 GRO95555 Johnstown LAP91900 Tatahouine Eucrites Chervony Kut EET87520 GRA98098 Ibitira MET01081 Moore County PCA91078 Serra de Mage Howardites EET87503 QUE94200 aMayne et al. (2011) b Burbine et al. (2009) Fs # MGM Calculated B1 Fs # MGM Calculated B2 Fs # Gaffey B1 Fs # Gaffey B2 Fs # 22a 25a 24b 23a 23b 42a 40a 51a 42a 45a 37a 40a 36a 37a 31a 26 ± 6 29 ± 6 25 ± 6 27 ± 6 27 ± 6 37 ± 6 45 ± 6 39 ± 6 39 ± 6 37 ± 6 37 ± 6 49 ± 7 32 ± 6 33 ± 6 29 ± 6 23 ± 4 26 ± 4 22 ± 4 26 ± 4 24 ± 4 43 ± 4 45 ± 4 41 ± 4 40 ± 4 39 ± 4 39 ± 4 46 ± 4 36 ± 4 34 ± 4 29 ± 4 25 ± 8 29 ± 8 24 ± 7 27 ± 8 26 ± 7 42 ± 8 54 ± 8 45 ± 8 46 ± 7 42 ± 8 43 ± 8 62 ± 8 35 ± 8 36 ± 8 29 ± 8 28 ± 4 31 ± 4 27 ± 4 32 ± 4 29 ± 4 52 ± 4 55 ± 4 50 ± 4 49 ± 4 48 ± 4 48 ± 4 56 ± 4 44 ± 4 42 ± 4 35 ± 4 Table 5: Measured and predicted Fs numbers for HED meteorites. In all cases, systematic error on the equations used to determine Fs number (see section 3.4) dominates over errors in determining band center. Both the MGM-derived and Gaffey et a. (2002) equations are sufficient to confine a given object to within one of the Fs regions shown in Figure 6, provided it falls on or near the HED "main sequence" on the band diagram. 23 Figure 6: Band diagram of HED meteorites. Diogenites are represented by circles, Howardites by triangles, Eucrites by squares, and 4 Vesta by a five-pointed star. The Fs number of each meteorite is shown next to each object. Three "Fs regions" are defined, which roughly corre- spond to the locations of Diogenites, Howardites, and Eucrites. The three regions combined cover the HED "main sequence," the area of band space where most HED meteorites and Vestoids are located. The top and right axes show Fs numbers derived from Band 1 and Band 2 centers using Equations 4 and 5, respectively. 24 1.851.901.952.00Band 2 Center (µm)0.920.930.940.95Band 1 Center (µm)AcapulcoitesLodranites 42 40 51 42 45 37 40 36 37 31 22 25 24 23 23DiogenitesHowarditesEucrites4 Vesta1325364860Band 1 Derived Fs Number1324344455Band 2 Derived Fs Number Object (4) Vesta (809) Lundia (956) Elisa (1468) Zomba (1929) Kollaa (2045) Peking (2371) Dimitrov (2442) Corbett (2511) Patterson (2566) Kirghizia (2579) Spartacus (2653) Principia (2763) Jeans (2795) Lepage (2823) van der Laan (2851) Harbin (2912) Lapalma (3155) Lee (3344) Modena (3657) Ermolova (3703) Volkonskaya (3782) Celle (4038) Kristina (4188) Kitezh (4215) Kamo (4796) Lewis (5481) Kiuchi (5498) Gustafsson (7800) Zhongkeyuan (9481) Menchu (9553) Colas (16416) 1987 SM3 (26886) 1994 TJ2 (27343) 2000 CT102 (33881) 2000 JK66 (36412) 2000 OP49 (38070) 1999 GG2 (50098) 2000 AG98 (97276) 1999 XC143 (349) Dembowska (1459) Magnya (21238) Panarea (22308) 1990 UO4 (40521) 1999 RL95 (105041) 2000 KO41 (7472) Kumakiri (10537) 1991 RY16 (14390) 1990 QP19 Albedo Temp. (K) B1 Corr. (µm) B2 Corr. (µm) Vesta and Vestoids 0.42 0.33 0.15 0.40 0.39 0.25 0.35 0.40 0.29 0.26 0.53 0.40 0.41 0.40 0.32 0.40 0.40 0.40 0.40 0.40 0.24 0.50 0.40 0.34 0.40 0.40 0.40 0.23 0.39 0.20 0.18 0.47 0.24 0.34 0.53 0.37 0.40 0.36 0.40 0.14 0.22 0.37 0.16 0.28 0.40 0.28 0.31 0.22 161 186 202 199 170 168 178 168 176 171 159 158 167 167 175 162 172 157 171 170 178 155 174 178 158 160 163 193 159 171 193 158 185 185 178 175 182 183 176 160 137 187 171 221 148 144 154 158 0.002 0.002 0.002 0.002 0.002 0.002 0.002 0.002 0.002 0.002 0.002 0.002 0.002 0.002 0.002 0.002 0.002 0.002 0.002 0.002 0.002 0.002 0.002 0.002 0.002 0.002 0.002 0.002 0.002 0.002 0.002 0.002 0.002 0.002 0.002 0.002 0.002 0.002 0.002 0.025 0.020 0.017 0.018 0.023 0.023 0.024 0.023 0.022 0.023 0.025 0.025 0.024 0.024 0.022 0.025 0.023 0.025 0.023 0.023 0.022 0.026 0.022 0.022 0.025 0.025 0.024 0.019 0.025 0.023 0.019 0.025 0.020 0.020 0.022 0.022 0.021 0.021 0.022 Outer Belt Objects 0.002 0.003 0.002 0.002 0.001 0.003 0.025 0.029 0.022 0.023 0.014 0.027 Unclassified Objects (See Section 4.1) 0.003 0.002 0.002 0.028 0.026 0.025 B1 Center (µm) 0.927 ± 0.002 0.927 ± 0.002 0.923 ± 0.001 0.927 ± 0.002 0.934 ± 0.002 0.933 ± 0.002 0.926 ± 0.003 0.925 ± 0.002 0.927 ± 0.002 0.930 ± 0.002 0.921 ± 0.000 0.925 ± 0.002 0.933 ± 0.002 0.926 ± 0.002 0.923 ± 0.002 0.917 ± 0.002 0.921 ± 0.002 0.920 ± 0.002 0.920 ± 0.002 0.923 ± 0.002 0.917 ± 0.002 0.927 ± 0.002 0.911 ± 0.001 0.935 ± 0.002 0.920 ± 0.002 0.929 ± 0.001 0.923 ± 0.002 0.936 ± 0.001 0.923 ± 0.000 0.929 ± 0.000 0.923 ± 0.002 0.922 ± 0.001 0.918 ± 0.003 0.921 ± 0.002 0.923 ± 0.002 0.927 ± 0.001 0.932 ± 0.002 0.926 ± 0.002 0.933 ± 0.003 0.931 ± 0.001 0.933 ± 0.000 0.929 ± 0.000 0.930 ± 0.000 0.931 ± 0.000 0.907 ± 0.000 1.015 ± 0.001 0.952 ± 0.000 0.973 ± 0.001 B2 Center (µm) 1.968 ± 0.003 1.965 ± 0.006 1.941 ± 0.005 1.977 ± 0.005 1.967 ± 0.005 1.970 ± 0.007 1.969 ± 0.024 1.951 ± 0.007 1.957 ± 0.005 1.958 ± 0.005 1.997 ± 0.014 1.980 ± 0.004 2.003 ± 0.003 1.965 ± 0.003 1.964 ± 0.027 1.929 ± 0.004 1.942 ± 0.005 1.914 ± 0.008 1.909 ± 0.006 1.948 ± 0.014 1.953 ± 0.009 1.947 ± 0.005 1.965 ± 0.006 1.965 ± 0.004 1.952 ± 0.008 1.948 ± 0.018 1.949 ± 0.011 1.981 ± 0.005 1.918 ± 0.032 1.939 ± 0.025 1.929 ± 0.008 1.971 ± 0.020 1.917 ± 0.025 1.926 ± 0.012 1.955 ± 0.013 1.971 ± 0.020 1.963 ± 0.016 1.961 ± 0.005 2.013 ± 0.013 1.911 ± 0.001 1.954 ± 0.000 1.934 ± 0.005 1.852 ± 0.001 1.941 ± 0.003 1.866 ± 0.002 1.963 ± 0.000 1.931 ± 0.001 1.994 ± 0.000 Table 6: MGM fit results for main belt basaltic asteroids. The columns in this table are: object designation, albedo, estimated mean surface temperature (see Equation 3), temperature correction to B1 center, temperature correction to B2 center, temperature-corrected B1 center, temperature correction to B2 center. 25 4. Discussion 4.1. Spectral Outliers Three objects initially classified as V-types based on visible wavelength spec- tra ((7472) Kumakiri, (14390) 1990 QP19, and (10537) 1991 RY16) are found to be spectrally unique from the remainder of our sample due to a wide and complex 1-µm absorption band and a wide and shallow 2-µm absorption band. The shape of the 1-µm band is likely due to overlap with an unusually large absorption feature or features in the 1.2-µm region. This is particularly visible in spectra of (14390) 1990 QP19 and (10537) 1991 RY16, both of which show a distinct rounded shoulder at the upper end of their 1-µm bands -- a feature which is far less obvious in the spectrum of asteroid (7472) Kumakiri. The shallowness of these objects' 2-µm absorption band results in a very different ratio of band depths between the 1- and 2-µm bands when compared to typical V-type asteroids. For most V-types, the depth of the 1-µm band is typically between 1 and 1.5 times the depth of the 2-µm band. However, in the case of all three of these objects, that ratio is significantly higher: 2.25 ((10537) 1991 RY16), 2.21 ((14390) 1990 QP19) and, in a much more extreme case, 10.74 (Kumakiri). The Bus-DeMeo taxonomy utility run by the Planetary Spectroscopy Group at MIT (DeMeo et al. 2009) classified (10537) 1991 RY16 as a V-type, though the fit was poor, and was unable to assign a spectral type to (7472) Kumakiri or (14390) 1990 QP19, though Burbine el al. (2011) suggested that (7472) Kumakiri might be a member of the O-type. As shown in Figure 2, (10537) 1991 RY16's 2-µm band is far shallower than that of a typical V-type, and neither (7472) Kumakiri nor (14390) 1990 QP19 is an exact match to the only spectroscopically confirmed O-type asteroid, 3628 Boznemcova. The distinctiveness of these objects is further underscored by where they fall in principal component space. Principal Component Analysis (PCA) reduces the dimensionality of data, using coordinate transformations to minimize variance along two axes: Principal Component 1 and Principal Component 2 (DeMeo et al., 2009). PCA is described in detail in Tholen (1984) and Bus (1999). 26 Figure 7 shows that the three objects are located in an almost entirely empty region in the range PC2 = -0.2 to -0.5, and PC1 = 0.3 to 1.0 (these values, and those shown in Figure 7, refer to NIR principal component values, which are distinct from their visible or visible + NIR counterparts). The only other object inhabiting this region is the V-type asteroid (2579) Spartacus, which is a known oddity among Vestoids (Burbine et al., 2001). Spartacus is known to have a much more complex 1-µm band than most V-types, and exhibits an unusually large Fs discrepancy for a Vestoid: Equations 4 and 5 predict Fs numbers of 28.8 and 41.5, respectively (straddling the range of Fs numbers that would be expected from a Howardite). However, neither it nor the Acapulcoites and Lodranites share any of the gross morphological features -- in particular, the extremely shallow 2-µm band -- which distinguish (10537) 1991 RY16, (14390) 1990 QP19, and (7472) Kumakiri from other spectral types. Based on these three asteroids' location in this sparse region of principal component space, they may require a significant expansion in principal component space of the V-type complex or the creation of a new spectral type. In previous examinations of Spartacus, it has been speculated that its com- plex 1-µm feature may be due either to compositional variation across the sur- face of the asteroid or to a much higher concentration of olivine than is present on the surface of other Vestoids (Burbine et al., 2001). While significant varia- tion in surface composition is uncommon among asteroids, the three overlapping olivine bands known to fall in the 1-µm region are a plausible explanation for this feature of Spartacus' spectrum and the spectra presented here. However, while its large Fs discrepancy is reminiscent of the Acapulcoites and Lodranites analyzed in this work, the fact that Equation 5 predicts a higher Fs number than Equation 4 is not what would be expected from olivine features in its spectrum. Furthermore, distinguishing between the three overlapping olivine bands and the two overlapping pyroxene bands in the 1-µm region is extremely challenging, and so we draw no conclusions regarding the specific mineralogy of these unusual objects. 27 Figure 7: Principal component diagram showing major asteroid types. Objects examined in this study are represented by five-pointed stars, with the exception of outer belt V-type (105041) 2000 KO41 (located outside the plot range at PCir2 = 0.6482, PCir1 = 1.4759). Kumakiri, Magnya, and (10537) 1991 RY16 fall in a region in PC space which is otherwise entirely empty save for the unusual V-type (2579) Spartacus. The only known O-type, 3628 Boznemcova, is also located off the plot at PCir2 = -0.6355, PCir1 = -0.1751. 4.2. Band centers and Fs numbers for outer belt objects As shown in Figure 8, the six outer belt V- and R-type objects all fall outside of the HED main sequence, and are primarily offset to the left relative to Vesta and the inner belt Vestoids. This supports the results of Ieva et al. (2015), who used MGM to locate (1459) Magnya and (21238) Panarea well outside the region of band space occupied by the Vestoid family and HED meteorites. 28 -0.4-0.20.00.20.4PCir2-0.50.00.51.01.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 Vesta105041(0.6482, 1.4759)2579 This distinction is further underlined by comparing a weighted mean of Band 1 and Band 2 centers for each group of objects: the mean band centers for the six outer belt V- and R-types, weighted by signal to noise, are 0.930 ± 0.025 and 1.918 ± 0.049, as opposed to 0.926 ± 0.013 and 1.96 ± 0.024 for Vesta and the Vestoids. This comparison also serves to further differentiate the three anomalous asteroids 14290, 10537, and (7472) Kumakiri, whose weighted mean band centers are much higher (particularly in Band 1) at 0.980 ± 0.032 and 1.971 ± 0.032 . Performing a Kolmogorov-Smirnov test on the distribution of Vestoid and outer belt V- and R-type Band 1 and 2 centers showed that the two populations are different in Band 1 center at 2.1 σ confidence and in Band 2 center at 2.35 σ confidence. After fitting the spectra of our outer belt asteroids using MGM, we calcu- lated their predicted Fs numbers with Equations 4 and 5. We determined error on these objects' Fs numbers by propagating the error on their band centers through Equations 4 and 5 and combining the result in quadrature with the RMS error on the equations themselves. Performing the same process on the weighted mean band centers calculated above allowed for mineralogical com- parisons between these groups of objects. However, as shown in Table 7 and Figure 8, the farther from Vesta and the Vestoids an object lies in the Band 1 vs Band 2 plane, the higher the discrepancy is between the Fs numbers predicted by Equations 4 and 5: (1459) Magnya sits relatively close to Vesta, and has a discrepancy of only 2 molar percent; however, (22308) Panarea is well to the left of Vesta and the entire Vestoid population, and has a discrepancy of 18 molar percent. Using this discrepancy as a proxy for the accuracy of these equations shows that these equations do not accurately estimate the Fs number of (7472) Kumakiri, (10537) 1991 RY16, and (14390) 1990 QP19: in the most extreme case, that of (7472) Kumakiri, the values predicted by the two equations differ by 65 molar percent. Fortunately, for most of our outer belt V- and R-types, these equations yield consistent results and thus can constrain each object's Fs number to a relatively narrow range, usually less than ten molar percent. The two exceptions to this, 29 as discussed in section 3.5, are (349) Dembowska and (22308) 1990 UO4. The former of these is known to be an olivine-rich body. The latter's location in band space is extremely close to the area inhabited by the Acapulcoites, and thus may have an Acapulcoite-like composition as well, which would imply the presence of olivine and a low Fs number. As a result, a meaningful comparison can be made between the Vestoids and most of our outer belt V-types using the average band centers of each, and the two objects which are not easily compared are nevertheless likely to be significantly lower in Fs content than the Vestoids. The weighted mean band centers for the Vestoids yield predicted Fs numbers of 32 ± 9 and 35 ± 4. This range is similar to a Howardite or low-Fs Eucrite meteorite (see Figure 6), which is to be expected given that Vestoids are com- prised primarily of material from Vesta's crust and upper mantle. These values are lower for our outer belt targets due to their significantly lower average Band 2 center. For these six objects, Equations 4 and 5 predict average Fs numbers of 35 ± 17 and 28 ± 8, respectively. While there is overlap between the Fs numbers predicted for the outer belt V- and R-types and the Vestoids, these results suggest that the typical Fs content of these objects may be lower than their inner-belt Vestoid counterparts. This is particularly true since the actual difference in Fs content may well be more significant than indicated by raw comparison of these numbers due to the possible presence of olivine artificially raising the Fs numbers of (349) Dembowska and (22308) 1990 UO4. However, these results are somewhat speculative and based on small sample sizes. As we discuss in section 4.3, more data are needed in order to further constrain the composition of other igneous asteroids in the outer main belt, which could in turn provide insights regarding the environments in which they formed. 30 Figure 8: Band diagram showing Vestoids and outer belt basaltic asteroids plotted over Dio- genite, Howardite, and Eucrite regions. Vesta is represented as a five-pointed star, main-belt Vestoids as squares, and outer belt basaltic asteroids as filled circles. The average band centers of comparison Lodranites and Acapulcoites are shown as filled triangles. 31 1.801.851.901.952.002.05Band 2 Center (µm)0.900.910.920.930.940.95Band 1 Center (µm)1325364860Band 1 Derived Fs Number1324344455Band 2 Derived Fs NumberDiogenitesHowarditesEucritesAcapulcoitesLodranites34914592123822308105041405214 Vesta Object (4) Vesta (956) Elisa (1468) Zomba (1929) Kollaa (2045) Peking (2371) Dimitrov (2442) Corbett (2511) Patterson (2566) Kirghizia (2579) Spartacus (2653) Principia (2763) Jeans (2795) Lepage (2823) van der Laan (2851) Harbin (2912) Lapalma (3155) Lee (3344) Modena (3657) Ermolova (3703) Volkonskaya (3782) Celle (4038) Kristina (4188) Kitezh (4215) Kamo (4796) Lewis (5481) Kiuchi (5498) Gustafsson (7800) Zhongkeyuan (809) Lundia (9481) Menchu (9553) Colas (16416) 1987 SM3 (26886) 1994 TJ2 (27343) (2000 CT102 (33881) 2000 JK66 (36412) 2000 OP49 (38070) 1999 GG2 (50098) 2000AG98 (97276) 1999XC143 (349) Dembowska (1459) Magnya (21238 Panarea (22308) 1990 UO4 (40521) 1999 RL5 (105041) 2000 KO41 (7472) Kumakiri (10537) 1991 RY16 (14390) 1990 QP19 Vesta and Vestoids MGM Predicted B1 Fs 33 ± 6 30 ± 6 32 ± 6 37 ± 6 37 ± 6 32 ± 6 31 ± 6 33 ± 6 35 ± 6 29 ± 6 31 ± 6 37 ± 6 32 ± 6 30 ± 6 26 ± 6 29 ± 6 28 ± 6 28 ± 6 30 ± 6 26 ± 6 33 ± 6 22 ± 6 38 ± 6 28 ± 6 34 ± 6 30 ± 6 38 ± 6 30 ± 6 32 ± 6 34 ± 6 30 ± 6 29 ± 6 26 ± 6 28 ± 6 30 ± 6 33 ± 6 36 ± 6 32 ± 6 37 ± 7 MGM Predicted B2 Fs 37 ± 4 32 ± 4 38 ± 4 36 ± 4 37 ± 4 36 ± 4 34 ± 4 35 ± 4 35 ± 4 41 ± 4 39 ± 4 42 ± 4 36 ± 4 36 ± 4 30 ± 4 32 ± 4 28 ± 4 27 ± 4 33 ± 4 34 ± 4 33 ± 4 36 ± 4 36 ± 4 34 ± 4 33 ± 4 33 ± 4 39 ± 4 28 ± 4 36 ± 4 32 ± 4 30 ± 4 37 ± 4 28 ± 4 30 ± 4 34 ± 4 37 ± 4 36 ± 4 35 ± 4 44 ± 4 35 ± 6 36 ± 6 35 ± 6 35 ± 6 35 ± 6 19 ± 6 91 ± 7 49 ± 6 63 ± 7 Outer belt objects 27 ± 4 34 ± 4 32 ± 4 17 ± 4 25 ± 4 19 ± 4 Unclassified objects (See Section 4.1) 36 ± 4 30 ± 4 41 ± 4 Gaffey B1 Fs 35 ± 7 30 ± 7 35 ± 7 42 ± 7 41 ± 7 34 ± 7 33 ± 7 35 ± 7 38 ± 7 29 ± 7 33 ± 7 41 ± 7 34 ± 7 31 ± 7 25 ± 7 29 ± 7 28 ± 7 28 ± 7 31 ± 7 25 ± 7 35 ± 7 19 ± 7 43 ± 7 28 ± 7 37 ± 7 31 ± 7 44 ± 7 31 ± 7 35 ± 7 37 ± 7 31 ± 7 30 ± 7 26 ± 7 29 ± 7 31 ± 7 35 ± 7 40 ± 7 34 ± 7 41 ± 8 39 ± 7 41 ± 7 40 ± 7 38 ± 7 39 ± 7 14 ± 7 125 ± 7 61 ± 7 82 ± 7 Gaffey B2 Fs 44 ± 4 39 ± 4 46 ± 4 44 ± 4 45 ± 4 44 ± 4 41 ± 4 42 ± 4 43 ± 4 51 ± 4 47 ± 4 52 ± 4 44 ± 4 44 ± 4 36 ± 4 39 ± 4 34 ± 4 32 ± 4 40 ± 4 41 ± 4 40 ± 4 44 ± 4 44 ± 4 41 ± 4 40 ± 4 41 ± 4 47 ± 4 34 ± 4 44 ± 4 39 ± 4 37 ± 4 45 ± 4 34 ± 4 36 ± 4 42 ± 4 45 ± 4 43 ± 4 43 ± 4 54 ± 4 33 ± 4 42 ± 4 37 ± 4 21 ± 4 29 ± 4 24 ± 4 44 ± 4 37 ± 4 50 ± 4 Table 7: Fs numbers for asteroids calculated from MGM-derived equations and Gaffey et al. (2002)'s equations. In nearly all cases, error is dominated by systematic uncertainty from Equations 4 and 5. The outputs of these equations correspond very closely to one another for Vesta and the Vestoids; however, the discrepancy between the predicted Fs number from each equation grows with the object's distance from the HED main sequence. 32 4.3. Compositional implications for formation These asteroids' lower average Fs numbers are of interest due to ferrosilite's ability to act as a tracer of the oxidation state of a mineral's formation envi- ronment. In oxidizing conditions, iron is preferentially incorporated into silicate minerals as they form, while reducing environments tend to send more iron into the metallic state. In the case of asteroids, this means that objects composed of high-Fs-number pyroxene are likely to have formed in a more oxidizing environ- ment than their low-Fs-number counterparts. That the average Fs number of our outer belt targets may be, on average, somewhat lower than the Vestoids' is indicative of these objects having formed in a more reducing environment. This fact is of particular significance when we consider it in light of the oxidation gradient that existed within the solar nebula. Due to the increased presence of water at lower temperatures, a newly-forming planetesimal was sub- ject to significantly more oxidizing conditions if it formed farther away from the sun. Currently, our target objects orbit well beyond Vesta -- in some cases, nearly one full astronomical unit farther from the Sun. However, they show much more dispersion in parameter space than the Vestoids, generally featuring lower Fs numbers and greater diversity in Fs number. This then implies that, despite their current orbits, at least some of these objects' progenitor plan- etesimals may have formed in other regions of the solar nebula or at different times than Vesta. If this is indeed the case, it may support the Bottke et al. (2006) model of scattering in the young solar system, which suggests that nu- merous differentiated asteroids formed interior to the main belt and were later scattered into stable main belt orbits. Alternatively, it may be evidence of the "Grand Tack" hypothesis of Walsh et al. (2012), in which Jupiter's inward and subsequent outward migration scatters material into the asteroid belt from two sources -- one inside Jupiter's formation region at approximately 3.5 AU, and one between Jupiter and Saturn. These observational results may be of use in establishing constraints on both of these models. 33 5. Summary We present visible and near-infrared spectral analysis of nine outer belt basaltic asteroids, 47 Vestoids, fifteen HED meteorites, and eight Acapulcoite and Lodranite meteorites. MGM band fitting allowed us to extract precise in- formation about the 1- and 2-µm absorption features in each spectrum. Our procedure for analyzing error in MGM-derived band centers is discussed for use in future works. Three outer belt objects had band centers and principal com- ponent values which were inconsistent with any currently defined taxonomic classification. We correlated the band centers of HED meteorites with known composition to their Fs numbers, then applied those relationships to the spec- tra of Vestoids and outer belt basaltic objects to determine their bulk pyroxene Fs content. Acapulcoite and Lodranite spectra were used as comparators to evaluate the accuracy of this technique for non-HED but still igneous composi- tions. Average Fs numbers of outer belt objects were, on average, slightly lower than the Fs numbers of Vestoids. Due to the gradient in oxidation state which existed in the solar nebula, this may suggest that these objects formed in a different location or time relative to Vesta. This result may support either the Bottke et al. (2006) model of planetesimal scattering in the early solar system or the Walsh et al. (2012) model of scattering due to giant planet migration. However, these conclusions are based on small sample sizes, and additional work is required to better understand compositions of these objects and the specific oxidation state of the solar nebula as a function of time and location. 6. Acknowledgements This work was partially supported by the NSF under its Research Experience for Undergraduates (REU) program at Northern Arizona University and Lowell Observatory. T. L. acknowledges the hard and skillful work of Kathy Eastwood as director of that program, the invaluable support of all of its participants, and the logistical assistance provided by Kathleen Stigmon and Natalie Shee- han. This work made use of meteorite reflectance spectra obtained by several 34 investigators at the NASA RELAB facility at Brown University, without which this study would never have been possible. Much of the data utilized in this publication were obtained and made available by the The MIT-UH-IRTF Joint Campaign for NEO Reconnaissance. The IRTF is operated by the University of Hawaii under Cooperative Agreement no. NCC 5-538 with the National Aeronautics and Space Administration, Office of Space Science, Planetary As- tronomy Program. The MIT component of this work is supported by NASA grant 09-NEOO009-0001, and by the National Science Foundation under Grants Nos. 0506716 and 0907766. N.M acknowledges support from an NSF Astronomy and Astrophysics Postdoctoral Fellowship (2012-2015) and Lowell Observatory. Acquisition of data from FIRE at Las Campanas Observatory were supported by a postdoctoral fellowship to N.M. by the Carnegie Institute of Washington Department of Terrestrial Magnetism. F.E.D. acknowledges support by the Na- tional Aeronautics and Space Administration under Grant No. NNX12AL26G issued through the Planetary Astronomy Program. Much of the data used in this study was taken using telescopes located at the summit of Mauna Kea, which holds enormous cultural and spiritual significance to the native Hawaiian community. We are immensely grateful for the opportunity to make use of this important site for astronomical observations. References References [1] Binzel, R. P., Masi, G., Foglia, S., Vernazza, P., Burbine, T. H., Thomas, C. A., DeMeo, F. E., Nesvorny, D., Birlan, M., Fulchignoni, M. 2007. Searching for V-type and Q-type Main-Belt Asteroids Based on SDSS Colors. Lunar and Planetary Science Conference 38, 1851. [2] Binzel, R. P., and 16 colleagues 2015. Spectral slope variations for OSIRIS- REx target Asteroid (101955) Bennu: Possible evidence for a fine-grained regolith equatorial ridge. Icarus 256, 22-29. 35 [3] Bottke, W. F., Nesvorn´y, D., Grimm, R. E., Morbidelli, A., O'Brien, D. P. 2006. Iron meteorites as remnants of planetesimals formed in the terrestrial planet region. Nature 439, 821-824. [4] Burbine, T. H., Buchanan, P. C., Binzel, R. P., Bus, S. J., Hiroi, T., Hinrichs, J. L., Meibom, A., McCoy, T. J. 2001. Vesta, Vestoids, and the howardite, eucrite, diogenite group: Relationships and the origin of spectral differences. Meteoritics and Planetary Science 36, 761-781. [5] Burbine, T. H., Buchanan, P. C., Dolkar, T., Binzel, R. P. 2009. Pyroxene mineralogies of near-Earth vestoids.. Meteoritics and Planetary Science 44, 1331-1341. [6] Burbine, T. H., Duffard, R., Buchanan, P. C., Cloutis, E. A., Binzel, R. P. 2011. Spectroscopy of O-Type Asteroids. Lunar and Planetary Science Conference 42, 2483. [7] Burbine, T. H., McCoy, T. J., Meibom, A., Gladman, B., Keil, K. 2002. Meteoritic Parent Bodies: Their Number and Identification. Asteroids III 653-667. [8] Bus, S. J., Binzel, R. P. 2002. Phase II of the Small Main-Belt Asteroid Spectroscopic Survey. A Feature-Based Taxonomy. Icarus 158, 146-177. [9] Bus, S. J., Binzel, R. P. 2002. Phase II of the Small Main-Belt Asteroid Spectroscopic Survey. The Observations. Icarus 158, 106-145. [10] Bus, S. J. 1999. Compositional structure in the asteroid belt: Results of a spectroscopic survey. Ph.D. Thesis 311. [11] Chabot, N. L., Haack, H. 2006. Evolution of Asteroidal Cores. Meteorites and the Early Solar System II 747-771. [12] Cushing, M. C., Vacca, W. D., Rayner, J. T. 2004. Spextool: A Spectral Extraction Package for SpeX, a 0.8-5.5 Micron Cross-Dispersed Spectro- graph. Publications of the Astronomical Society of the Pacific 116, 362- 376. 36 [13] DeMeo, F. E., Binzel, R. P., Slivan, S. M., Bus, S. J. 2009. An extension of the Bus asteroid taxonomy into the near-infrared. Icarus 202, 160-180. [14] Dunn, T. L., McSween, H. Y., McCoy, T. J., Cressey, G. 2008. Mineralog- ical and Chemical Evidence for Intragroup Oxidation State Variations in Equilibrated Ordinary Chondrites. Lunar and Planetary Science Confer- ence 39, 1306. [15] Feierberg, M. A., Larson, H. P., Fink, U., Smith, H. A. 1980. Spectroscopic evidence for two achondrite parent bodies - Asteroids 349 Dembowska and 4 Vesta. Geochimica et Cosmochimica Acta 44, 513-524. [16] Floss, C. 1999. Fe, Mg, Mn-bearing phosphates in the GRA 95209 me- teorite; occurrences and mineral chemistry. American Mineralogist 84, 1354-1359. [17] Gaffey, M. J., Burbine, T. H., Binzel, R. P. 1993. Asteroid spectroscopy - Progress and perspectives. Meteoritics 28, 161-187. [18] Gaffey, M. J., Cloutis, E. A., Kelley, M. S., Reed, K. L. 2002. Mineralogy of Asteroids. Asteroids III 183-204. [19] Gaffey, M. J. 1997. Surface Lithologic Heterogeneity of Asteroid 4 Vesta. Icarus 127, 130-157. [20] Goswami, J. N., Vanhala, H. A. T. 2000. Extinct Radionuclides and the Origin of the Solar System. Protostars and Planets IV 963. [21] Grimm, R. E., McSween, H. Y. 1993. Heliocentric zoning of the asteroid belt by aluminum-26 heating. Science 259, 653-655. [22] Hammergren, M., Gyuk, G., Puckett, A. 2006. (21238) 1995 WV7: A New Basaltic Asteroid Outside the 3:1 Mean Motion Resonance. ArXiv Astrophysics e-prints arXiv:astro-ph/0609420. [23] Hardersen, P. S., Gaffey, M. J., Abell, P. A. 2004. Mineralogy of Asteroid 1459 Magnya and implications for its origin. Icarus 167, 170-177. 37 [24] Hevey, P. J., Sanders, I. S. 2006. A model for planetesimal meltdown by 26Al and its implications for meteorite parent bodies. Meteoritics and Planetary Science 41, 95-106. [25] Huaman, M. E., Carruba, V., Domingos, R. C. 2014. Dynamical evolution of V-type photometric candidates in the outer main belt. Monthly Notices of the Royal Astronomical Society 444, 2985-2992. [26] Ieva, S., Dotto, E., Lazzaro, D., Perna, D., Fulvio, D., Fulchignoni, M. 2016. Spectral characterization of V-type asteroids - II. A statistical anal- ysis. Monthly Notices of the Royal Astronomical Society 455, 2871-2888. [27] Ivezi´c, Z., and 32 colleagues 2001. Solar System Objects Observed in the Sloan Digital Sky Survey Commissioning Data. The Astronomical Journal 122, 2749-2784. [28] Lazzaro, D., Michtchenko, T., Carvano, J. M., Binzel, R. P., Bus, S. J., Burbine, T. H., Moth´e-Diniz, T., Florczak, M., Angeli, C. A., Harris, A. W. 2000. Discovery of a Basaltic Asteroid in the Outer Main Belt. Science 288, 2033-2035. [29] Mainzer, A., and 34 colleagues 2011. Preliminary Results from NEOWISE: An Enhancement to the Wide-field Infrared Survey Explorer for Solar System Science. The Astrophysical Journal 731, 53. [30] Marzari, F., Cellino, A., Davis, D. R., Farinella, P., Zappala, V., Vanzani, V. 1996. Origin and evolution of the Vesta asteroid family.. Astronomy and Astrophysics 316, 248-262. [31] Mayne, R. G., Sunshine, J. M., McSween, H. Y., Bus, S. J., McCoy, T. J. 2011. The origin of Vesta's crust: Insights from spectroscopy of the Vestoids. Icarus 214, 147-160. [32] Mayne, R. G., Sunshine, J. M., McSween, H. Y., McCoy, T. J., Corrigan, C. M., Gale, A. 2010. Petrologic insights from the spectra of the unbrec- 38 ciated eucrites: Implications for Vesta and basaltic asteroids. Meteoritics and Planetary Science 45, 1074-1092. [33] McCoy, T. J., Keil, K., Clayton, R. N., Mayeda, T. K., Bogard, D. D., Garrison, D. H., Huss, G. R., Hutcheon, I. D., Wieler, R. 1996. A petro- logic, chemical, and isotopic study of Monument Draw and comparison with other acapulcoites: Evidence for formation by incipient partial melt- ing. Geochimica et Cosmochimica Acta 60, 2681-2708. [34] McCoy, T. J., Keil, K., Clayton, R. N., Mayeda, T. K., Bogard, D. D., Gar- rison, D. H., Wieler, R. 1997. A petrologic and isotopic study of lodranites: Evidence for early formation as partial melt residues from heterogeneous precursors. Geochimica et Cosmochimica Acta 61, 623-637. [35] McSween, H. Y., and 21 colleagues 2013. Composition of the Rheasilvia basin, a window into Vesta's interior. Journal of Geophysical Research (Planets) 118, 335-346. [36] Moskovitz, N. A., Jedicke, R., Gaidos, E., Willman, M., Nesvorn´y, D., Fevig, R., Ivezi´c, Z. 2008b. The distribution of basaltic asteroids in the Main Belt. Icarus 198, 77-90. [37] Moskovitz, N. A., Lawrence, S., Jedicke, R., Willman, M., Haghighipour, N., Bus, S. J., Gaidos, E. 2008c. A Spectroscopically Unique Main-Belt Asteroid: 10537 (1991 RY16). The Astrophysical Journal 682, L57. [38] Moskovitz, N. A., Willman, M., Burbine, T. H., Binzel, R. P., Bus, S. J. 2010. A spectroscopic comparison of HED meteorites and V-type asteroids in the inner Main Belt. Icarus 208, 773-788. [39] Moskovitz, N., Gaidos, E. 2011. Differentiation of planetesimals and the thermal consequences of melt migration. Meteoritics and Planetary Sci- ence 46, 903-918. 39 [40] Moskovitz, N. A., Walker, R. J. 2011. Size of the group IVA iron meteorite core: Constraints from the age and composition of Muonionalusta. Earth and Planetary Science Letters 308, 410-416. [41] Neumann, W., Breuer, D., Spohn, T. 2012. Differentiation and core forma- tion in accreting planetesimals. Astronomy and Astrophysics 543, A141. [42] Pieters, C. M. 1983. Strength of mineral absorption features in the trans- mitted component of near-infrared reflected light - First results from RE- LAB. Journal of Geophysical Research 88, 9534-9544. [43] Rayner, J. T., Toomey, D. W., Onaka, P. M., Denault, A. J., Stahlberger, W. E., Vacca, W. D., Cushing, M. C., Wang, S. 2003. SpeX: A Medium- Resolution 0.8-5.5 Micron Spectrograph and Imager for the NASA In- frared Telescope Facility. Publications of the Astronomical Society of the Pacific 115, 362-382. [44] Roig, F., Gil-Hutton, R. 2006. Selecting candidate V-type asteroids from the analysis of the Sloan Digital Sky Survey colors. Icarus 183, 411-419. [45] Roig, F., Nesvorn´y, D., Gil-Hutton, R., Lazzaro, D. 2008. V-type asteroids in the middle main belt. Icarus 194, 125-136. [46] Rubin, A. E., Wasson, J. T. 1995. Variations of Chondrite Properties with Heliocentric Distance. Meteoritics 30, . [47] Simcoe, R. A., Burgasser, A. J., Bernstein, R. A., Bigelow, B. C., Fishner, J., Forrest, W. J., McMurtry, C., Pipher, J. L., Schechter, P. L., Smith, M. 2008. FIRE: a near-infrared cross-dispersed echellette spectrometer for the Magellan telescopes. Ground-based and Airborne Instrumentation for Astronomy II 7014, 70140U. [48] Solontoi, M. R., Hammergren, M., Gyuk, G., Puckett, A. 2012. AVAST survey 0.4-1.0 µm spectroscopy of igneous asteroids in the inner and mid- dle main belt. Icarus 220, 577-585. 40 [49] Sunshine, J. M., McFadden, L.-A., Pieters, C. M. 1993. Reflectance Spec- tra of the Elephant Moraine A79001 Meteorite: Implications for Remote Sensing of Planetary Bodies. Icarus 105, 79-91. [50] Sunshine, J. M., Pieters, C. M., Pratt, S. F. 1990. Deconvolution of min- eral absorption bands - An improved approach. Journal of Geophysical Research 95, 6955-6966. [51] Tachibana, S., Huss, G. R. 2003. The Initial Abundance of 60Fe in the Solar System. The Astrophysical Journal 588, L41-L44. [52] Tedesco, E. F., Noah, P. V., Noah, M., Price, S. D. 2002. The Supplemen- tal IRAS Minor Planet Survey. The Astronomical Journal 123, 1056-1085. [53] Tholen, D. J. 1984. Asteroid taxonomy from cluster analysis of photome- try. Ph.D. Thesis . [54] Thomas, C. A., Binzel, R. P. 2010. Identifying meteorite source regions through near-Earth object spectroscopy. Icarus 205, 419-429. [55] Walsh, K. J., Morbidelli, A., Raymond, S. N., O'Brien, D. P., Mandell, A. M. 2012. Populating the asteroid belt from two parent source regions due to the migration of giant planets-"The Grand Tack". Meteoritics and Planetary Science 47, 1941-1947. 41
1903.06746
1
1903
2019-03-15T18:44:14
Effect of Different Angular Momentum Transport mechanisms on the Distribution of Water in Protoplanetary Disks
[ "astro-ph.EP" ]
The snow line in a protoplanetary disk demarcates regions with H$_2$O ice from regions with H$_2$O vapor. Where a planet forms relative to this location determines how much water and other volatiles it forms with. Giant planet formation may be triggered at the water snow line if vapor diffuses outward and is cold-trapped beyond the snow line faster than icy particles can drift inward. In this study we investigate the distribution of water across the snow line, considering three different radial profiles of the turbulence parameter $\alpha(r)$, corresponding to three different angular momentum transport mechanisms. We consider the radial transport of water vapor and icy particles by diffusion, advection, and drift. We show that even for similar values of $\alpha$, the gradient of $\alpha$(r) across the snow line significantly changes the snow line location, the sharpness of the volatile gradient across the snow line, and the final water/rock ratio in planetary bodies. A profile of radially decreasing $\alpha$, consistent with transport by hydrodynamic instabilities plus magnetic disk winds, appears consistent with the distribution of water in the solar nebula, with monotonically-increasing radial water content and a diverse population of asteroids with different water content. We argue that $\Sigma(r)$ and water abundance $N_{\rm H_2O}(r)/N_{\rm H_2}(r)$ are likely diagnostic of $\alpha(r)$ and thus the mechanism for angular momentum transport in inner disks.
astro-ph.EP
astro-ph
Draft version March 19, 2019 Preprint typeset using LATEX style AASTeX6 v. 1.0 EFFECT OF DIFFERENT ANGULAR MOMENTUM TRANSPORT MECHANISMS ON THE DISTRIBUTION OF WATER IN PROTOPLANETARY DISKS Anusha Kalyaan School of Earth and Space Exploration, Arizona State University, PO Box 871404, Tempe, AZ 85287-1404 9 1 0 2 r a M 5 1 . ] P E h p - o r t s a [ 1 v 6 4 7 6 0 . 3 0 9 1 : v i X r a School of Earth and Space Exploration, Arizona State University, PO Box 871404, Tempe, AZ 85287-1404 Steven J. Desch (Received December 14, 2018; Revised March 6, 2019; Accepted March 7, 2019) ABSTRACT The snow line in a protoplanetary disk demarcates regions with H2O ice from regions with H2O vapor. Where a planet forms relative to this location determines how much water and other volatiles it forms with. Giant planet formation may be triggered at the water snow line if vapor diffuses outward and is cold-trapped beyond the snow line faster than icy particles can drift inward. In this study we investigate the distribution of water across the snow line, considering three different radial profiles of the turbulence parameter α(r), corresponding to three different angular momentum transport mechanisms. We consider the radial transport of water vapor and icy particles by diffusion, advection, and drift. We show that even for similar values of α, the gradient of α(r) across the snow line significantly changes the snow line location, the sharpness of the volatile gradient across the snow line, and the final water/rock ratio in planetary bodies. A profile of radially decreasing α, consistent with transport by hydrodynamic instabilities plus magnetic disk winds, appears consistent with the distribution of water in the solar nebula, with monotonically-increasing radial water content and a diverse population of asteroids with different water content. We argue that Σ(r) and water abundance NH2O(r)/NH2 (r) are likely diagnostic of α(r) and thus the mechanism for angular momentum transport in inner disks. Keywords: protoplanetary disks, planets and satellites: formation, methods: numerical 1. INTRODUCTION A snow line is the boundary in a protoplanetary disk between the region near the star where a condensible volatile, especially water, is present as vapor, and far from the star where it is present as a solid. The loca- tion of a snow line depends on the pressure-temperature conditions in the disk (Hayashi 1981; Stevenson & Lu- nine 1988). Icy solids drift inwards into the inner disk through the snow line region as they lose angular mo- mentum by moving against the pressure-supported gas disk (Weidenschilling 1977). Ice on solids sublimates into vapor as it approaches the snow line. This vapor is able to diffuse through the gas both inwards towards the star and outwards back through the snow line (Steven- son & Lunine 1988). A bidirectional flow of water is thus established across the snow line region. The location of the snow line in the disk is straightforwardly set by the pressure and temperature of the disk, and is typically where the midplane temperature is about 160 K to 180 K (Lodders 2003). In contrast, the distributions of water and volatiles across the snow line region -- whether the abundance of water ice is enhanced or depleted beyond the snow line, or whether the water vapor abundance inside the snow line is enhanced or depleted -- depend subtly on the relative rates of different radial transport processes. The mechanics of radial transport affects not just the distribution of water, but other volatiles as well. The chemical inventories of planetesimals (asteroids) and planets forming in a disk will depend on the ra- dial distributions of these species in the disk. Some ma- jor condensible species, e.g. CO, have their own snow lines ( Oberg et al. 2011). Other volatiles, e.g., NH3, are trace species, but are expected to condense with water (Dodson-Robinson et al. 2009). The chemical equilib- ria of these and other species are affected by the abun- dances of volatiles in the disk at different radii r (Cuzzi & Zahnle 2004; Najita et al. 2013). Modeling the ra- dial distribution of all volatile species therefore depends on understanding how volatile distribution at the water 2 snow line operates. Besides affecting the distribution of chemical species in the disk, snow lines also can directly affect the growth of planets. Water ice also can enhance coagulation rates of icy particles over those of bare silicate parti- cles because of ice's higher sticking coefficient (Gunlach & Blum 2015). An enhancement in solid mass density beyond the snow line is also possible due to the cold- trapping of vapor diffused across the snow line (Steven- son & Lunine 1988; Ros & Johansen 2013). This can di- rectly enhance the coagulation rate as well, but may also increase the solids-to-gas ratio above the critical thresh- old for triggering planetesimal growth via the stream- ing instability (Johansen et al. 2007). The increase in solids-to-gas ratio also can lower the ionization of the gas beyond the snow line. If the disk is evolving by magnetorotational instability, this would lead to a local decrease in the angular momentum transport, and a lo- cal build up of gas. This in turn can lead to a localized pressure maximum in which particles can concentrate, further enhancing planet growth (Kretke & Lin 2008). These factors may have led to rapid formation of Jupiter at the snow line in the solar nebula. As depicted in Figure 1, the location of the snow line and the radial distribution of water and volatiles de- pend on the thermal structure of the disk and the rela- tive rates of several transport processes, including advec- tion and diffusion of vapor and particles, and radial drift of particles by aerodynamic drag (e.g., Cuzzi & Zahnle 2004; Ciesla & Cuzzi 2006). All of these processes are strongly affected by the angular momentum transport in the disk, parameterized by the dimensionless param- eter α. Not only the strength of the turbulence (the magnitude of α), but the spatial structure of α (how it varies in the disk with r, the distance from the star) can affect these processes. The goal of this paper is to quan- tify how the strength and spatial structure of turbulent viscosity, parameterized by α(r), affect the radial distri- bution of water across a protoplanetary disk. This will improve first-principles models, allowing them to pre- dict how much water an exoplanet may have accreted, based on other observable data. 1.1. Observations of the Water Snow Line Region Observational constraints greatly assist these model- ing efforts but by themselves do not allow firm predic- tions of exoplanet water contents. Constraints on the ra- dial distribution of water in disks most commonly come from studies of the water content of solar system bod- ies, or from infrared or millimeter observations of water vapor. Within the solar system, a clear gradient in water con- tent is observed. Of the terrestrial planets, the bulk Earth (crust + mantle + core) likely accreted a few Kalyaan & Desch earth oceans in the mantle and therefore up to 0.1-0.2 % by wt (Mottl et al. 2007; Wu et al 2018). Mars may have accreted ∼ 1wt% water (Wanke and Dreibus 1994). S-type asteroids present largely inside 2.7 AU (Gradie & Tedesco 1982; De Meo and Carry 2014) are associated with ordinary chondrites which accreted with ∼ 1 wt% water (Hutchison et al. 1987; Alexander et al. 1989,2013), whereas C-type asteroids beyond 2.7 AU are associated with carbonaceous chondrites with up to 13wt% H2O (as hydrated silicates; Alexander et al. 2013). The icy satellites of Jupiter and the other plan- ets, Pluto as well as the Kuiper Belt Objects, are roughly 50wt% ice (Brown 2012). Therefore, radially outward from the sun, there is an indication of a sharp increase in the bulk water abundance existed towards the outer disk, potentially in the asteroid belt or closer. However, it is not clear how much this distribution was affected by the presence of Jupiter (e.g., Morbidelli et al. 2016), or by the particular transport mechanisms acting in our protoplanetary disk that might not act in other disks. Infrared observations of existing protoplanetary disks provide useful information about the abundance of warm water vapor inside the snow line, but the snow line itself is not resolved by such observations, as its close proxim- ity (< 5 AU) to the central star means it subtends only tens of milliarcseconds at typical disk distances (Pon- toppiddan et al. 2014; Notsu et al. 2016). Instead, high- resolution spectroscopy in combination with models is used to infer the distribution of water. An 'approxi- mate' radial location of the snow line has been inferred from high resolution spectroscopy, using the inferred temperature associated with the mid-infrared lines of water vapor present in the disk atmosphere (Meijerink et al. 2009, Zhang et al. 2013). Meijerink et al. inferred that water vapor remained high in the inner disk up to a few AU from the star, and then rapidly decreased in concentration in the disk atmosphere. They argued that the vapor may have frozen out onto dust grains that sedimented toward the disk midplane. Zhang et al. (2013) performed the same analysis for TW Hya and found a dramatic drop in vapor abundance at around 4 AU. Zhang et al. (2015) infer multiple snow lines due to different volatiles at different gaps in the substruc- ture seen in HL Tau (ALMA Partnership 2015) with a chemical model, arguing that gaps in the substruc- ture resulted from enhanced pebble accretion at these condensation fronts. This effort is however based on a number of assumptions of disk chemistry and does not directly access the water snow line. Notsu et al. (2016, 2017, 2018) have discovered a new method implement- ing high-dispersion spectroscopy to place the water snow line by the selection of specific emission lines of water vapor (with low Einstein Aul and high upper-state en- ergies) that are likely to originate from the innermost Snow Lines and Different α Profiles 3 Figure 1. Schematic figures (a, b, c) that show the various processes of radial transport of volatiles across the snow line, and the implications of the redistribution of volatiles over disk evolution. a) A schematic that shows the various radial transport processes that move volatiles bidirectionally through the snow line. All these processes can be sequentially contextualized as follows: 1) particles from the outer disk face a headwind from the pressure-supported gas and therefore spiral inwards; 2) small icy particles well-coupled to the gas also diffuse inwards 2) ice on these particles sublimate on reaching warmer regions of the inner disk; 3) some of this vapor diffuses back through the snow line to refreeze onto solids; 4) ice-bereft particles also diffuse back through, and may gain some of their icy mantles; 5) icy chondrules continue to diffuse both inwards back through the snow line as well as into the outer disk; 6) with time, the particles grow and/or are accreted into asteroids. b) In the early stages of disk evolution, viscous dissipation contributes significantly to the thermal structure of the nebula. Mass accretion rates are initially high. Eventually they drop down and so does the temperature of the inner disk, causing the snow line to move inwards with time. A peak in the ice abundance of chondrules forms just beyond the snowline; as the snow line moves inward, this peak follows with time. c) With time, the collective signature of the redistribution of volatiles through the above processes manifest as the bulk abundance of water available in different bodies at different heliocentric distances. See text for detailed discussion. warm disk. They argue that this is within ALMA's cur- rent capabilities. Finally, serendipitous observations have constrained the instantaneous position of a snow line of an FU Ori- onis star in the midst of an outburst (Cieza et al. 2016). In this case, the drastic change in the thermal structure of the disk following the outburst may have moved the water snow line out to a distance of 40 AU from the star, far enough for the snow line to be spatially resolved. The rapid improvement in observational techniques makes it likely that the location of the snow line will be constrained around many disks in the foreseeable fu- ture; but additional observations would be needed to constrain the abundance of water ice or water vapor on either side of the snow line. Substantial modeling is still required to build predictive models of exoplanet water content. 1.2. Models of the Water Snow Line Region Several models (eg. Ciesla & Cuzzi 2006; Garaud & Lin 2007; Dodson-Robinson et al. 2009; Min et al. 2011; Desch et al. 2018) of the snow line calculate the ther- mal structure of the solar nebula in order to determine the radial location where water vapor finds the right range of temperatures and pressures to condense onto small solids. These models differ in assumptions of key parameters that affect the mid-plane temperature, in- cluding the choice of opacity of the disk material κ (and its variation with r, if included) and the mass accretion rates. Min et al. (2011) who performed full 3-D radiative transfer simulations summarized the effects of these two properties in Table 2 of their paper, and compared the results of simulations of other works. In their canonical run, rice (i.e., the radius where fractional abundance of ice reaches 50% at the midplane) varies from 16.1 AU to M(cid:12) = 10−6 to 10−9 M(cid:12)/yr for fixed opac- 0.7 AU, for ity. Alternatively, for fixed accretion rate 10−8 M(cid:12)/yr, rice varies from 1.2 AU to 4.8 AU across the range of κ t=0 t>0 rSL rSL Chondrule fH2O peak Radially drifting particles Ice sublimation Particle diffusion Volatile diffusion Particle growth Vapor freezes out on solids Early evolution After a few Myr Asteroid volatile content Macc κ 4 Kalyaan & Desch employed in different works (Min et al. 2011 and Davis 2005a). Yet, while rice can vary considerably based on different inputs, it is relatively straightforward to calcu- late once those inputs are fixed. In addition to fixing the location of the snow line, models try to determine the concentration of H2O vapor (the molar ratio of H2O gas to H2), inside and just out- side rice, as well as the concentration of H2O ice particles outside and just inside rice. Higher ice/gas ratio over av- erage beyond the snow line can encourage planetesimal growth. For example, the streaming instability is sen- sitive to solids-to-gas ratio (Youdin & Goodman 2005). Higher vapor/gas ratio within the snow line may provide a more oxidizing environment for chemistry. Whether vapor is enhanced interior to the snow line or ice is more enhanced exterior to the snow line depends on the rel- ative rates of evaporation of ice and diffusion of vapor, and freezing of vapor onto icy solids and rate of drift of these solids beyond the snow line. Stevenson & Lunine (1988) first argued that solids outside of the snow line would act as cold trap for water vapor diffusing from the inner nebula outward through the snow line. They argued that this would in time dehydrate the inner neb- ula, but also provide more solid material to enhance the growth of planetesimals beyond the snow line by a factor of ∼ 75, aiding formation of planets (and likely Jupiter) at the snow line, presumably at 5 AU. Cuzzi & Zahnle (2004) later performed semi-analytical calculations for volatile transport and argued that different inputs could give rise to a variety of outcomes; besides the possibility outlined by Stevenson & Lunine (1988), there could be a time in the disk's evolution when the inner disk was enhanced in water vapor, if the flux of inward-drifting icy particles were higher than the outward flux of dif- fusing water vapor. The inner disk could then become depleted in water vapor at later times, as the inward flux of drifting particles decreased (among other things, they are accreted into larger planetesimals). This was later verified by the numerical simulations of Ciesla & Cuzzi (2006) who included the transport of vapor, dust, fast migrators and immobile asteroids within a disk evolu- tion model and tracked their transport with time. Some models also consider that icy solids may be made of up multiple silicate particles held together by ice so that evaporation of ice can lead to the release of multiple small grains, dynamically coupled to the gas (Ida & Guillot 2016; Schoonenberg & Ormel 2017). While, the enhancements and/or depletions of vapor/ice inside and outside of the snow line affect the planet forming poten- tial and the chemistry near the snow line, their effects also reach far beyond it as the snow line does not re- main static but moves significantly with time, as noted above with disk models assuming different mass accre- tion rates. The relative rates of transport processes that produce variations in radial water abundance is strongly depen- dent on the turbulent viscosity parameter α, which reg- ulates both the diffusion of vapor within the snow line and diffusion and drift of solids beyond the snow line. This can be seen by computing the timescale for volatile diffusion, tdiff ≈ r2/D ≈ r2/ν, where r is the distance from the star, D is the diffusion coefficient, which is as- sumed to be D ≈ ν, volatile viscosity. The drift velocity of solids, which depends on the disk density is also is greatly affected by ν. Both diffusion and drift depend on the sound speed cs and therefore temperature, which depends on viscous heating and therefore also the tur- bulent viscosity ν. Moreover, as we argue below, it is not just the magnitude of ν that matters but also its spatial variation in the disk. 1.3. Turbulent Viscosity in Protoplanetary Disks Determining the extent of turbulence in protoplane- tary disks is key to understanding the various processes of disk evolution and planet formation, as well as deter- mining their associated timescales. Knowledge of the turbulence in disks provides insight into mass accre- tion rates and disk dissipation timescales (Lynden-Bell & Pringle 1974; Hartmann et al. 1998), the rates of solid and volatile transport processes such as vertical settling of dust particles (Dullemond & Dominik 2004), turbu- lent concentration of these particles to grow into larger particles (Cuzzi et al. 2003), radial diffusion of vapor (Cuzzi & Zahnle 2004; Ciesla & Cuzzi 2006; Desch et al. 2017) radial diffusion and drift of particles (Estrada et al. 2016; Desch et al. 2017), as well as the rates of planet growth by pebble accretion (Xu et al. 2017) and final pebble isolation mass (Ataiee et al. 2018). The α parameter enters the disk evolution equations as the widely-used Shakura-Sunyaev (1973) parameterization of disk viscosity ν. Here, ν is assumed to be a fraction of the product of the maximum velocity (cs, i.e., local sound speed) and size (H, i.e., disk scale height) scales of turbulent eddies that could mix material in the disk (i.e., ν = αcsH). This prescription has been useful to formulate turbulence in disk evolution even without the knowledge of the physical mechanism behind it. One possible constraint on α emerges from primitive material in the solar system. The size distribution of chondrules present in various chondrites are similar to a log-normal distribution around a mean size suggesting that they may have been aerodynamically sorted. A leading idea that could yield such a sorted distribution is turbulence (Cuzzi et al. 2001). If this is the case, then α can be inferred to be ∼ 10−4 at distances of ∼ few AU from the sun at around 1-3 Myr, thus providing a region- and time-specific reference point for α for the solar nebula (Cuzzi et al. 2001; Desch et al. 2007). Snow Lines and Different α Profiles 5 Observations of other protoplanetary disks have placed some more constraints on α. Hartmann et al. (1998) considered the mass accretion rates onto T Tauri stars and the viscous spreading of their disks over times, concluding that α ∼ 10−2 represented a typical spa- tially and time-averaged value of α. Other studies reach similar conclusions: Hueso & Guillot (2005) found α = 10−3 − 10−1 for DM Tau and 4 × 10−4 to 10−2 for GM Aur; Andrews et al. (2009, 2010) found α ≈ 5×10−4 to 8 × 10−2 for disks in Ophiuchus. More recently, the Atacama Large Millimeter Array (ALMA) has been used to constrain the turbulent broadening of CO emission lines in the outer regions (> 30 AU) of resolved disks, finding α < 10−4 − 10−3 (Hughes et al. 2011; Guil- loteau et al. 2012; Simon et al. 2015; Flaherty et al. 2015, 2017, 2018; Teague et al. 2016). These authors consider these values to be an upper limit, and low val- ues of α are corroborated by other studies. Pinte et al. (2016) and recently, Dullemond et al. (2018) esti- mated α ∼ 10−5 − 10−4 from the lack of smearing of concurrent dust rings and gaps in observed disks. Ob- servational surveys of disks in the Lupus star-forming region find disks too compact to have been viscously spreading (Ansdell et al. 2018). Rafikov (2017) argued that the lack of correlation between α (determined from mass accretion rates onto the central star) and other disk properties (e.g., disk mass, size, surface density, stellar mass, radius, or luminosity) means that turbulent vis- cosity cannot be the major driver of disk evolution, and therefore α must be low. In combination, these observa- tions suggest α ≈ 10−4−10−3 in the outer regions of pro- toplanetary disks. However, constraining its exact value by observations has been difficult by itself. Attempting to detect any spatial variation in this parameter is more difficult, and requires insight from models. Many disk models have considered magnetohydrody- namics (MHD) mechanisms as the cause of angular mo- mentum transport, especially the magnetorotational in- stability (MRI) of Balbus & Hawley (1991). Under the assumption of ideal MHD, the MRI is expected to yield a uniform α with high value (α ∼ 10−2) throughout the disk (Balbus et al. 1998 Rev Mod Phys). It has long been recognized, however, that protoplanetary disks are subject to non-ideal MHD effects, especially Ohmic dis- sipation, which will suppress the MRI at the disk mid- plane over a large range of distances from the star (Jin 1996; Gammie 1996). The result would be 'dead zones' between a few times 0.1 AU and about 10 AU, in which α is very low (perhaps 10−4 or less) Additional effects of ambipolar diffusion (Kunz & Balbus 2004; Desch 2004; Bai & Stone 2011; Simon et al. 2013a,b; Gressel et al. 2015) and the Hall effect (Wardle & Ng 1999; Bai 2014, 2015; Lesur et al. 2014; Simon et al. 2015) have been considered as well. Within 10 AU, ambipolar diffusion will suppress the MRI in the surface layers away from the disk midplane. The Hall effect depends on the orien- tation of the magnetic field relative to the disk rotation, and can lead to more or less angular momentum trans- port (Bai 2014; Lesur et al. 2014; Simon et al. 2015; see §2). In general, Hall effects suppress the MRI inside 10 AU except in regions of modest density in certain circumstances. The recognition from models that the MRI is effi- ciently suppressed in large regions of protoplanetary disks, combined with observations of α in outer disks far lower than expected for MRI-active disks, has led to a resurgence of disk evolution models including purely hy- drodynamic instabilities, or hydrodynamic instabilities in concert with magnetically driven winds. Hydrody- namic instabilities such as vertical shear instability, or VSI (Stoll & Kley 2014; Flock et al. 2017), convective overstability (Klahr & Hubbard 2014; Lyra 2014), and Zombie Vortex Instability (Marcus et al. 2013, 2015) typically yield α ∼ 10−4 throughout the disk (Maly- gin et al. 2017; Lyra & Umurhan 2018). The variation of α across the disk would depend on the local cooling timescale, and probably would not be uniform across the disk. Magnetically driven winds are expected to act inside about 1 AU, augmenting α due to hydrodynamic turbulence (Suzuki et al. 2016). Overall, theoretical models strongly suggest that α could vary across the disk. In our modeling of snow lines we therefore adopt three different profiles of α(r). We consider a uniform α case, but also a case due to the MRI acting in a disk with a dead zone, and a case in which VSI is augmented by magnetic disk winds in the inner disk. These profiles of α(r) are graphically represented in Figure 2. 1.4. Overview In this paper we investigate, for the first time, the ef- fects of spatially varying α on the distribution of water and volatiles in protoplanetary disks. The relative rates of diffusion and particle drift depend on the magnitude of α, because the flux of water vapor transported by dif- fusion out beyond the snow line is proportional to α. But the flux of particles inward across the snow line depends on the density of gas, and tends to be higher for lower gas densities. Lower gas densities are associated with high α, which leads to mass rapidly accreting onto the star. Both the inward drift of icy particles and the outward diffusion of water vapor increase with α. The concentra- tion of vapor inside the snow line and the concentration of icy solids beyond the snow line depend sensitively on the balance between these two rates. Therefore it may matter whether α is decreasing or increasing with r across the snow line. The different possible physical mechanisms for angular momentum transport lead to 6 Kalyaan & Desch different profiles for α(r) which, we hypothesize, could affect the distribution of water across the snow line. We consider three distinct, physically motivated radial profiles of α (see Figure 2), and we investigate the effect of each on disk evolution and the radial distribution of water across the snow line -- and ultimately throughout the disk. We examine how the distribution of water and volatiles depends not just on the overall value of α in a disk, but on the gradient of α(r) in a disk, especially across the snow line. The paper is organized as follows. In §2 we describe the details of our code for calculating disk evolution and volatile transport, and the assump- tions underlying our snow line model. We also present our evolutionary models for the three different α pro- files. In §3 we describe the results of a suite of disk sim- ulations assuming different α(r) and a parameter study conducted to understand the effects of various assump- tions about particle properties on the radial distribution of water in the protoplanetary disk. In §4 we compare these numerical outcomes with theoretical and observa- tional studies of disks, and with solar system data from meteorites. We conclude that different profiles of α(r) yield subtle but distinct radial distributions of water in protoplanetary disk. 2. METHODS 2.1. Disk Model 2.1.1. Structure and Flow of the Bulk Gas Our underlying disk model is described in detail in our previous paper (Kalyaan et al. 2015; hereafter Paper I). The 1(+1)D explicitly propagated finite-difference disk code features a protoplanetary disk of 60 AU in size, discretized into a logarithmic grid from 0.1 - 60 AU, and evolved using the canonical equations of Lynden-Bell & Pringle (1974) of an axissymmetric viscously evolving disk. In what follows we assume the disk is vertically well-mixed. We consider a disk passively heated by starlight and actively heated by viscous dissipation, as described be- low. For passive heating, the temperature in the disk will vary with distance r from the star as T (r) ∝ r−q0, with q0 = 3/7 (Chiang & Goldreich 1997). The lumi- nosity of the young Sun is considered to vary with time as per Baraffe et al. (2002), leading to a drop in temper- atures over the first few Myr while the disk is present. In a passively heated disk, the temperature profile is (cid:18) t (cid:19)−1/7 (cid:16) r (cid:17)−3/7 1 Myr 1 AU Tpass(r) = 171.4 K. (1) We also calculate the heating from viscous accretion us- ing the results of the detailed 3D radiative transfer sim- ulations of Min et al. (2011): (cid:20) 27 128 k µσ (cid:21)1/3 Tvisc(r) = Σ(r)2 κ α(r) Ω(r) , (2) where k and σ are Boltzmann's constant and the Stefan- Boltzmann constant, µ = 2.33 proton masses is the mean molecular weight, Σ(r) is the surface density of gas, and Ω(r) is the Keplerian orbital frequency. We as- sume that a population of fine dust (∼ 0.01 − 1µm) uniformly distributed and well-coupled to the bulk gas contributes to a radially-uniform and temperature- independent opacity of the disk material equivalent to κ = 5 cm2 g−1 (as calculated by Semenov et al. (2003) for iron-rich composite aggregate grains) at 200-400 K. We note that the dust is a very small fraction (< 1%) of the overall mass of the gas. Here the turbulence pa- rameter α(r) is assumed to vary spatially according to the three cases described below in detail. Note that it affects the temperature of the disk, in addition to af- fecting the gas and particle dynamics. To combine the effects of passive and active disk heating, we combine the two temperatures to get the total temperature, as follows: T (r) =(cid:2)Tvisc(r)4 + Tpass(r)4(cid:3)1/4 . (3) We neglect temperature-dependent variations in κ at temperatures below silicate vaporization at ≈ 1400 K. Above that temperature the lack of opacity precludes a temperature gradient, so we assume a maximum mid- plane temperature of 1400 K. 2.1.2. Transport of Vapor We consider water vapor to be a trace species in the bulk disk gas. We take the equations governing their evolution via advection and diffusion as follows (similar to those adopted by Clarke & Pringle 1988, Gail 2001, Bockelee-Morvan et al. 2002; see Desch et al. 2017): ∂Σvap ∂t = 1 ∂ Mvap 2πr ∂r , where the mass flux of vapor is Mvap = 2πr ΣDvap ∂c ∂r . (4) (5) Here Σvap and Σ are the surface densities of vapor and bulk gas, c = Σvap/Σ is the concentration of the vapor, and Dvap is the diffusivity of the vapor. This does not necessarily equal the turbulent viscosity of the gas, ν, but we take the ratio to be Sc = ν/Dvap = 1, where Sc is the Schmidt number. In this work, we don't consider the effect of variation in Sc on α for gaseous species (we do however consider different Sc for solids as mentioned in later sections). Gaseous diffusivity or Sc will be dependent on the mech- anism of turbulence as well as on the presence of a strong Snow Lines and Different α Profiles 7 Figure 2. The three α profiles considered in this work depicted with possible physical scenarios that would produce such a radial profile: a) the Uniform α profile, that is constant across radius; the range of constant α values explored in this study is also depicted as red-dashed lines; ii) MRI-α profile, with an active innermost disk and a radially increasing α towards the outer disk; and iii) Hybrid α profile, a turbulent inner disk with radially decreasing α towards the outer disk. See §2.2 for detailed discussion. magnetic field, as studied by Johansen et al. 2006 (see Table 1 of their work). We also note that in order to simplify our computations, we also don't consider the vertical diffusivity of the tracer simply assuming that material is well-mixed with height. However, it need not be so and a 2-D model then will be necessary to understand the effects of varying vertical diffusion. We use the donor cell method to convert mass flows from one grid zone to another. At the inner boundary, we calculate the mass flux using the zero-torque bound- ary condition, as described in Paper I; if it is outward, we multiply by the concentration c in the innermost zone to get the mass flux of vapor into the innermost zone. 2.1.3. Transport of Solids Into this gaseous disk with vapor, we add small par- ticles of 0.06 cm diameter, analogs to chondrules, which are round, millimeter-sized particles found in abundance in chondritic meteorites (Scott & Krot 2014; see their Table 1 for chondrule sizes). In addition to diffusion and advection, intermediate-sized solid particles also drift relative to the gas, usually inward, towards the star. This happens as a result of gas orbiting the star more slowly than the Keplerian orbital velocity rΩ, by an amount η rΩ, due to partial support by the pressure gra- dient force (usually outward). Typically η ∼ 10−3. Par- ticles orbiting at the Keplerian velocity feel a headwind that robs them of angular momentum, causing them to spiral inward (e.g., Weidenschilling 1977). The rate at which particles drift inward depends on their Stokes number: St = π 2 ρpap Σ , (6) Adapted'from'Simon+2018'UNIFORM(ALPHA(DISK(MRI(ACTIVE(DISK(TURBULENT(INNER(DISK(B(Mass/angular(momentum((loss(via(disk(winds(PoorlyGionized((MRIGinacJve(dead(zone(HYBRID α-DISK UNIFORM α-DISK Kalyaan & Desch 8 where ρp = 3 g cm−3 is the internal density and ap is the radius of the particle. This can be recast in terms of the aerodynamic stopping time of particles, tstop: St = 1 Ωtstop . (7) In terms of the Stokes number, the drift velocity of par- ticles with respect to the gas is Vdrift = −St2 Vg,r − St η rΩ 1 + St2 , (8) where Vg,r is the radial component of the gas velocity (Takeuchi & Lin 2002). This expression is valid in the Epstein regime, when ap < λ, where λ is the mean free path of gas molecules; this is typically a good assump- tion throughout much of a protoplanetary disk. At 1 AU, particles of radius ∼ 30 − 100 cm and St = 1 can drift inward on timescales as short as 50 years, but larger and smaller particles drift more slowly; chondrule- sized particles with radii 0.03 cm would typically take 105 − 106 yr to drift inward (Weidenschilling 1977). We don't track the population of fine dust that produces the opacity of the disk material, and assume that it is well-coupled and homogeneously mixed throughout the disk. The diffusion rate of particles also depends on their Stokes numbers. Their diffusivity is Dp = Dgas 1 + St2 , (9) where we assume the diffusivity of the bulk gas, like the vapor, is equal to the turbulent viscosity, ν. We apply the same boundary conditions on the flow of solids as we do on the flow of vapor. 2.1.4. Vaporarization & Condensation Because our motivation is to calculate the spatial vari- ation in water-to-rock ratio, we track the following flu- ids independently: bulk disk gas; water vapor; ice-free or 'rocky' chondrules (made of silicates); 'icy' chon- drules that carry the mass of ice on particles the size of chondrules; ice-free, 'rocky' asteroids (large silicate bodies too large to drift); and 'icy' asteroids that carry the mass of ice on large bodies. In reality, ice would coat the surfaces of rocky chondrules, slightly increas- ing their radius; in practice we assume two populations of identical-size objects (chondrules or asteroids), one pure rock and the other pure ice. We initialize the disk with a uniform concentration of vapor c = 10−4, a uni- form abundance of rocky chondrules (with surface den- sity = 0.005 × Σ) and icy chondrules (also with surface density = 0.005 × Σ), and no icy or rocky asteroids. Ice in icy chondrules can convert to vapor if in a region warm enough. The rate at which this occurs depends on the local saturation water vapor pressure over ice, which is Peq(T ) = 0.1 exp (28.868 − 6132.9/T ) dyn cm−2, T > 169 K (10) from Marti & Mauersberger (1993), and Peq(T ) = 0.1 exp (34.262 − 7044.0/T ) dyn cm−2, T ≤ 169 K (11) from Mauersberger & Krankowsky (2003). The equi- librium vapor pressure relates to the surface density of water vapor as Σvap,eq = (2π)1/2 H, (12) (cid:18) Peq (cid:19) c2 H2O where H = C/Ω is the scale height of the disk, C the sound speed in the bulk gas, and cH2O the sound speed in water vapor. If Σvap,eq exceeds the total amount of water in icy chondrules and gas at radius r, we assume that all of the water there is in vapor form; otherwise we assume Σvap = Σvap,eq and assume the remaining water is in the form of water ice (in icy chondrules). 2.1.5. Particle Growth To simulate the growth of particles into planetesimals, we assume a fraction of the chondrule population at each radius r is converted into asteroid bodies every timestep of the code. Specifically, asteroids are presumed to grow on a timescale tgrow, so that in a timestep dt the mass of chondrules is reduced by an amount Σchon(dt/tgrow), and the same mass is added to the mass of asteroids. We assume typically tgrow = 1 Myr. We do not include detailed models of fragmentation or growth of these par- ticles into larger bodies, rather assuming that a fraction of the mass grows per time interval into large bodies by growth mechanisms such as streaming instability or pebble accretion (Johansen et al. 2008; Lambrechts & Johansen 2014). We do not model the size of asteroids, since bodies of any size more than a few km would take (cid:29) 1 Myr to radially drift (Weidenschilling 1977). We ignore migration of asteroids by other mechanisms (e.g., dynamical resonances or scatterings). 2.2. Turbulence Radial Profiles We consider three different radial profiles for the tur- bulence parameter α(r) and examine the response of wa- ter vapor and ice to each profile. A major goal of this paper is to study the effect of these different profiles -- not just the magnitude of α, but the variation of α with distance r from the star -- on the distribution of water in the disk. The most consistent way to do this would be to adopt power law profiles α ∝ r−a, with a carrying different slopes, but the value of α at the snow line re- maining fixed. The problem with this approach is that the different values of α inside and outside the snow line Snow Lines and Different α Profiles 9 also lead to different distributions of mass and Σ(r) pro- files. Because the temperature depends on accretional heating and therefore Σ(r), the location of the snow line would also vary. A different approach would be to adopt the profiles of α(r) predicted from first principles by dif- ferent theories. The problem with this approach is that first-principles approaches to deriving α(r) are not es- pecially robust or predictive. They also would probably depend sensitively on inputs such as temperature, den- sity, ionization levels, etc. Improvements in the models, or just differing input assumptions, are likely to lead to different α(r) profiles. As a compromise, we consider three α(r) profiles that capture the flavor of different physical mechanisms. Case I considers a uniform value of α throughout the disk. Case II considers a disk sub- ject to the MRI, with a dead zone and low α at interme- diate radius, bracketed by larger α at other radii. Case III considers a disk subject to purely hydrodynamic in- stabilities yielding uniform α, augmented by enhances transport due to magnetic disk winds in the inner disk. Across the snow line region (typically at several AU), α is uniform in Case I, increasing with r in Case II, and decreasing with r in Case III. Below we present the three α(r) profiles used, and discuss their physical motivations and justifications. We also discuss the effect each α(r) has on the disk surface density and temperature across r. 2.2.1. Uniform α profile The uniform α profile is one that is most commonly used in disk models as this profile makes the least as- sumptions regarding the specific physical mechanism that contributes to the viscosity in the disk material and transports angular momentum. As mentioned be- fore, α features in the parameterization of viscosity ν as ν = αcsH, where cs is the sound speed, and H is the disk scale height; its choice decides how large tur- bulent eddies can be and how fast they can flow, and therefore, how efficiently material is mixed and trans- ported around in the disk. Since the dominant physical mechanism behind momentum transport has been hard to determine and still requires more observational veri- fication (see §1), this profile remains attractive for wide use in disk models. We therefore define "CASE I" as follows: αI = 3 × 10−4 (all r). (13) We also consider a range of globally uniform values α between 10−5 and 10−3. This choice of α is motivated by the observations of protoplanetary disk evolution, but also by numerical simulations of hydrodynamic tur- bulence such as the VSI, and by the concentration of chondrules by turbulence (see §1.3). While continued simulations of VSI and other hydrodynamic instabilities may ultimately predict variations in α with position, at the current time a uniform α throughout the disk is not inconsistent with purely hydrodynamic turbulence. 2.2.2. MRI α Profile While the uniform α assumption is widely used and simple to implement, it is unlikely that a protoplanetary disk would have uniform α. This is certainly true if an- gular momentum transport is dominated by the MRI, as it only operates at full efficiency in regions that are sufficiently ionized. The cold, poorly-ionized midplane regions from inside 1 AU, out to beyond 10 AU ("dead zones") are unlikely to be very MRI-active; however, the regions close to the Sun may be MRI-active due to ther- mal ionization (Desch & Turner 2015), and the regions far from the Sun may be MRI-active due to cosmic-ray ionization. These effects alone introduce radial varia- tions, as the innermost and outermost portions might have relatively high turbulence, which we arbtirarily take to be α ∼ 10−3, but the dead zone regions would have lower α. In the dead zone regions, the low-density upper layers are likely to be ionized either to a uniform surface density Σa ∼ 100 g cm−2 (Gammie 1996) or to radially-dependent surface densities Σa < 10 g cm−2 due to X-ray ionization (Lesniak & Desch 2012; Kalyaan et al. 2015). The vertically averaged value of α would be αaΣa/Σ, where αa is the value in the active layer. In general the variations in Σ and Σa with r are likely to lead to radial variations of αa within the active layer as well. In order to understand the structure of a disk evolving with the MRI, Kalyaan et al. (2015) employed formulations from Bai & Stone (2011) for MRI accretion which considered the non-ideal MHD effects of ambipo- lar diffusion. The model disk was assumed to be ionized by radially dependent X-rays from the protostar, ambi- ent cosmic rays, as well as an internal ionizing source of short-lived radionuclides. A simple chemical model was assumed that included both recombination of ions and electrons in the gas phase as well as grain surfaces. α was determined from the local ion density; a vertically- averaged α(r) profile therefore assumed the shape of a power law with a positive slope, i.e., α increased with r, as the disk became more and more MRI-active (due to increasing ion density) with increasing r. The only ex- ception was the innermost disk (< 0.3 - 1 AU), which is likely to have temperatures exceeding ∼ 1000 K. In the limit of ideal magnetohydrodynamics, high temperature ionization processes such as the collisional ionization of alkali metals (Armitage 2011) as well as thermionic emis- sion of dust grains (Desch & Turner 2015) render the en- tire vertical extent of the innermost region MRI-active. Therefore, neglecting some non-ideal effects, the shape of the MRI-α profile as shown in Figure 2 seems rea- sonable to assume. The vertically averaged α(r) value calculated in Kalyaan et al. (2015) varied strongly with 10 Kalyaan & Desch Figure 3. Plots show the evolution of the surface density profiles Σ(r) for the three disk evolution models considered in this work: a) the Uniform α profile, b) MRI-α profile, and c) Hybrid α profile. Note the structure of the disk driven by the MRI, with much of the mass concentrated at ∼ 1 AU region, as well as the structure of the hybrid α disk with largely constant Σ up to 10 AU. Different colors represent different times: 0 (dashed), 20kyr (red), 50kyr (orange), 100kyr(yellow), 200kyr (light green), 500kyr (green), 1Myr (light blue), 2Myr (blue), 4Myr (dark blue), and 5Myr (violet). Figure 4. Plots show the evolution of the radial temperature profiles T (r) for the three cases: a) the Uniform α profile, b) MRI-α profile, and c) Hybrid α profile. Different colors represent different times: 0 (dashed), 20kyr (red), 50kyr (orange), 100kyr(yellow), 200kyr (light green), 500kyr (green), 1Myr (light blue), 2Myr (blue), 4Myr (dark blue), and 5Myr (violet). r as 1 ×10−5 in the inner disk within 1 AU to 0.01 in the outer disk beyond 20 AU, rather than the one-order of magnitude difference between the lowest and highest α assumed here. The second α profile we consider ("CASE II"), mo- tivated by studies of the magnetorotational instability (MRI), therefore is given by  αII(r) = r ≤ 0.3 AU 1 × 10−3, −2.718 , 0.3 AU < r ≤ 0.7 AU 1 × 10−4 (r/0.3 AU) 0.7 AU < r ≤ 1.5 AU 1 × 10−4, 1 × 10−4 (r/1.5 AU)+0.769 , 1.5 AU < r ≤ 30 AU 1 × 10−3, 30 AU < r (14) We note that this profile is consistent with α ≈ 3× 10−4 in an averaged sense. It equals 3 × 10−4 at 6.2 AU. At 3 AU, α = 1.7 × 10−4. To verify the consistency of this profile with results of recent studies, we note that Flaherty et al. (2015,2017) find an upper limit of turbulent α ∼ 10−3 in the outer disk (> 30 AU) of HD163296. Bai (2015) present simu- lations that yield α ∼ 10−4 (10−6) if ΩB > 0 (ΩB < 0) at 5 AU. We assume a scenario where turbulent trans- port begins to gradually increase in efficiency towards larger radii ( ∼1 AU < r < ∼30 AU). Beyond 30 AU, we assume turbulent α reaches 1 × 10−3 an order of magnitude higher than the inner disk. Despite these uncertainties, we feel our case II profile captures a key attribute of the MRI, which is that it should be less ac- tive in dead zone regions, with α increasing in magnitude with increasing r beyond the dead zone. 2.2.3. Hybrid α profile At the current time, models invoking hydrodynamic instabilities such as VSI, would predict a low level of turbulence acting throughout this disk, with α ∼ 10−4; these models are not developed to the point that radial variations can be strongly argued for. A disk evolv- ing purely by non-magnetic, hydrodynamic instabilities such as the VSI is probably represented well by a uni- form α disk as in our Case I profile. Disks evolving by magnetic instabilities such as the MRI are probably characterized by our Case II profile. The third example we consider is a disk that is evolving primarily by hy- drodynamic instabilities such as VSI, but augmented by magnetically controlled angular momentum transport, not through the MRI but through disk winds. For example, Simon et al. (2018) explain the low levels Snow Lines and Different α Profiles 11 of turbulence measured by Flaherty et al. (2015, 2017) in the outer portions of some disks, by invoking atten- uation of cosmic rays by strong winds launched from the inner disk. These winds, in turn, would be gener- ated by magnetocentrifugal outflows relying on large- scale magnetic fields. Suzuki et al. (2016) have recently performed global numerial simulations of disk winds, considering the global energy budget of a protoplane- tary disk. Suzuki et al. (2016) find that the disk winds launched by the outflows would induce a torque in the disk characterized by αφ,z ∼ 10−4. This is not the same as the traditional α parameter, which relates di- rectly to the αr,φ component of the Reynolds or Maxwell stresses, but it plays a similar role in that the disk wind would drive angular momentum transport, lead to heat- ing (by ambipolar diffusion) and the Reynolds stress could help particles diffuse. Other simulations find a role for the MRI in the upper layers of the disk in driv- ing the disk winds. Simulations by Bai et al. (2015) find α ∼ 10−6−10−4 at 5 - 15 AU, depending on whether the magnetic field is aligned or anti-aligned with the disk's rotation. These results suggest a hybrid model in which the disk overall is characterized by low levels of α ∼ 10−5 throughout, except in the inner disk, where α > 10−4 may obtain due to magnetic disk winds, possibly in con- cert with the MRI. Such a profile was recently con- sidered by Desch et al. (2018) in their disk evolution model to explain the abundances of CAIs (calcium-rich, aluminum-rich inclusions) and refractory elements in different meteorite types. This model demands a low level of α ∼ 10−5 in the outer disk (> 10 AU) to pre- vent mixing of gas with the inner disk, which is depleted in refractory elements; higher values of α would not al- low CI chondrites to chemically match the solar photo- sphere. The model also demands a higher value of α in the inner disk (< 1 AU), up to 5 × 10−4, so that CAIs will be efficiently transported outward in the disk; lower values of α would not lead to efficient transport of CAIs to the carbonaceous chondrite-forming region, but higher values of α would drain the inner disk of material before ordinary and enstatite chondrites could form. Based in part on the success of the Desch et al. (2018) model, and in part on the physical plausibility of VSI augmented by disk winds, the third α profile we consider ("CASE III") is given by  1 × 10−3, 1 × 10−3 (r/1 AU) 1 × 10−4, αIII(r) = r ≤ 1 AU −1 , 1 AU < r ≤ 10 AU (15) We note that like the case II profile, the typical value of α in the case III profile is ≈ 3 × 10−4, in an average 10 AU < r sense, and equals that value at 3 AU. 2.2.4. Effect of Different α(r) on Disk Structure and Evolution A radially-varying α profile considers a disk with a variable radial efficiency of mass transport. Such a disk will see more rapid mass transport of the bulk gas or larger variations in local density in its more turbulent regions (i.e., higher α) over time than its less turbulent and therefore quiescent regions. This yields distinctly different structure than the smooth Σ profiles of the tra- ditional uniform α disk (e.g. Hartmann et al. 1998). In this work, we consider two variable α evolutionary disk models: i) MRI-driven disk; and ii) Hybrid α disk. The case of an MRI-driven disk (CASE II) was inves- tigated in detail by Kalyaan et al. (2015), where we as- sumed an MRI-active inner disk, and an outer disk that was increasingly turbulent with heliocentric distance due to decreasing gas density and increased ionization from cosmic ray penetration. Here we adopt a time-averaged parameterized form of the same α profile (i.e., CASE II). This α(r) yields a tenuous inner disk and a mas- sive central region that results in mass pile-up from the outer disk. The resulting structure from such an α pro- file closely resembles that of the structure of transition disks as pointed out by Pinilla et al. (2016). We also consider another radially varying α profile (Case III) where we assume higher turbulent α in the in- ner disk that gradually decreases with increasing r. Such a profile yields a distinctly different structure where Σ(r) is flat throughout the inner disk upto ∼ 20 AU. Beyond this, Σ(r) exponentially drops, varying little from the initially assumed profile. Two varying α profiles thus produce two different disk structure, and yield different evolutionary timescales. It is in this context that we consider the transport of tracer volatiles and solids that are mixed within the bulk gas to varying extents according to the local α(r) and study its effects. 3. RESULTS In this work, we carry out a suite of simulations of an evolving disk with volatile transport under different pre- scriptions of turbulence. We explore a range of globally uniform α values as well as radially varying α profiles that are assumed to be motivated by different mecha- nisms of angular momentum transport. The results of these simulations are complex and interdependent on several parameters. Therefore, we also perform a set of simulations where we only change one parameter at a time to tease out important trends while keeping the other parameters constant, as many of these effects oc- cur concurrently in the radially varying α simulations. First, we discuss our canonical uniform α disk in de- tail. We then vary some parameters (namely, size of the 12 Kalyaan & Desch drifting particle, diffusivity of vapor and small solid par- ticles, opacity κ of the disk gas and finally the timescale of growth of asteroids) that regulate properties of either particles, vapor or the general disk material/processes. Finally, we explore volatile transport in disks with dif- ferent global α and radially varying α profiles. 3.1. Our Canonical Uniform α case We first discuss the results of our canonical case of the uniform α disk in detail, where α=3 × 10−4 at all r, which is shown in Figure 5. The figure comprises of three plots that help us trace the water content in the disk in different useful ways. In the figure, the first plot (Figure 5a) shows the radial distribution of total abun- dance of water in ice and vapor, i.e., sum of Σvap and Σ of ice in solids (chondrules + asteroids)/Σgas at each r at different times ranging from t = 20,000 yr - 5 Myr. The location of the snow line is evident as the radius at 2 AU at which this quantity changes abruptly at each timestep. The dark blue region denotes radii always be- yond the snow line and is composed of only ice between times shown (20000 yr - 5 Myr). The gray shaded region denotes the disk radii always within the snow line that is composed of only vapor throughout the course of our simulation up to 5 Myr. The light blue region between them denotes regions that start with H2O as vapor, but later see H2O only as solid ice as the snow line moves in- ward as accretional heating diminishes. Figure 5b shows the fraction of the mass in small particles that is ice, at different radii r and times t. Far out in the disk, the particles assume cosmic abundances and are assumed to be 50% ice and 50% rock. Inside the snow line, no ice exists in small particles. Just beyond the snow line, water can be cold-trapped on small particles, which see an enhancement in the ice fraction above cosmic abun- dances. These small particles are subject to rapid radial drift and are also the precursor materials for asteroidal bodies that grow directly from them, e.g., by streaming instability. Asteroidal material at a given location is as- sumed to capture at any time in the disk a fraction of the rocky chondrule mass and the icy chondrule mass at each location. These asteroids are assumed not to radi- ally migrate after that. Figure 5c shows the fraction of asteroidal material at various radii r and times that is ice. Note that the radial scale is linear, not logarithmic as in Figures 5a and 5b. At early times, asteroid mate- rial has cosmic abundances of water (50% ice fraction) outside the snow line at about 5 AU, and no ice inside that. As the snow line moves inward, more asteroidal material between 3 and 5 AU can acquire water ice. Figure 5 illustrates several processes shaping the dis- tribution of volatiles across the snow line and through- out the disk. The concentration of vapor in the inner disk depletes with time as it is either diffused inward and accreted onto the star, or is back-diffused through the snow line to deposit as more ice on solids (Figure 5). We also have to note that vapor diffusion is accom- panied by diffusion of chondrules in two ways through the snow line: as inward diffusion of icy chondrules, and as both inward and outward diffusion of silicate chon- drules (holding to our assumption that ice on chondrule evaporates instantaneously upon approaching the snow line). Silicate chondrule diffusion inward from the snow line to the inner nebula however is counteracted by the outward silicate diffusion towards the outer nebula that have equal rates. Therefore, silicate diffusion through the snow line does not enrich or reduce the abundance of ice beyond the snow line. Effectively, it is only the vapor that is back-diffused through the snow line to the outer nebula, which piles up as more ice with time just beyond the snow line. This leads to increase in the frac- tion of water-ice in chondrules from initial concentra- tion of 0.5 (assumed from equal abundances of ice and rock) to peak abundance of 0.8 just beyond 2.0 AU at 5 Myr (Figure 5). The peak abundance of ice beyond the snow line eventually forms the radially smeared-out peak abundance in the ice-fraction of asteroids in Fig- ure 5c. The radial smearing occurs as the snow line moves inward and the chondrules and the ice distribu- tion moves inward, leaving the asteroids that form in their wake to remain in place. To facilitate comparisons to this canonical case, in Figure 6 we present the timescales of different trans- port mechanisms in the disk, defined as 2πr Σ/(∂ M /∂r), where Σ and M refer to the relevant fluid (chondrules, or H2O, etc.). Throughout much of the disk, the diffu- sion of chondrules and vapor are equally rapid and the most rapid processes in the disk, with typical timescales ∼ 105 yr. Chondrule drift is almost as rapid, with asso- ciated timescales ∼ 3 × 105 yr. 3.2. Parameter Study of a Uniform α Disk We perform a parameter study for some of the most important input parameters in our disk model that is able to significantly influence the water distribution across the disk. We performed simulations for a range of chondrule sizes, Schmidt number, timescale for growth of asteroids and opacity of disk material. These param- eter studies were performed with the canonical uniform α disk with global α = 3 × 10−4 for comparison. The re- sults of this parameter study are highlighted in Figures 7, 8, 9, 10, and 11 and discussed in detail below. 3.2.1. Effect of varying chondrule size In this study, the chondrule radius was varied from 0.01 cm, to 0.03 cm (our canonical case), to 0.06 cm, to explore the changes in water distribution due to differ- ent chondrule sizes. Increasing the chondrule radius has Snow Lines and Different α Profiles 13 Figure 5. Plots show the results of our canonical uniform α case, i.e, CASE I. Here, α is considered to be 3 × 10−4 at all r. (a) shows total water content/Σg with r, where total water = Σvap + Σchon + Σast; (b) shows radial variation in the water-rock ratio in chondrules; (c) shows radial variation in the water-rock ratio in asteroids that grow from chondrules. Figure 6. Plot shows the timescales of the various radial transport processes of volatiles for the uniform α case as depicted in Figure 5. The gold dotted lines show the radial variation in chondrule diffusion rates at the simulation times plotted in all other plots (20,000 yr - 5 Myr); the blue dotted lines (overlapped by the gold) show the radial variation in vapor diffusion rates; and grey lines show the radial variation in drift rates of chondrules. All transport rates computed for a 1 AU region. We note that on approaching the snow line, drift rates of icy chondrules becomes irrelevant as they are sublimated to vapor, and beyond the snow line, vapor is immediately converted into ice on chondrules, rendering diffusion rates of vapor here invalid. It is nevertheless shown here for completion and because these rates are still valid for silicate chondrules. a dramatic effect on the water distribution, as seen in Figure 7. With increasing chondrule size, the ice frac- tion of chondrules just beyond the snow line increases greatly; at 5 Myr it is 60wt% for a = 0.01 cm, 80wt% for a = 0.03 cm, and nearly 100wt% for a = 0.06 cm. We attribute this to a change in the relative rates of dif- fusion and drift. The diffusion rate of particles can vary with particle size through the Stokes number (Equation 9), but in the limit St (cid:28) 1, appropriate for chondrules, the diffusivity does not appreciably vary with particle size. The drift rate, on the other hand, is proportional to St and therefore particle size. Larger chondrules dif- fuse as quickly as small chondrules, but large chondrules drift more rapidly; for the large chondrules the drift rates approach the diffusion rates. Whatever water va- por diffuses outward across the snow line is more rapidly brought back across the snow line into the inner disk; but, simultaneously, drift from the outermost portions of the disk is more rapid, and more water overall is brought to the snow line. This has the additional effect of limit- ing the depletion of water vapor from the inner disk in- side the snow line. While the inner disk content steadily is more depleted over time in the case with a = 0.01 cm, the case with a = 0.03 cm has roughly constant wa- ter vapor in the inner disk for about 0.5 Myr, implying the inner disk is temporarily 'flooded' with H2O. For the case with a = 0.06 cm, the vapor abundance is en- hanced for 1 Myr before rapidly depleting. Because α is the same in all three cases, the snow line is at the same location for all three cases. 3.2.2. Effect of varying particle diffusivity To more closely examine the importance of changing the diffusivity vs. drift of particles, we explored a range of Schmidt number, from Sc = 0.3, to 1 (our canonical normal'101100101r/[AU]103102101100101102103Transport Rates (Myr)chondrule driftvapor diffusionchondrule diffusion 14 Kalyaan & Desch Figure 7. Plots show results of variation of parameter chondrule size a in an uniform α disk. Increasing chondrule size leads to higher chondrule ice-rock ratios beyond the snow line; an indirect effect of more rapidly drifting particles bringing in more water to the inner nebula which is then back-diffused through the snow line to refreeze as ice. Figures and colors same as previous plots. case), to 3. The particle diffusivity is Dp = ν/Sc, so higher Schmidt number yields lower particle diffusivity, but does not affect the drift rate or evolution of the disk. Results are plotted in Figure 8. The cases with Sc = 3 have the lowest diffusivity and therefore the highest rel- ative rate of particle drift to particle diffusion. As in the case with a = 0.06 cm, which also had a high relative rate of drift to particle diffusion, this leads to a higher ice fraction of chondrules beyond the snow line. Com- paring the different Schmidt number cases, those with the highest Sc are marked by the narrowest radial ex- tent of enhanced chondrule ice fraction, and those with the lowest Sc have the broadest radial extent of high water fraction. This is understood in terms of icy chon- drules diffusing further 'upstream' beyond the snow line, widening the distribution, when Sc is low and their dif- fusivity high. The water vapor abundance in the inner disk is also slightly lower for low Schmidt number, as chondrules can diffuse back across the snow line more rapidly, and can more effectively capture water vapor beyond the snow line. 3.2.3. Effect of varying asteroid growth timescale We explored the effect of varying the timescale tgrow over which chondrules are converted into larger, asteroid-sized bodies. Asteroids are assumed to grow by converting a fraction of the mass (both silicate and ice) present at a given radius into large bodies that from that point on are immovable. Beside our canonical case of tgrow = 1 Myr, we consider tgrow = 0.3 Myr, basically as- suming that planetesimals form three times faster than in our canonical case. The results are plotted in Figure 9. Comparing the tgrowth = 0.3 Myr case to our canon- ical case, chondrules are more efficiently converted into asteroids before they can drift or diffuse. This signifi- cantly reduces Σchon, both inside and outside the snow line. A dip in the water vapor abundance just inside the snow line (Figure 9a) reveals that inward transport of ice (through diffusion and also drift) is not fast enough to replenish water vapor, which continues to diffuse out- ward across the snow line. This results in an enhanced ice-to-rock ratio beyond the snow line. These effects transport water outward so effectively, the snow line is moved slightly outward. 3.2.4. Effect of varying the opacity We explored the effect of varying the opacity of the gas (due to dust), from κ = 4 cm2 g−1, to κ = 5 cm2 g−1 (our canonical case), to κ = 6 cm2 g−1, to κ = 8 cm2 g−1. Increased opacity leads to higher disk temperatures in actively accreting disks. Far from the star, the disk tem- perature conforms to the passively heated disk limit, but the snow line location is usually located where accretion has increased the disk temperature above the passively Adust&&a&=&0.01&cm&a"="0.03"cm"(canonical)"a&=&0.06&cm& Snow Lines and Different α Profiles 15 Figure 8. Plots show results of variation of parameter Schmidt number (Sc) in an uniform α disk. Higher Sc (lower diffusivity) leads to higher peak abundance of ice in chondrules beyond the snow line, with a narrow width. Lower Sc leads to lower peak abundance but with a broader width, dependent on the diffusion of icy chondrules in the outer nebula. Figures and colors same as previous plots. Figure 9. Plots show results of variation of parameter growth timescale of asteroids tgrow in an uniform α disk. More rapid growth of asteroids depletes the pool of chondrules for diffusion through the snow line. Figures and colors same as previous plots. heated disk case. Therefore higher opacities can increase disk temperatures and move the snow line outward. Fig- ure 10 plots the effects on the radial distribution of wa- ter, which turn out to be minimal. The location of the snow line hardly moves. As Equation 2 demonstrates, the temperature in the disk varies as κ1/3. As the tem- peratures around the snow line tend to fall as r−1, the location of the snow line tends to vary as rsnow ∝ κ−1/3. Across the entire range of opacities we considered, the location of the snow line varies by only 25%, from about 2 AU to 2.5 AU. Increases in global disk temperatures slightly increase the turbulent viscosity, since ν = αCH = α(kT / ¯m)/Ω. This does not change the viscosity of gas or diffusivity of sc#Sc#=#1.0#(canonical)#Sc#=#0.3#Sc#=#3.0#tgr$tgrow&=&1.0&Myr&(canonical)&tgrow$=$0.3$Myr$$ 16 Kalyaan & Desch particles at the snow line, which is at more-or-less fixed temperature ≈ 160 K. The diffusion and drift of vapor and particles across the snow line would not differ from the canonical case. But global increases in temperature and ν would decrease the disk evolution timescale ∝ r2/ν, making the disk evolve slightly faster (≈ 25%). At any instant in time this would lead to greater ice-to-rock ratios of chondrules beyond the snow line. Overall the radial distribution of water ice (especially in asteroids) is little changed. 3.2.5. Summary The cases considered above show the effects of varying chondrule size and Schmidt number, as well as asteroid growth timescale and disk opacity. Disk opacity is found to have little effect on the radial distribution of water, and especially the final water content of asteroids. One of the most important parameters affecting the distri- bution of water is the timescale on which chondrules are converted into asteroids. For shorter timescales tgrowth, chondrules are more depleted from beyond the snow line and can carry less water inward across the snow line by drift. This enhances the water-to-rock ratio. The other two parameters (a and Sc) are of moderate importance, and highlight the subtle interplay between drift and dif- fusion of vapor and particles. We plot in Figure 11 the timescales of chondrule drift and diffusion and vapor dif- fusion in these different model disks. As previously seen, the cases where particle drift timescales are shorter than the chondrule diffusion and vapor diffusion timescales in the 1-10 AU region are the cases that lead to the highest water-to-rock ratios beyond the snow line. This is some- what unexpected, as faster-drifting chondrules might be expected to carry ice out of this region, into the inner disk, faster than water vapor could diffuse back across the snow line. We attribute the enhancement in water- to-rock ratio to an overall increase in the water being brought to the snow line region from far out in the disk by drifting chondrules. Having examined the roles of different physical pro- cesses in the disk, we next explore the effect of vary- ing the uniform value of α across the disk, considering cases ranging from α = 3 × 10−5, to α = 1 × 10−4, to α = 3 × 10−4 (our canonical case), to α = 3 × 10−3. The results are plotted in Figure 12. With increasing values of α, the inner disk within the snow line is in- creasingly depleted in water vapor. At 5 Myr, the water abundance declines from 4.5 × 10−3 for α = 3 × 10−5, to (cid:28) 10−4 for α = 1 × 10−3. As α is increased, water vapor back-diffuses outward across the snow line more quickly, but water ice on particles also diffuses inward more quickly. Increasing α also would increase the in- ward velocity of gas, Vg,r, but from Equation 9, as long as St (cid:28) 1, the drift speed of chondrules is little affected by changes in α. Therefore increasing α increases the relative importance of particle diffusion to drift. This conclusion is supported by Figure 11, which shows that drift is more rapid at low values of α, but for high val- ues of α diffusion is more rapid than drift. At higher α, icy chondrules beyond the snow line are better able to diffuse outward, giving them more opportunity to be incorporated into asteroidal material. For all values of α, there is a zone beyond the snow line in which the ice- to-rock fraction is increased. As α increases, the width of that zone increases, but the peak ice-to-rock concen- tration decreases. This trend is somewhat echoed in the plots of asteroid water-ice fraction plots (third column of Figure 12). Higher values of α yield a broader range of radii over which asteroid water fractions increase from 0wt% to 50wt%. Smaller values of α allow for slightly higher water-ice fractions, but the effect is muted. In summary, higher values of α yield higher accumulations of ice beyond the snow line, but do not yield the highest peak abundances of water ice beyond the snow line, as the water content beyond the snow line is distributed over a broader radial region. 3.3. Water distribution with an MRI α profile Case I examines the distribution of water in a disk with uniform α = 3 × 10−4. Case II, presented in Fig- ure 14, shows the distribution of water in a disk subject to turbulent viscosity like that driven by the MRI. An immediate result is that the ice-to-rock fraction of chon- drules beyond the snow line is much higher than in Case I, with an ice fraction of 0.97. The difference in α at the snow line is not large, less than a factor of 2. Larger differences in α than this did not lead to such large differences in water content (§3.1). We attribute the increased water-ice fraction to several factors. One is the surface density, Σ(r), which we plot in Figure 3. The effect on the temperature is seen in Figure 4, and the timescales for diffusion and drift are plotted in Figure 6. As seen in Figure 3, by as early as 0.5 Myr, the surface density at 2 AU is almost a factor of 4 larger in Case II than in Case I. This is due to α being lower in this region than surrounding regions in Case II, causing mass to pile up in this same region where the snow line is located. The lower values of α in this region also lead to vapor and chondrules diffusing less rapidly there, but more rapidly in the surrounding regions. At the snow line itself vapor and icy chondrules do not diffuse as efficiently. This leads to the dip in the profile of the total water content to bulk gas just within the snow line (Figure 14a). At the same time, the higher densities lead to lower Stokes numbers, which leads to smaller drift speeds. Moreover, the snow line does not move as much as in the uniform α case. This retains a higher concentration of ice-rock over a smaller region Snow Lines and Different α Profiles 17 Figure 10. Plots show results of variation of parameter κ in an uniform α disk. Higher κ leads to a warmer disk, that enhances both diffusion and drift rates to a similar extent. Diffusion is still predominant; therefore higher κ therefore leads to a slight increase in peak chondrule ice abundance beyond the snow line. Figures and colors same as previous plots. than in the uniform α case. Over and above this, the icy chondrules quickly diffuse outward into the outer disk where they can grow and be accreted in asteroids and stay there, without the risk of the ice being lost as vapor near the snow line. All of these factors combine to bring water ice or vapor to the snow line region, but to inhibit its escape from the snow line region. 3.4. Water distribution with the hybrid α profile Case III, presented in Figure 16, shows the distribu- tion of water in a disk subject to turbulent viscosity like that driven by VSI plus disk winds in the inner disk. An important difference between this case and the others is that the surface density Σ(r) is flat and significantly lower than the other cases, with Σ(r) ≈ 500 g cm−2 throughout the snow line region by 0.5 Myr (Figure 3). The higher values of α in the inner disk keep temper- atures in the inner disk hotter than in Cases I and II, and allows the inner disk to evolve and lose mass more rapidly. Compared to the other cases, initially the tem- peratures are higher in the inner disk due to the higher α, but then as surface densities decrease, the tempera- tures drop more precipitously and the snow line moves in rapidly. Between 0.02 and 5 Myr the snow line moves in from 5 AU to 1 AU as temperatures drop (Figure 4) Rapid inward drift of the snow line keeps the peak abundance at about the same level throughout the sim- ulation from 1 - 5 AU. This results in a large "transi- tion" region in the asteroid water-ice fraction (Figure 16 c) between the water ice solid material beyond 5 AU to water-depleted inner disk within the snow line. The va- por content is seen to deplete to around the same levels in the inner disk as the uniform α case. 3.5. Convergence tests We also performed convergence tests to ensure that our simulations returned similar results with increase in our grid resolution. We performed the same simulation with 200, 300 and 400 zones, and found that with higher resolution, the peak water-to-rock ratio beyond the snow kappa$κ$=$$4.0$cm2/g$κ"=""5.0"cm2/g"(canonical)"κ$=$$6.0$cm2/g$κ$=$$8.0$cm2/g$ 18 Kalyaan & Desch Figure 11. Plots show relative timescales of the different radial transport processes for all the simulations performed in our parameter study, that include varying i) size of the chondrule; ii) diffusivity of water vapor; iii) timescale for growth of chondrules to asteroids; and finally iv) opacity of the disk material. While opacity has little effect on water distribution, achon and tgrowth matter most, significantly affecting the abundance of water ice beyond the snow line. line dropped by less than 4%, and that none of our re- sults or conclusions is significantly changed. We find no other significant variation due to increase in resolution (Figure 17). 4. DISCUSSION 4.1. Sensitivity Analysis of water ice in the disk and in planetesimals that would eventually form planets. These included chondrule size, Schmidt number, the growth timescale of asteroids, the opacity, and the value of α. Our sensitivity analysis allowed us to identify which of these factors had the biggest impact on the distribution of water ice, both in chondrules and in the planetesimals formed from them. In our parameter study of uniform-α cases, we consid- ered the effects of several parameters on the distribution 4.1.1. Parameters affecting water distribution in the disk Increasing chondrule size Increasing tracer diffusivity Increasing κ Increasing tgrowth Transport Timescales (Myr) Transport Timescales (Myr) Transport Timescales (Myr) Transport Timescales (Myr) Transport Timescales (Myr) Transport Timescales (Myr) Transport Timescales (Myr) Transport Timescales (Myr) Transport Timescales (Myr) Transport Timescales (Myr) Snow Lines and Different α Profiles 19 Figure 12. Plots show the results of the different uniform α cases that employ a range of the globally uniform α as indicated. (a) shows total water content/Σg with r, where total water = Σvap + Σchon + Σast; (b) shows radial variation in the water-rock ratio in chondrules; (c) shows radial variation in the water-rock ratio in asteroids that grow from chondrules. Figures and colors same as previous plots. We find that for variations across the likely range of each input, the size of particles, i.e., chondrule radius a, had the largest effect on the water-ice fraction in chon- drules (Figure 7). In our canonical case with a = 0.03 cm, the water ice fraction in asteroids just beyond the snow line at 5 Myr slightly exceeded 80%. For smaller particles, a = 0.01 cm, the fraction barely exceeded 60%, while for slightly larger particles, a = 0.06 cm, the ice fraction approached 100%. Changing the size of small particles onto which water vapor can condense as ice, by factors of just 2 to 3, led to very large changes in the ice-to-rock ratio on chondrules. The cases with larger chondrules also saw the water-ice fraction reach elevated levels at far greater radii beyond the snow line, The next most important parameter is the growth timescale of asteroids tgrow (Figure 9). For our canon- ical case with tgrow = 1 Myr, the water-ice fraction is ≈ 80% beyond the snow line at 5 AU. For a faster con- version rate of chondrules into planetesimals, tgrow = 0.3 Myr, the water-ice fraction approaches 100%. Far fewer chondrules are available for the water ice to adhere to, leading to higher water-ice fractions. The width of the water-ice region beyond the snow line is slightly larger as well. Variations in the value of α were important but not as prominent as the changes in the above parameters (Figure 12). The water-ice fraction beyond the snow line at 5 Myr is ≈ 80% in our canonical case with α = 3 × 10−4. For lower values of α = 3 × 10−5 this fraction approaches 90%, while for larger values of α = 1 × 10−3 it is 75%. We also find that as α is increased, we find that vapor concentration decreases by just short of two orders of magnitude at t = 5 Myr between our lowest and highest adopted α (10−5 to 10−3) values. Variations in the diffusivity of small particles, via the Schmidt number Sc, led to notable but smaller changes in the water-ice fraction in chondrules (Figure 8). Be- yond the snow line at 5 AU, the water-ice fraction is Uni$alpha$α$=$3$×$10/5$α$=$1$×$10/4$α"="3"×"10(4!(canonical)"α$=$1$×$10/3$ 20 Kalyaan & Desch Figure 13. Plots show the timescales of the various radial transport processes of volatiles for different uniform α cases (including the canonical case of uniform α). Colors and lines same as Figure 6. Note that diffusion becomes quicker than drift and therefore is more important as α is increased from 3 × 10−5 (upper left plot) to 1 × 10−3 (bottom right plot). Figure 14. Plots show results of the MRI-α profile, i.e., CASE II. (a) shows total water content/Σg with r, where total water = Σvap + Σchon + Σast); (b) shows radial variation in the water-rock ratio in chondrules; (c) shows radial variation in the water-rock ratio in asteroids that grow from these chondrules. See Figures 3 and 4 for corresponding surface density and temperature radial profile plots. Colors and lines same as in Figure 5. Figure 15. Plots show the timescales of the various radial transport processes for volatiles for the 3 α(r) profiles considered in this work; i) the Uniform α profile, ii) MRI-α profile, and iii) Hybrid α profile. Colors same as in Figure 6. Transport Timescales (Myr) Transport Timescales (Myr) Transport Timescales (Myr) Transport Timescales (Myr) α = 3 × 10-5 α = 1 × 10-4 α = 3 × 10-4 α = 1 × 10-3 mri$Transport Timescales (Myr) Transport Timescales (Myr) Transport Timescales (Myr) CASE I: UNIFORM α CASE II: MRI α CASE III: HYBRID α Snow Lines and Different α Profiles 21 Figure 16. Results of the hybrid-α profile, i.e., CASE III. (a) shows total water content/Σg (total water = Σvap + Σchon + Σast) with r; (b) shows radial variation in the water-rock ratio in chondrules; (c) shows radial variation in the water-rock ratio in asteroids. See Figures 3 and 4 for corresponding surface density and temperature radial profile plots. Colors and lines same as in Figure 5. Figure 17. Plots show the results of our convergence study. Resolution increases from top to bottom with number of radial zones increasing as 100, 200, 300 and 400 radial zones dw#Convergence)) 22 ≈ 80% in our canonical case with Sc = 1, dropping slightly to 75% for Sc = 0.3, and rising to 90% for Sc = 3. The width of the enhanced water-ice region is more strongly affected by the particle diffusivity, reach- ing to much greater radii for greater diffusivity (lower Sc). Surprisingly, changes in the value of the opacity, κ, had the smallest effect on the distribution of water ice in the disk. The cases with κ = 4 cm2 g−1 and κ = 8 cm2 g−1 showed almost no difference in peak water-ice content or distribution of water ice beyond the snow line, compared to our canonical case with κ = 5 cm2 g−1. Opacity sets the temperature in the disk and the location of the snow line, but it is not a large effect. From Equation 2, the temperature in actively accreting disks is T ∝ κ1/3, and from Figure 4 tempera- ture also is dropping roughly as T ∝ r−1 in the snow line region, meaning that the location of the snow line varies roughly as κ1/3, being only tens of percent closer in or farther out across this range of opacity. More impor- tantly, the distribution of ice and vapor is little affected by these changes. We conclude that "planetary" parameters like the ra- dius of small particles (chondrules) or the rate at which they are converted into planetesimals, dominate the dis- tribution of water ice in the disk, and not such "disk" parameters like α or opacity κ. 4.1.2. Parameters affecting distribution of water ice in planetesimals The distribution of water ice in a disk translates into water content of planets only by affecting the water con- tent of asteroidal planetesimals, which are the building blocks of planets. Planetesimals, we assume, grow con- tinuously over the lifetime of the disk, converting small chondrules into large asteroids with an e-folding time of tgrow. Because asteroids grow over a long timescale, instantaneous distributions of water ice in the disk are averaged out. Most of the cases considered above lead to similar distributions of water ice in planetesimals. Peak water-ice contents rarely exceed 50%, sometimes approaching 60%. Beyond the snow line, almost all plan- etesimals in all cases have ≈ 50% ice. The quantity that is strongly affected is the radial distribution of asteroids with intermediate water ice content < 50%. In our canonical case, that distribution at 1 Myr ex- tends from about 2.5 AU, where water ice contents are barely above 0%, to about 5 AU, where they approach 50%. Most parameters have little effect on this distri- bution: this same pattern is observed across the range we consider for chondrule size a, Schmidt number Sc, asteroid growth timescale tgrow, and opacity κ (Figures 7, 8, 9, and 10). The radial distribution of asteroid water-ice content Kalyaan & Desch seems most affected by α (Figure 12). The largest val- ues of α ∼ 10−3 allow the disk to evolve more rapidly; asteroids less reflect a snapshot of the disk and more reflect an average over time. The peak water ice con- centrations do not exceed 50% anywhere. For the small- est values of α = 3 × 10−5, the asteroids beyond the snow line can reach 60% ice, with a drop-off beyond that. To a lesser extent, the asteroid growth timescale matters, with tgrowth = 0.3 Myr yielding asteroids be- yond the snow line with > 55% water ice. The existence of a local maximum in the asteroid water ice content appears sensitive to the timescale of planetesimal for- mation (tgrowth) relative to the disk evolution timescale (∼ r2/ν ∝ α−1). Faster asteroid growth and/or slower disk evolution allows a local maximum. We conclude that water ice content of planetary ma- terials is affected equally by "disk" parameters (α) and "planetary" parameters (tgrowth). The sensitivity of the water ice distribution in the pro- toplanetary disk to various input parameters can be un- derstood largely in terms of the relative rates of diffu- sion and drift. As seen in Figures 11, 13 and 15, in general diffusive transport of vapor and chondrules is faster than drift of chondrules, by almost an order of magnitude throughout much of the disk in most of the presented simulations. In the common situation where drift is much slower than diffusion, the water ice content just beyond the snow line is close to the cosmic abun- dance of water ice, with a water-ice fraction of 50%. But as the rate of drift is comparable to, or faster than, the rate of diffusion, then the water-ice fraction in solids beyond the snow line can be become quite large. 4.2. Physical processes affecting water distribution The radial distribution of water ice in chondrules be- yond the snow line can be understood in part as a rel- ative rate of particle drift vs. diffusion of particles and vapor. For example, Figure 12 shows that the peak water-ice fraction is higher at 5 Myr (≈ 90%) for the case with α = 3×10−5 than it is (≈ 75%) at 5 Myr in the disk with α = 1 × 10−3. This is not due to temporal differences: these water-ice enhancements occur late in disk evolution, so if anything the disk with higher α should develop the same water-ice enhancements at earlier times. (The evolution timescale scales as r2/ν ∝ α−1) It is also not due to a relatively higher general rates of diffusion. Disks with large α should cause water vapor to diffuse outward past the snow line faster, leading to greater enhancements; or, including the inward diffusion of icy chondrules through the snow line, one might ex- pect the two effects to cancel each other out, leading to no differences. Figure 13 demonstrates that the diffu- sion timescales of vapor and particles at the snow line Snow Lines and Different α Profiles 23 vary from ∼ 2 Myr for α = 3 × 10−5, to ∼ 0.02 Myr for α = 1 × 10−3, without a significant change in the water-ice fraction beyond the snow line. A change in just the diffusion rate of particles but not vapor leads to somewhat larger changes. Figure 8 demonstrates that as the diffusion rate of particles alone is varied by an order of magnitude, as Sc is varied from 0.3 to 3, the water-ice fraction does change, but only from about 75% to 90%. With increased Sc, particle diffusivity decreases, meaning that icy chondrules diffuse inward more slowly, leading to a lower loss of icy solids from the region beyond the snow line. Particle diffusion is not the dominant loss mechanism, however. Rather, the dominant effect appears to be particle drift, and factors affecting drift rates appear to be the main determinant of the water-ice distribution. Figure 7 shows that as chondrule radii are increased from 0.01 cm to 0.06 cm, there is very little change in the diffu- sion timescale. This is because D = ν (1 + St2)−1, and St (cid:28) 1 for chondrules of all these sizes. The drift rate changes considerably, though, as the drift velocity is pro- portional to St and therefore a/Σ. The large chondrules drift 6 times faster than the small chondrules, and the drift timescales reflect this. Larger drift rates might be expected to more quickly drain icy chondrules out of the region beyond the snow line, lowering the water-ice frac- tion there, opposite to the observed trend ; but a more important factor is the influx of ice to this region from farther locales in the disk. The drift velocity of parti- cles is Vdrift ∝ StηVK ∝ ηVK(a/Σ). The mass flux of particles is therefore ΣVdrift ∝ ηVKa. The influx of icy particles into the region beyond the snow line therefore is 6 times larger for the 0.06 cm radius chondrules than for the 0.01 cm radius chondrules. This ice ultimately vaporizes inside the snow line, but as long as as signif- icant fraction is returned to beyond the snow line, the water-ice fraction there will ultimately be higher. A re-examination of Figures 12 and 13 suggests that the cause of the increased water-ice fraction in cases with low α is primarily due to differences in drift rate. Be- cause disks with low α evolve more slowly, they main- tain high surface densities for longer. Higher Σ leads to slower drift rates. Indeed, the cases with low α have the longest drift timescales, and and the cases with highest α have the shortest drift timescales. The short drift timescales lead to more overall water ice beyond the snow line for the high-α cases, as seen in the first column of Figure 13. The high α additionally has the effect of broadening and lessening the peak water-ice fraction. These overall trends help explain why factors like the disk opacity have little effect on the water-ice content. Changes in opacity have a relatively minor effect on the temperature, which is proportional to κ1/3. Factor of 2 increases in κ have a minor effect on the disk viscos- ity, increasing it only by a factor of 1.26. This would affect the diffusion of both particles and vapor, which would not directly change the distribution of ice much. The surface density at any time would be slightly de- creased by the higher viscosity, leading to higher drift of particles that slightly increases the water-ice fraction. But because the effects are indirect, the change in the water-ice fraction is small. In this work, we don't consider the growth of the opacity-yielding fine dust grains into larger particles or into chondrules. We assume that they are created in collisions between chondrules and don't track water ice deposited on them. We argue that these assumptions are valid as our simulations reveal that the timescales of diffusion are smaller by an order of magnitude than drift, and chondrule diffusion (unlike drift) is almost indepen- dent of grain-size for the smallest sizes of dust grains (where St << 1). As long as the rates of grain growth by coagulation and fragmentation balance each other, and the collisional size distribution stays constant across the inner disk where the water snow line is located (inde- pendent of the power law of the size distribution, e.g., Brauer et al. 2008; Birnstiel et al. 2011), our model re- mains valid as diffusion is faster than drift. We find that our assumption of collisional equilibrium across the in- ner disk is consistent with Birnstiel et al. (2015; 2011), who argue for efficient fragmentation in the inner disk. Then, the smaller dust grains only carry a fraction of the total water mass but the trends of total water abundance and ice fraction in smaller dust grains across and beyond the snow line remain the same as with chondules. On the contrary, if chondrules beyond the snow line requires higher relative velocities for fragmentation, more water is carried by the largest particles which upon drifting inward more rapidly than smaller particles yield more vapor that is back diffused into the snow line, leading finally to more ice accumulation beyond the snow line. Ultimately, the water-ice content of small solids be- yond the snow line is set by subtle interplays between viscosity, diffusion, and drift, with factors affecting drift playing the most important role. 4.3. Effects of changing α(r) With these insights, we can begin to understand the water distributions in our three different α(r) profiles. Changing α(r) in the disk, either by varying a globally uniform value, or by adopting a radially varying α(r) profile, changes the distribution of ice by altering the diffusion rates of particles and vapor, and by changing the drift rates of particles. Figure 16 shows that Case III, with the hybrid α(r) profile due to hydrodynamic instabilities and and mag- netic disk winds, and Case I, have the highest overall amounts of water beyond the snow line, with water frac- 24 Kalyaan & Desch tions reaching 4 times the overall disk mass fraction of 5 × 10−3. Case II has water abundances approaching these levels, 3 times the background levels, but over a narrower range of radii. Figure 14 shows that Case II, with the α(r) profile resembling that due to the MRI, has the highest local concentrations of water-ice beyond the snow line. Some regions show water-ice fractions of solids > 95%, in con- trast to the peak water-ice fractions ≈ 60% in Case III. Case I with uniform α has intermediate peak values at 5 Myr of ≈ 80%. Since all three cases have similar over- all water contents, we attribute these effects largely to the different values of α at the snow line, which tends to smear out and decrease the peak distributions. At 5 Myr, the value of α at the snow line is only 1.7 × 10−4 in Case II, allowing for sharp distributions. At the snow line at 5 Myr, α = 3 × 10−4 in Case I, and is as large as 8 × 10−4 in Case III. The different cases yield different positions of the snow line over time. In Case I, the snow line starts at about 4.5 AU and moves in to about 2 AU by 5 Myr. In Case II, the snow line starts at about 5 AU and moves in to 3 AU, and for Case III it moves in from 5 AU to about 1.2 AU by 5 Myr. Case III shows the snow line moving in the most because the higher values of α at small radii deplete the inner disk of gas the fastest, leading to decreased accretional heating. An important side effect of the snow line moving in so much more in Case III is that it leads to a greater diver- sity of asteroid water-ice fractions in the inner disk. In Case III, asteroids forming at 2.5 AU between 0.5 Myr and 5 Myr might have anywhere from 0% to 50% ice, with similar ranges seen for asteroids forming anywhere between 2 and 4 AU. In contrast, in Case II, asteroid water content only varies in a relatively narrower range of radii, from 3 to 4 AU. Case I, with uniform α, is somewhat intermediate between these cases, with aster- oids showing variable water content between 2.5 and 3.5 AU. These results in particular suggest ways to use the water distributions of asteroids to probe the α(r) profile of the solar nebula. In this study, we do not account for the temporal variation in α(r) and merely assume that it is a func- tion of location in the disk. In the cases of MRI- and wind-driven disks, as Σ changes with time, α is likely to change as well (eg. Kalyaan et al. 2015; Suzuki et al. 2016). While this is important to consider in the future, we argue that the focus of this paper is to understand the effect of the radial variation of α on the radial distri- bution of water. As α(r) itself has subtle effects on the various interconnected physical processes at the snow line, the effect of addition of α(t) at each r adds another layer of complexity with subtle effects that can be more difficult to separate. 4.4. Observational Tests 4.4.1. Astronomical Observations of Disks Our work highlights the connections between differ- ent α(r) profiles and the distribution of water vapor in the inner disk or water-ice on chondrules or in aster- oids in the outer disk. It is not currently possible to directly measure α(r) in protoplanetary disks, although some observations may constrain this parameter. High- resolution imaging of disk regions by Pinte et al. (2016) and Dullemond et al. (2018) show that the sharpness of the gaps and rings in the outer portions of many proto- planetary disks requires low α < 10−4 more compatible with Cases I or III, but not with Case II. Higher res- olution imaging potentially could similarly constrain α in the inner disk in the future, but this would be an observational challenge. Novel tests with sophisticated chemical models to detect heating from viscous accre- tion, such as detecting molecular emission in the cooler (90-400 K) layers of the disk, as put forth by Najita & ´Ad´amkovics (2017), or detect distinct chemical imprints that betray either a spatial variation in turbulence or a difference between low- and high-viscosity disks may also constrain α(r). It would advantageous to use observations of water va- por in the inner portions of disks to constrain α there. Differences in water vapor abundance take time to de- velop, but by 5 Myr, in our Case I uniform-α disk, with α = 3 × 10−4 in the inner disk, the water vapor mass fraction is 1 × 10−4. For case II, α = 1 × 10−3 in the inner disk and the water vapor mass fraction is 4×10−4. And in Case III, α = 1 × 10−3 throughout a broader re- gion in the inner disk, and the water vapor abundance is 2 × 10−4. The relationship between α and water vapor abundance is therefore complicated, but forward mod- els with α(r) as an input should be able to test models against observations. A particular feature of our Case II disk that might be observable is the dip in water abun- dance by a factor of almost 2 just inside the snow line, between 2 and 3 AU. We highlight that the Σ profile itself as well as the water abundance NH2O/NH2 across the inner disk are by themselves a diagnostic of the α profile in the in- ner disk and thus the dominant mechanism of angular momentum transport operating in disks. 4.4.2. Solar System Observations Our results suggest it may be possible to use the spa- tial distribution of asteroid water content to constrain the α(r) profile of the solar nebula disk. The water con- tent of asteroids is discernible using spectroscopy and comparison to meteorites. Gradie & Tedesco (1982) determined that different classes of asteroids are predominantly present in specific locations of the asteroid belt; S-type asteroids dominate Snow Lines and Different α Profiles 25 in mass and number in the inner belt inside about 2.7 AU, while C-type asteroids dominate in mass and num- ber in the outer belt beyond about 2.7 AU (see Fig. 3 of DeMeo & Carry 2014; the largest asteroids corre- spond to all the mass above the horizontal dotted line at ∼ 1 × 1018 kg). S-type asteroids show little evidence of hydration fea- tures in their spectra, and are spectrally associated with ordinary chondrites, which are generally water- poor, with only 0.1 − 1 wt% water (Hutchison et al. 1987; Alexander et al. 1989, 2013). C-type asteroids do show hydration bands in their spectra, and are as- sociated with water-rich carbonaceous chondrites, with up to ∼ 10wt% (structurally bound) water (Alexan- der et al. 2013). Additionally, E-type asteroids at 2.0 AU are spectrally associated with enstatite chondrites, which accreted essentially no water (Jacquet et al. 2017). Rarer R chondrites appear to have been water-rich and formed perhaps at 2.6 AU (Desch et al. 2018). DeMeo & Carry (2014) were able to extend the study of Gradie & Tedesco (1982) to include the wealth of information attained since then for the smaller asteroids down to 5 km in size, finding that while a sharp radial gradient exists for the distributions of the largest asteroids' spec- tral types (consistent with the conclusions of Gradie & Tedesco 1982), the smallest asteroids of different classes have been significantly radially mixed. This is com- monly interpreted as a radial mixing of asteroids scat- tered into that region, with S-types at r < 2 AU scatter- ing out, mixing with C-types from r > 3 AU scattering in (e.g., the Grand Tack model of Walsh et al. 2011). This is a very viable interpretation of the data, but our mod- eling offers an alternative explanation, that asteroids in the 2-3 AU region may have formed in place, with dry asteroids forming early and wet asteroids forming later, after the snow line swept inward. We imagine that asteroids form by a rapid mechanism like streaming instability, and are snapshots of the disk composition at that place and time. The number of asteroids created per time, per area of the disk, would be N = (Σ / tgrowth) / M , where Σ is the surface density of solids, and M is the mass of a typical asteroid. For large radii that are always outside the snow line, all of the asteroids are "wet". For small radii that are always inside the snow line, all of the asteroids are "dry". For intermediate radii that the snow line sweeps through, some asteroids are wet and some are dry. Therefore at the time the asteroids in the inner disk formed, at ≈ 2−3 Myr (Desch et al. 2018), the snow line must not have extended inward of 2.0 AU; and the asteroids formed between 2 AU and 3 AU or more around this time appear to have sampled a variety of water-ice fractions. Comparing these findings about the asteroid belt to our various disk models, Case II seems least consistent with our asteroid belt. Its snow line does not move inward of 3.5 AU by 3 Myr, and the range of radii over which the asteroid water-ice fractions vary is narrow, < 1 AU. Case II is also marked by a region in which the water-ice fraction of asteroids exceeds the canonical ratio. Case I is broadly consistent with the asteroid belt, with the snow line at about 2.5 AU at 3 Myr, and a broad range of radii over which the asteroid water- ice fractions vary. The gradient of water-ice fractions would be monotonically increasing with distance so that no asteroids would have ice fractions > 50%, for α > 10−3. Finally, Case III appears most consistent with the asteroid belt, with the snow line at 2.0 AU at 3 Myr, and a monotonically increasing water-ice fraction in the inner disk. None of these cases includes the effects of Jupiter opening a gap in the disk, which Morbidelli et al. (2016) have shown to be potentially very important in setting the water content of the inner disk. Nevertheless, our analysis shows how one might use the distribution of as- teroids and water content to infer α(r). Future work is encouraged, but so far disk evolution by MRI does not appear consistent with the constraints from our asteroid belt. We also note that different initial surface density pro- files would lead to different amount of ice that would be locked beyond the snow line in chondrules and even- tually asteroidal material. A shallower surface density profile would lead to more ice being accumulated beyond the snowline, than a more steeper profile. This is an in- triguing result that warrants a more detailed parameter study. 5. SUMMARY AND CONCLUSIONS 5.1. Summary The distribution of water and other volatiles in planets depends on how these volatiles are transported through the protoplanetary disk, as diffusing and advected va- por, and as diffusing, advected and drifting icy parti- cles. The partitioning of a volatile between vapor and ice depends on the temperature, and the "snow line" in a disk, demarcating ice-rich and vapor-rich regions, is sensitive to the degree of accretional heating in a disk. The transport of particles and diffusion of vapor and particles depends on the strength of turbulence trans- porting angular momentum. These processes ultimately depend on the magnitude and spatial variation of α, the turbulence parameter. We have investigated the effects of different α(r) pro- files on the distribution of water in protoplanetary disks. We have considered three α(r) profiles in particular. Our Case I assumes a uniform value of α throughout the disk; our canonical case considers α = 3× 10−4, but 26 Kalyaan & Desch we also have explored higher and lower values. In Case II we investigate a profile of α(r) motivated by simula- tions of how the MRI would operate in protoplanetary disks, with higher values of α close to the star, where temperatures are hot, and far from the star, where low densities permit high ionization fractions, but low at in- termediate radii, where the disk is dominated by MRI "dead zones". This α(r) profile varies between 10−4 and 10−3, and is on average close to 3 × 10−4. This profile is an approximation of the results of Kalyaan et al. (2015). In Case III, we adopt an α(r) profile similar to that of Desch et al. (2018) constructed to explain a variety of meteoritic data. This profile is marked by low α = 1 × 10−4 throughout the disk, but rising at inter- mediate radii to a an elevated value α = 1 × 10−3 in the inner disk. As in Case II, this profile varies between 10−4 and 10−3 and is on average close to α = 3 × 10−4. This profile is consistent with hydrodynamic instabili- ties like vertical shear instability acting throughout the disk, augmented by magnetic disk winds in the inner disk. With these α(r) profiles as inputs, we conducted 1-D disk simulations including the transport of vapor and small particles by advection and diffusion, plus trans- port of chondrule-sized (1 mm) particles by advection, diffusion, and radial drift due to aerodynamic drag. We included the condensation of vapor to ice on chondrule surfaces and the sublimation of ice to vapor. We also ac- counted for growth of planetesimals from the population of chondrules or icy chondrules, predicting the ice frac- tions of these asteroidal bodies formed at each distance from the star. 5.2. Important Conclusions We present the following important results from this study: 1. Radial water distribution in the protoplanetary disk is sensitive to the subtly-interconnected roles of the various physical processes of advection, dif- fusion and drift of vapor and solids. However, fac- tors that affect drift play the most important role in setting abundances of vapor and ice inside and outside of the snow line, in disk gas and solids, and eventually planetesimals. The rates of drift deter- mine how much ice mass is carried into the inner nebula, half of which is inevitably moved outward via vapor diffusion through the snow line, lead- ing to high water-rock ratios for chondrules here. If chondrule diffusion is efficient beyond the snow line, these icy chondrules are less likely to drift back towards the snow line, and are accreted into asteroids that remain there. 2. Disks with lower (global) α show higher peak abundances of water in asteroids right beyond the snow line, with a narrower peak width. This trans- lates to sharper volatile gradients in asteroids that may form at different distances. Slight enhance- ments and depletions in the radial abundance pro- file are maintained for a longer duration in less vis- cous disks, which may provide unique localized en- vironments for interesting chemistry. Vapor abun- dance in the inner disk also stays relatively high. 3. Disks with higher α lead to efficient depletion of water vapor in the inner disk over few Myr timescales. Due to enhanced diffusivity of chon- drules beyond the snow line, a shallower peak abundance with a broad peak width results, and the snow line sweeps a greater distance with time. Such disks evolve more rapidly, leaving less time for planets to accrete mass, unlike less viscid disks and even radially varying α disks that have less turbulent regions. 4. While the early location of the snow line is de- pendent on α, the final location (at 5 Myr) is not sensitive to the choice of the globally uniform α. 5. Ultimately, for uniform α disks, the order of im- portance of the various factors determining the magnitude of the peak ice fraction in chondrules beyond the snow line is as follows: achon > tgrowth > α > Sc > κ, i.e., planetary proper- ties are more important than disk properties. The order of importance of factors affecting the peak ice fraction in asteroids beyond the snow line is as follows: α > tgrowth > achon > everything else. Here, disk and planetary parameters seem to be as important as the other. 6. The inner disk water vapor content is sensitive to the size of the drifting solids that brings water in- ward from the outer nebula. Even a slight increase in particle size by a factor of 2 (i.e., achon = 0.03 to 0.06 cm) produces a persistent enhancement in water vapor in the inner disk interior to the snow line for at least ∼ 0.5 Myr. 7. Assuming a more rapid growth timescale of aster- oids (i.e., 0.3 Myr; comparable to that of chondrule and vapor diffusion) results in the depletion of the chondrule pool beyond the snow line before they diffus inward, as well as back diffusion of the vapor inward of the snow line diffusing outward, thus af- fecting the subtle 'equilibrium' achieved by these diffusive processes. This causes the snow line itself to slightly move outward. 8. Radially varying α disks yield disk structure dis- tinctly different from the smooth uniform α disk. Snow Lines and Different α Profiles 27 An MRI-active disk may have a tenuous inner disk with a mass pile up at distances of a few AU. On the contrary, the hybrid α case with the more tur- bulent inner disk may have a flat (and low) Σ(r) upto 20 AU within 0.5 Myr. These effects conse- quentially lead to different radial distribution of volatiles in each disk. 9. In CASE II (MRI-α disk), much of the bulk gas (along with the vapor and chondrules) collects at around 1 AU. The snow line also does not move much with time. Both of these processes have the effect of producing large ice-rock ratios of chon- drules beyond the snow line. 10. For the hybrid α disk, higher α in the inner disk leads to high accretion rates, quick depletion of the disk mass, and therefore drastic drop in the temperature of the inner disk. This leads to sig- nificant movement of the water snow line that does not allow for abundances of ice to reach locally as high values as in CASE II. 11. We argue that CASE III appears most like the solar nebula, and yields asteroids with a diverse population of differing water content as well as a monotonically increasing amount of water content. This scenario is also consistent with the forma- tion of enstatite chondrites at < 2 AU, ordinary chondrites at 2-4 AU region (after a few Myr) and carbonaceous chondrites beyond 3-4 AU (Desch et al. 2018). This suggests that we might be able to constrain α(r) from the radial distribution of asteroidal water content. 12. In summary, we argue that Σ(r), radial water abundance NH2O/NH2 as well as radial distribu- tion of asteroidal water content provide additional avenues for constraining α(r) and therefore the mechanism of disk evolution. The authors are grateful for helpful discussions with Jacob Simon, Kevin Flaherty, Cornelis Dullemond and Prajkta Mane. The authors are also grateful for the insightful suggestions from the anonymous reviewer. REFERENCES Alexander, C. M. O., Barber, D. J., & Hutchison, R. 1989, Brauer, F., Dullemond, C. P., & Henning, T. 2008, A&A, 480, Geochim. Cosmochim. Acta, 53, 3045 859 Alexander, C. M. O. '., Howard, K. T., Bowden, R., & Fogel, Brown, M. E. 2012, Annual Review of Earth and Planetary M. L. 2013, Geochim. Cosmochim. Acta, 123, 244 Sciences, 40, 467 ALMA Partnership, Brogan, C. L., P´erez, L. M., et al. 2015, ApJL, 808, L3 Andrews, S. M., Wilner, D. J., Hughes, A. M., Qi, C., & Dullemond, C. P. 2009, ApJ, 700, 1502 Andrews, S. M., Wilner, D. J., Hughes, A. M., Qi, C., & Dullemond, C. P. 2010, ApJ, 723, 1241 Ansdell, M., Williams, J. P., Manara, C. F., et al. 2017, AJ, 153, 240 Chapman, C. R., & Salisbury, J. W. 1973, Icarus, 19, 507 Chiang, E. I., & Goldreich, P. 1997, ApJ, 490, 368 Ciesla, F. J., & Cuzzi, J. N. 2006, Icarus, 181, 178 Cieza, L. A., Casassus, S., Tobin, J., et al. 2016, Nature, 535, 258 Clarke, C. J., & Pringle, J. E. 1988, MNRAS, 235, 365 Cuzzi, J. N., & Zahnle, K. J. 2004, ApJ, 614, 490 Cuzzi, J. N., Hogan, R. C., Paque, J. M., & Dobrovolskis, A. R. 2001, ApJ, 546, 496 Ansdell, M., Williams, J. P., Trapman, L., et al. 2018, ApJ, 859, Cuzzi, J. N., Davis, S. S., & Dobrovolskis, A. R. 2003, Icarus, 21 166, 385 Ataiee, S., Baruteau, C., Alibert, Y., & Benz, W. 2018, arXiv:1804.00924 Bai, X.-N., & Stone, J. M. 2011, ApJ, 736, 144 Bai, X.-N., & Stone, J. M. 2013, ApJ, 769, 76 Bai, X.-N., Ye, J., Goodman, J., & Yuan, F. 2016, ApJ, 818, 152 Bai, X.-N. 2013, Protostars and Planets VI Posters, 2 Bai, X.-N. 2014, ApJ, 791, 137 Bai, X.-N. 2015, ApJ, 798, 84 Bai, X.-N. 2016, ApJ, 821, 80 Balbus, S. A., & Hawley, J. F. 1991, ApJ, 376, 214 Baraffe, I., Chabrier, G., Allard, F., & Hauschildt, P. H. 2002, A&A, 382, 563 Bell, J. F., Davis, D. R., Hartmann, W. K., & Gaffey, M. J. Dauphas, N. 2017, Nature, 541, 521 Davis, S. S. 2005, ApJ, 620, 994 DeMeo, F. E., & Carry, B. 2014, Nature, 505, 629 Desch, S. J. 2007, ApJ, 671, 878 Desch, S. J., Estrada, P. R., Kalyaan, A., & Cuzzi, J. N. 2017, ApJ, 840, 86 Desch, S. J., Kalyaan, A., & O'D. Alexander, C. M. 2018, ApJS, 238, 11 Desch, S. J., & Turner, N. J. 2015, ApJ, 811, 156 Dodson-Robinson, S. E., Willacy, K., Bodenheimer, P., Turner, N. J., & Beichman, C. A. 2009, Icarus, 200, 672 Dullemond, C. P., & Dominik, C. 2004, A&A, 421, 1075 Dullemond, C. P., Birnstiel, T., Huang, J., et al. 2018, ApJL, 1989, Asteroids II, 921 869, L46 Birnstiel, T., Ormel, C. W., & Dullemond, C. P. 2011, A&A, Estrada, P. R., Cuzzi, J. N., & Morgan, D. A. 2016, ApJ, 818, 525, A11 200 Birnstiel, T., Andrews, S. M., Pinilla, P., & Kama, M. 2015, Flaherty, K. M., Hughes, A. M., Rosenfeld, K. A., et al. 2015, ApJL, 813, L14 ApJ, 813, 99 Bockel´ee-Morvan, D., Gautier, D., Hersant, F., Hur´e, J.-M., & Flaherty, K. M., Hughes, A. M., Rose, S. C., et al. 2017, ApJ, Robert, F. 2002, A&A, 384, 1107 843, 150 28 Kalyaan & Desch Flaherty, K. M., Hughes, A. M., Teague, R., et al. 2018, ApJ, Najita, J. R., Carr, J. S., Pontoppidan, K. M., et al. 2013, ApJ, 856, 117 766, 134 Flock, M., Nelson, R. P., Turner, N. J., et al. 2017, ApJ, 850, 131 Gail, H.-P. 2001, A&A, 378, 192 Gammie, C. F. 1996, ApJ, 457, 355 Garaud, P., & Lin, D. N. C. 2007, ApJ, 654, 606 Gomes, R., Levison, H. F., Tsiganis, K., & Morbidelli, A. 2005, Nature, 435, 466 Gradie, J., & Tedesco, E. 1982, Science, 216, 1405 Gressel, O., Turner, N. J., Nelson, R. P., & McNally, C. P. 2015, ApJ, 801, 84 Guilloteau, S., Dutrey, A., Wakelam, V., et al. 2012, A&A, 548, A70 Gundlach, B., & Blum, J. 2015, ApJ, 798, 34 Hartmann, L., Calvet, N., Gullbring, E., & D'Alessio, P. 1998, ApJ, 495, 385 Hayashi, C. 1981, Progress of Theoretical Physics Supplement, 70, 35 Hueso, R., & Guillot, T. 2005, A&A, 442, 703 Hughes, A. M., Wilner, D. J., Andrews, S. M., Qi, C., & Hogerheijde, M. R. 2011, ApJ, 727, 85 Hutchison, R., Alexander, C. M. O., & Barber, D. J. 1987, Geochim. Cosmochim. Acta, 51, 1875 Nakamura, T., Noguchi, T., Tanaka, M., et al. 2011, Science, 333, 1113 Notsu, S., Nomura, H., Ishimoto, D., et al. 2016, ApJ, 827, 113 Notsu, S., Nomura, H., Ishimoto, D., et al. 2017, ApJ, 836, 118 Notsu, S., Nomura, H., Walsh, C., et al. 2018, ApJ, 855, 62 Oberg, K. I., Murray-Clay, R., & Bergin, E. A. 2011, ApJL, 743, L16 Oka, A., Nakamoto, T., & Ida, S. 2011, ApJ, 738, 141 Pinilla, P., Flock, M., Ovelar, M. d. J., & Birnstiel, T. 2016, A&A, 596, A81 Pinte, C., Dent, W. R. F., M´enard, F., et al. 2016, ApJ, 816, 25 Pontoppidan, K. M., Salyk, C., Bergin, E. A., et al. 2014, Protostars and Planets VI, 363 Rafikov, R. R. 2017, ApJ, 837, 163 Raymond, S. N., & Izidoro, A. 2017, Icarus, 297, 134 Ros, K., & Johansen, A. 2013, A&A, 552, A137 Sasselov, D. D., & Lecar, M. 2000, ApJ, 528, 995 Schoonenberg, D., & Ormel, C. W. 2017, A&A, 602, A21 Scott, E. R. D., & Krot, A. N. 2014, Meteorites and Cosmochemical Processes, 65 Semenov, D., Henning, T., Helling, C., Ilgner, M., & Sedlmayr, E. 2003, A&A, 410, 611 Ida, S., & Guillot, T. 2016, A&A, 596, L3 Jacquet, E., Piani, L., & Weisberg, M. K. 2017, Chondrules and the Protoplanetary Disk, 1963, 2001 Shakura, N. I., & Sunyaev, R. A. 1973, A&A, 24, 337 Simon, J. B., Hughes, A. M., Flaherty, K. M., Bai, X.-N., & Armitage, P. J. 2015, ApJ, 808, 180 Johansen, A., Oishi, J. S., Mac Low, M.-M., et al. 2007, Nature, Simon, J. B., Bai, X.-N., Flaherty, K. M., & Hughes, A. M. 2017, 448, 1022 Kalyaan, A., Desch, S. J., & Monga, N. 2015, ApJ, 815, 112 Klahr, H., & Hubbard, A. 2014, ApJ, 788, 21 Kretke, K. A., & Lin, D. N. C. 2007, ApJL, 664, L55 Kunz, M. W., & Balbus, S. A. 2004, MNRAS, 348, 355 Lambrechts, M., & Johansen, A. 2014, A&A, 572, A107 Lodders, K. 2003, ApJ, 591, 1220 Lynden-Bell, D., & Pringle, J. E. 1974, MNRAS, 168, 603 Lyra, W. 2014, ApJ, 789, 77 Lyra, W., & Umurhan, O. 2018, arXiv:1808.08681 Malygin, M. G., Klahr, H., Semenov, D., Henning, T., & Dullemond, C. P. 2017, A&A, 605, A30 Marcus, P. S., Pei, S., Jiang, C.-H., & Hassanzadeh, P. 2013, Physical Review Letters, 111, 084501 Marcus, P. S., Pei, S., Jiang, C.-H., et al. 2015, ApJ, 808, 87 Marti, J., & Mauersberger, K. 1993, Geophys. Res. Lett., 20, 363 Mauersberger, K., & Krankowsky, D. 2003, Geophys. Res. Lett., 30, 1121 Meijerink, R., Pontoppidan, K. M., Blake, G. A., Poelman, D. R., & Dullemond, C. P. 2009, ApJ, 704, 1471 arXiv:1711.04770 Stevenson, D. J., & Lunine, J. I. 1988, Icarus, 75, 146 Stoll, M. H. R., & Kley, W. 2014, A&A, 572, A77 Suzuki, T. K., Ogihara, M., Morbidelli, A., Crida, A., & Guillot, T. 2016, A&A, 596, A74 Takeuchi, T., & Lin, D. N. C. 2002, ApJ, 581, 1344 Teague, R., Guilloteau, S., Semenov, D., et al. 2016, A&A, 592, A49 Tsiganis, K., Gomes, R., Morbidelli, A., & Levison, H. F. 2005, Nature, 435, 459 Turner, N. J., Fromang, S., Gammie, C., et al. 2014, Protostars and Planets VI, 411 Walsh, K. J., Morbidelli, A., Raymond, S. N., O'Brien, D. P., & Mandell, A. M. 2011, Nature, 475, 206 Wanke, H., & Dreibus, G. 1994, Philosophical Transactions of the Royal Society of London Series A, 349, 285 Weidenschilling, S. J. 1977, MNRAS, 180, 57 Wu, J., Desch, S. J., Schaefer, L., et al. 2018, Journal of Geophysical Research (Planets), 123, 2691. Xu, Z., Bai, X.-N., & Murray-Clay, R. A. 2017, ApJ, 847, 52 Yang, L., & Ciesla, F. J. 2012, Meteoritics and Planetary Min, M., Dullemond, C. P., Kama, M., & Dominik, C. 2011, Science, 47, 99 Icarus, 212, 416 Morbidelli, A., Levison, H. F., Tsiganis, K., & Gomes, R. 2005, Nature, 435, 462 Najita, J. R., & ´Ad´amkovics, M. 2017, ApJ, 847, 6 Youdin, A. N., & Lithwick, Y. 2007, Icarus, 192, 588 Zhang, K., Pontoppidan, K. M., Salyk, C., & Blake, G. A. 2013, ApJ, 766, 82
1808.08870
1
1808
2018-08-27T14:57:06
Role of gaseous giants in the dynamical evolution of terrestrial planets and water delivery in the habitable zone
[ "astro-ph.EP" ]
In the present research, we study the effects of a single giant planet in the dynamical evolution of water-rich embryos and planetesimals, located beyond the snow line of systems around Sun-like stars, in order to determine what kind of terrestrial-like planets could be formed in the habitable zone (hereafter HZ) of these systems. To do this, we carry out N-body simulations of planetary accretion, considering that the gas has been already dissipated from the disk and a single giant planet has been formed beyond the snow line of the system, at 3 au. We find that a giant planet with a value of mass between Saturn-mass and Jupiter-mass, represents a limit from which the amount of water-rich embryos that moves inward from beyond the snow line starts to decrease. From this, our research suggests that giant planets more massive than one Jupiter-mass become efficient dynamical barriers to inward-migrating water-rich embryos. Moreover, we infer that the number of these embryos that survive in the HZ significantly decreases for systems that host a giant planet more massive than one Jupiter-mass. This result has important consequences concerning the formation of terrestrial-like planets in the HZ with very high water contents and could provide a selection criteria in the search of potentially habitable exoplanets in systems that host a gaseous giant around solar-type stars.
astro-ph.EP
astro-ph
MNRAS 000, 1 -- 10 (2018) Preprint 28 August 2018 Compiled using MNRAS LATEX style file v3.0 Role of gaseous giants in the dynamical evolution of terrestrial planets and water delivery in the habitable zone Mariana B. Sanchez1,2,(cid:63) Gonzalo C. de El´ıa1,2, and Luciano A. Darriba1,2 1Facultad de Ciencias Astron´omicas y Geof´ısicas, Universidad Nacional de La Plata Paseo del Bosque s/n, La Plata, B1900FWA, Buenos Aires, Argentina 2Instituto de Astrof´ısica de La Plata, CCT La Plata-CONICET-UNLP Paseo del Bosque s/n, La Plata, B1900FWA, Buenos Aires, Argentina Accepted XXX. Received YYY; in original form ZZZ ABSTRACT In the present research, we study the effects of a single giant planet in the dynami- cal evolution of water-rich embryos and planetesimals, located beyond the snow line of systems around Sun-like stars, in order to determine what kind of terrestrial-like planets could be formed in the habitable zone (hereafter HZ) of these systems. To do this, we carry out N-body simulations of planetary accretion, considering that the gas has been already dissipated from the disk and a single giant planet has been formed beyond the snow line of the system, at 3 au. We find that a giant planet with a value of mass between Saturn-mass and Jupiter-mass, represents a limit from which the amount of water-rich embryos that moves inward from beyond the snow line starts to decrease. From this, our research suggests that giant planets more massive than one Jupiter-mass become efficient dynamical barriers to inward-migrating water-rich embryos. Moreover, we infer that the number of these embryos that survive in the HZ significantly decreases for systems that host a giant planet more massive than one Jupiter-mass. This result has important consequences concerning the formation of terrestrial-like planets in the HZ with very high water contents and could provide a selection criteria in the search of potentially habitable exoplanets in systems that host a gaseous giant around solar-type stars. Key words: giant planets -- terrestrial planets: dynamic evolution and habitability 1 INTRODUCTION At the present time, we know that planetary systems are common in the Universe. They could be found around different stars and be composed by all kind of planets with a huge variety of parameters. Up to date, there are 3824 confirmed exoplanets and 2859 planetary systems (http://exoplanet.eu/), and more objects are waiting to be confirmed. As the years pass, the observational techniques improve and the theoretical models become more refined. In fact, observational studies as Cumming et al. (2008) and Howard (2013), and theoretical works as Mordasini et al. (2009), Ida et al. (2013), and Ronco et al. (2017) have shown the existence of a wide diversity of planetary systems, sug- gesting that those systems consisting only of rocky planets would seem to be the most common in the Universe. Of particular interest are the terrestrial-like planets lo- cated in the so-called "habitable zone" (hereafter HZ) of a given system, which is defined as the circumstellar region (cid:63) E-mail: [email protected] © 2018 The Authors inside which a planet can retain liquid water on its surface. However, the location of a terrestrial-like planet in the HZ is a necessary condition but not enough to say that such a planet could host life as we know today. In fact, the mainte- nance of habitable conditions on a planet requires to satisfy other conditions, which are related to the existence of a suit- able atmosphere, organic material, the presence of a mag- netic field, plate tectonics that replenish the atmosphere of CO2, among others (Martin et al. 2006). Several authors worked with N-body simulations in or- der to describe the possible formation and evolution of a planetary system and the water delivery in the HZ in dif- ferent dynamical scenarios. In particular, many works fo- cused on the study of planetary systems that host at least one gaseous giant. For example, Raymond et al. (2004) and Raymond et al. (2006) explored the accretion process and dynamics of terrestrial planets around a Sun-like star un- der the effects of a Jovian planet in the outer disk, while Mandell et al. (2007) studied the formation of Earth-like planets during and after giant planet migration in solar- type stars, considering systems with a single migrating gi- 2 M. B. S´anchez et al. ant planet and other ones with an inner migrating gas giant and an outer non-migrating giant planet. Moreover, Ray- mond et al. (2011) studied the terrestrial-like planet forma- tion and water delivery in systems with multiple unstable gas giant planets. On the other hand, Haghighipour & Ray- mond (2007) studied the habitable planet formation consid- ering one Jovian-type planet around binary systems. They worked with a Sun-like star as the primary star and took values of 0.5 M(cid:12), 1 M(cid:12), and 1.5 M(cid:12) for the secondary star of the binary system, focusing their attention on the for- mation of Earth-like planets in the HZ of the primary star. More recently, Quintana & Lissauer (2014) worked with ter- restrial planet formation around a Sun-like star considering a massive planet in the system with values of masses be- tween 1 M⊕ and 1 Mjup, while Izidoro et al. (2015) used dynamical simulations to show that gas giant planets act as barriers to the inward migration of super-Earths initially located in distant orbits around a Sun-like star, consider- ing interactions with a gaseous protoplanetary disk in the middle stages of their formation. Lately, Zain et al. (2018) worked with planetary formation and water delivery in the HZ around a Sun-like star considering different scenarios: one with a Jovian-like giant, one with a Saturn-like giant and other ones without giant planets in the system. They found that planets with high amount of water in mass, were formed in the HZ of all their work scenarios. All these works focus their attention in the late stages of terrestrial planet formation, assumed that water was delivered to planets via collisions, considering the condensation of material beyond the snow line, located about 3 au. In the present work, we use N-body simulations in or- der to study the dynamical evolution of systems that host a massive gaseous giant just beyond the snow line around a Sun-like star, when the gas has been already dissipated of the disk. The main goal of our research is to understand how the giant planet of a system affects the formation of the terrestrial ones, in particular those potentially habitable. To do this, we propose different scenarios, considering only one giant planet per system around the snow line, whose mass ranges from 0.5 Msat to 3 Mjup, where Msat and Mjup repre- sent the planetary mass of Saturn and Jupiter, respectively. This work is structured as follows: in Section 2, we describe the model and the numerical method that we used for se- lecting the initial conditions of our work. In Section 3, we present the N-body code and specify the physical and orbital parameters of the bodies that participate of the numerical simulations. In Section 4, we show the HZ model that we used in order to classify the potentially habitable planets. In Section 5, we expose the results obtained from the N- body simulations. At the end, we give the conclusions of the present research in Section 6. 2 MODEL AND NUMERICAL METHOD In this section, we describe the model used for the proto- planetary disk together with the parameters chosen for gen- erating the initial conditions of our numerical simulations. From these initial conditions, we calculate the distribution of embryos, planetesimals, and the location of the giant planet at the beginning of the post gas phase of each system, in order to carry out the N-body simulations on this last stage of formation of a planetary system. 2.1 Model of the protoplanetary disk The parameter that determines the distribution of the ma- terial in a protoplanetary disk is the surface density. The gas-surface density profile Σg(R) and the solid-surface den- sity profile Σs(R) that we adopted for our model of proto- planetary disk are given by Σg(R) = Σ0g e−(R/Rc)2−γ (cid:18) R (1) , (cid:19)−γ (cid:18) R Rc Rc (cid:19)−γ Σs(R) = Σ0sηice e−(R/Rc)2−γ , (2) (Lynden-Bell & Pringle 1974; Hartmann et al. 1998) where R is the radial coordinate in the mid plane of the protoplan- etary disk, Rc the characteristic radius of the disk, and γ the factor which determines the density gradient. Moreover, the parameter ηice represents an increase in the amount of solid material due to the condensation of water beyond the snow line Rice. The normalization constant Σ0g is determined assuming axial symmetry for the material of the disk. Un- der such conditions, we can express the protoplanetary disk mass by Þ ∞ 0 Md = 2πRΣg(R)dR, from which we obtain Σ0g = (2 − γ) Md 2πR2 c . (3) (4) In order to determine Σ0s we used the relation between the gas and solid surface densities given by Lodders et al. (2009). As we are considering a 1 M(cid:12) star with solar metallicity, this relation is confined to Σ0s = z0Σ0g, where z0 is the primordial abundance of heavy elements in the Sun and has a value of z0 = 0.0153. It is worth remarking that the systems of work that we propose in the present research host a giant planet around the snow line at the end of the gaseous phase. Thus, we must select appropriate parameters concerning the disk mass and the gas and solid density profiles that lead to the forma- tion of such work scenarios. To do this and following the study developed by Zain et al. (2018), we adopt a disk mass Md = 0.1 M(cid:12), which gives the amount of material neces- sary around the snow line to form a gaseous giant planet in the range of masses we are working with. Despite of the existence of observed less massive protoplanetary disks, An- drews et al. (2010) found evidence of more massive proto- planetary disks in the 1 Myr-old Ophiuchus star-forming re- gion ≥ 10% of the stellar-mass, such as Elias 24 and DoAr 25 with 11.7% and 13.6% of the stellar-mass respectively. We used a characteristic radius Rc = 25 au, and an exponent γ = 0.9, which are consistent with the observations carried by Andrews et al. (2010). According to Lodders et al. (2009), we assume ηice = 0.5 if R < Rice and ηice = 1 if R > Rice, being Rice = 2.7 au for a solar luminosity star (Ida & Lin 2004). Furthermore, we assume that the protoplanetary disk presents a radial compositional gradient. In fact, we consider that bodies beyond Rice present a water content of 50% by MNRAS 000, 1 -- 10 (2018) mass, while bodies inside Rice do not have water. This wa- ter distribution is assigned to each body in our simulations based on its initial location. 2.2 Post gas stage: initial distributions From the surface density profiles specified in the previous section, we determine the initial position of the giant planet and the embryo and planetesimal distributions in the post gas stage of the system. We remark that the region of study of the present research is confined between 0.5 au ≤ R ≤ 9.5 au, and it includes the HZ of the system, the snow line, and an outer region with water-rich embryos and planetesimals. To give a better understanding of the different body distri- butions, we divide the region of work in the following three parts: (i) Inner region, with 0.5 au ≤ R < 2.5 au, (ii) Central region, with 2.5 au ≤ R < 3.5 au, (iii) Outer region, with 3.5 au ≤ R ≤ 9.5 au. In the following, we describe the considerations adopted to determine the distributions of the different bodies in each of these three regions. 2.2.1 Inner region In the inner region of the system, we consider only the ex- istence of planetary embryos since we assume that all plan- etesimals were already accreted by embryos in the previous stages such as it was shown by Zain et al. (2018). For a suit- able inner embryo distribution, we express the mass of an embryo located at R growing in the oligarchic growth mode by M = 2πR∆RHΣs(R) f , (5) (Kokubo & Ida 2000) where Σs(R) is the solid-surface den- sity, f a factor that represents the planetesimal fraction ac- creted by the embryo, and ∆RH the orbital separation be- tween two consecutive embryos of mass M in terms of their mutual Hill radii, which is given by RH = R , (6) being M(cid:63) the mass of the central star. Replacing Eqs. (2) and (6) in Eq. (5), we obtain an expression for the mass of each embryo as a function of the distance R, which is given by (cid:18) 2 (cid:19) 1 3(cid:18) R (cid:19)−γ 3M(cid:63) Rc (cid:17)2−γ(cid:33) 3 2 −(cid:16) R Rc e M = 2πR2∆Σ0sηice f In the inner system, we assume that f = 1, since it is as- sumed that all planetesimals were accreted by the inner em- bryos at the end of the gaseous phase (Zain et al. 2018). For the initial mass of the first embryo, which is located at R0 = 0.5 au, we derive a value of M0 = 0.11 M⊕. Assuming a ∆ = 5, we calculate the initial locations and masses for the rest of the inner embryos by the expressions (cid:19) 1 3 (cid:18) 2M 3M(cid:63) (cid:32) (cid:19) 1 3 (cid:18) 2Mi 3M(cid:63) MNRAS 000, 1 -- 10 (2018) Ri+1 = Ri + ∆Ri , (8) (cid:32) Mi+1 = A (cid:18) 2 3M(cid:63) (cid:19) 1 3(cid:18) Ri+1 (cid:19)−γ Rc −(cid:16) Ri+1 Rc e (cid:17)2−γ(cid:33) 3 2 , 3 (9) for i = 0, 1, . . . , being A = 2πR2∆Σ0sηice f . By doing this up to a value of R < 2.5 au, we obtain a population of 37 inner embryos, which has a total mass of 16.78 M⊕. 2.2.2 Central region In the central region of the system, we only located the gi- ant planet at R = 3 au, just beyond the snow line, as we assumed that it had already accreted all the surrounding material in the previous stages. We considered that in an environment around the snow line is where the surface den- sity is maximized and the influence zone is suitable for an embryo to accrete enough mass in order to form a giant core (Brunini & Benvenuto 2008). According to a core instability model and an oligarchic growth regime of solid protoplanets for the growth of a core of a giant planet, the expected nec- essary mass is one between 10 M⊕ - 15 M⊕, (Mizuno 1980; Bodenheimer & Pollack 1986; Pollack et al. 1996), which is the amount of mass of the surrounding material, taking into account embryos and planetesimals, which were located in this region in the previous stages if we consider a disk mass of 0.1 M(cid:12) (Zain et al. 2018). Assuming also that the giant planet was formed in situ, we defined each work scenario in relation with the giant planet mass chosen by: - Scenario 1: Mgiant = 0.5Msat. - Scenario 2: Mgiant = 1Msat. - Scenario 3: Mgiant = 1Mjup. - Scenario 4: Mgiant = 1.5Mjup. - Scenario 5: Mgiant = 2Mjup. - Scenario 6: Mgiant = 3Mjup. We chose masses between 0.5Msat and 3Mjup because we wanted to study massive giant planets effects as barriers of water-rich material which moves inward the inner system and their relation with the amount of water in planets into the HZ. We chose a range in mass which includes the mass of the gaseous giant planets of our Solar System, considering an arbitrary lower limit of a half of the Saturn-Mass and an upper limit of 3 times the Jupiter-mass, to analyze the effects of little more massive giant planets than the massive one in our Solar-System (less massive planets formed around the snow line of a system which orbits around a Sun-like star have been studied in Zain et al. (2018)). In the outer region of the system, we distribute both em- bryos and planetesimals. As for the embryo distribution, we consider that the factor f is a function of the radial distance R from the central star, which is constructed in a consistent way with the outer embryo distribution observed in Zain et al. (2018). Then, we use the Eq. (8) to determine the initial locations of the outer embryos and the Eq. (9) to calculate their masses. From this, we obtain a total mass of 33.22 M⊕, which is distributed in 12 outer embryos. In Fig. 1, we can see the variation of the mass of each embryo . (7) 2.2.3 Outer region 4 M. B. S´anchez et al. Figure 1. Initial distribution of embryo masses and the giant planet of 1 Mjup as a function of the distance R from the central star. Figure 2. Solid initial density profile which models the plan- etesimals distribution at the beginning of the post gas stage as a function of the distance R. and the giant planet of 1 Mjup as a representative giant, as a function of the distance R. Regarding planetesimals, we compute the total mass contained in such a population from the difference between the total mass of solids in the outer region and the total mass contained in the outer embryo population. From this, the total mass of the planetesimal population associated to the outer region of the system is equal to 55 M⊕. To analyze the planetesimal distribution, we multiply the solid-surface density profile given by Eq. (2) for 1 − f . As we have al- is a ready mentioned, function of the radial distance R in the outer region, which is constructed in a consistent way with the outer embryo distribution observed in Zain et al. (2018). From this, we calculate the planetesimal surface density profile at the be- ginning of the post gas stage, which is shown in Fig. 2. As the reader can see, the planetesimal surface density profile is zero in the inner region of the system, since we assume that the embryos efficiently accreted all planetesimals of such a region during the gaseous phase. f = 1 in the inner region, while f According to this, we can express the differential mass of planetesimals contained in a ring centered at R with a width dR by dM = 2πRΣs(R)(1 − f )dR. If we consider that all the planetesimals have the same in- dividual mass, we can write the mass dM as a function of the individual mass of a planetesimal mp and the amount of planetesimals dN in the ring by dM = mpdN. (11) (10) Then, we sample the amount of planetesimals with a distri- bution function F(R) that is function of the distance R from the central star dN = F(R)dR. Finally, from the Eqs. (10), (11), and (12), we obtain the following expression for the distribution function F(R) F(R) = (13) This function F(R) will be used to generate the planetesimal population for developing our N-body simulations. Σs(R)(1 − f ). 2πR mp (12) 3 N-BODY SIMULATIONS To carry out our study, we use the so-called N-body code Mercury, which was developed by Chambers (1999). In our work, we make use of the hybrid integrator, which uses a second order symplectic algorithm to treat interaction be- tween objects with separations greater than 3 Hill radii and a Bulirsch-stoer method for resolving closer encounters. We integrate each simulation for 100 Myr, which is consistent with the Earth-formation timescale according to Jacobson et al. (2014). To develop the integration, we adopted a time step of 3 days, which is shorter than 1/40th of the orbital period of the innermost body in our simulation. Moreover, in order to avoid any numerical error for small-perihelion orbits, we assume a non-realistic Sun's radius of 0.1 au. Mercury evolves the orbits of embryos and planetesi- mals allowing collisions between them. These collisions are treated as perfectly inelastic ones, conserving the mass and water content of the bodies of the simulations. In order to re- duce the CPU time, we assume that embryos interact gravi- tationally with all other bodies of the simulation, while plan- etesimals are not self-interacting (Raymond et al. 2006). To make use of the Mercury code, it is necessary to give physical and orbital parameters for the giant planet, em- bryos and planetesimals of our simulations. Regarding giant planet, we change its mass in each work scenario, adopt- ing values of 0.5 Msat, 1 Msat, 1 Mjup, 1.5 Mjup, 2 Mjup and 3 Mjup, such as it was proposed in Sect. 2.2.2. For the most massive giant planets (1 Mjup, 1.5 Mjup, 2 Mjup and 3 Mjup), we consider a physical density of 1.3 g cm−3, while for the less massive giants (0.5 Msat and 1 Msat) we assume a physical density of 0.7 g cm−3. Moreover, each giant planet has an initial semi-major axis of a = 3 au. Finally, we con- sider initial quasi-circular and co planar orbits for every gi- ant planet, with values of perihelion argument ω, ascending node longitude Ω, and mean anomaly M randomly selected between 0◦ and 360◦. In all our work scenarios, we consider the same dis- tribution of masses and semi-major axis for the planetary embryos. In fact, each of them is assigned with an initial semi-major axis and an initial mass given by Eqs. (8) and (9), respectively. For all embryos, we consider a physical den- sity of 3 g cm−3. Moreover, initial eccentricities and inclina- MNRAS 000, 1 -- 10 (2018) 0.1 1 10 100 1000 0 1 2 3 4 5 6 7 8 9 10Mass (M⊕)R (au) 0 1 2 3 4 5 6 7 8 0 1 2 3 4 5 6 7 8 9Solid superficial density (g cm-2)R (au) tions lower than 0.02 and 0.5◦, respectively, are randomly assigned, while the initial angular parameters ω, Ω and M are randomly determined between 0◦ and 360◦ for each em- bryo. In all our simulations, we include 1000 planetesimals with an individual mass of 0.055 M⊕. To determine the ini- tial values of the semi-major axis of the planetesimals, we adopt the acceptance-rejection method developed by John von Neumann using the function F(R) given by Eq. (13). For all planetesimals, we consider a physical density of 1.5 g cm−3. Finally, just like for embryos, the planetesimals are randomly assigned with initial eccentricities and inclinations lower than 0.02 and 0.5◦, respectively, while ω, Ω and M are randomly given between 0◦ and 360◦. Because of the stochastic nature of the accretion pro- cess, we remark that it is very necessary to carry out a set of N-body simulations for each of our 6 work scenarios in order to analyze the results in a statistical way. Thus, for scenarios 2-6, we carry out 30 numerical simulations, while that for scenario 1 we develop only 13 simulations because they require much more CPU time. It is worth noting that the energy is conserved better than 1 part in 104 in all cases. 4 HABITABLE ZONE MODEL In the present work, we use the HZ definition given by Kopparapu et al. (2013a), Kopparapu et al. (2013b) and Kopparapu et al. (2014). Using an updated 1-D radiative- convective, cloud free model, they estimated the boundaries of the HZ around Sun-like stars. In their studies, they took new H2O and CO2 absorption coefficients, derived from the HITRAN 2008 and HITEMP 2010 line-by-line databases. In this model, they supposed that into the inner HZ, atmo- spheres are dominated by water, while in the outer HZ, are dominated by carbon dioxide. Moreover, considering these kind of atmospheres and the relation between the pressure of surrounding nitrogen and the planetary radius, they found a dependence between the planetary mass and the width of the HZ. In fact, they studied the limits of the conserva- tive HZ for planets of 0.1 M⊕, 1 M⊕ and 5 M⊕ which orbit around a Sun-like star, founding a changing internal limit of 1.005 au, 0.9504 au and 0.9174 au respectively. They also found a fix external limit of 1.676 au in all cases. We could notice that the internal limit gets closer to the star while the planetary mass increases what concludes that with a more massive planet a wider HZ is related. From the values of the internal limit in relation with the planetary mass, we did an interpolation in order to associated our planets with their limits of the HZ. In our work, we considered that a planet with its semi- major axis, aphelion and perihelion completely contained into the HZ, will be a potentially habitable planet. How- ever, in our simulations, we found some planets with high eccentricities. In case that a planet has a very eccentric orbit and its perihelion or aphelion scape from the limits of the HZ, Williams & Pollard (2002) proposed that is the average temporal flux in an orbit which determines the habitability conditions. We adopted this average flux criterion when the perihelion or aphelion of a planet is not completely contained into the limits of the HZ but near them. For a planet with eccentricity e, semi-major axis a, and assuming a Sun-like MNRAS 000, 1 -- 10 (2018) 5 central star, the average flux normalized to terrestrial flux is Seff = 1 √ 1 − e2 . a2 (14) On the other side, Kopparapu et al. (2014) calculated values of the average flux for 0.1 M⊕, 1 M⊕ and 5 M⊕ founding changing maximum average fluxes of 0.99, 1.107 and 1.188 respectively, which are normalized to terrestrial flux. They also found a fix minimum average flux of 0.356, normalized to terrestrial flux, in all cases. We could noticed that the maximum flux gets bigger while the planetary mass increases. We used those maximum average flux values in an interpolation with their planetary mass in order to calculate maximum average fluxes for our planets and related them with the orbital elements semi-major axis and eccentricity through the Eq. (14), in order to determinate flux curves of maximum average flux that allow us to express the semi-major axis as a function of the eccentricity. Using that flux criterion, we can say that a planet is considered inside the HZ if the evolution of its semi-major axis and eccentricity, in the last period of the integration, are contained into the maximum and minimum average flux curves. In this context, we assumed both criterion to consider a planet as a potentially habitable one. We will see this classification in detail in section Results. 5 RESULTS Once the work scenarios were defined, our main goal is to analyze the influence of a single giant planet located around the snow line in the dynamical evolution of each planetary system of our simulations. In particular, we study how a sin- gle giant planet affects the evolution of water-rich embryos and planetesimals located beyond the snow line, in order to understand what kind of planets could be formed in the HZ of the system and their amount of water in mass in each of our six work scenarios. 5.1 Dynamical evolution of outer embryos First of all, we study how a single giant planet located around the snow line of a system is able to affect dynam- ical evolution of water-rich embryos and planetesimals of the outer disk, which should have important implications in the type of planets formed in the HZ of the system in each work scenario. To study the evolution of the outer em- bryo population, we started analyzing the removal process of outer embryos in each scenario in order to determine the percentage of removed outer embryos as a function of the giant's mass. To do this, we calculated the number of re- moved outer embryos in every simulation for each scenario of study. Then, we computed averaged percentage adding the number of removed outer embryos in each simulation and normalizing it to the total amount of initial outer em- bryos in all simulations corresponding to each scenario. In Fig. 3, we present the averaged percentages of removed outer embryos with error bars for the six different scenarios. From this, we can observe that the percentage of removed outer 6 M. B. S´anchez et al. Figure 3. Average percentage of the amount of removed outer embryos in each work scenario with their associated error. Figure 4. Average percentages of the amount of ejected outer embryos over the total amount of the removed outer embryos in each work scenario with their associated error. embryos is an increasing function of the giant planet's mass. In fact, while a giant planet of of 1 Msat removed ∼ 40% of the outer embryos, the massive giants of 2 Mjup and 3 Mjup removed ∼ 60% and ∼ 70% of them, respectively. Once it is made, we calculated the percentage of outer embryos ejected from the system in each of our scenarios of study. In the N-body simulations, one body was assumed to be ejected if it reached a distance from the central star greater than 100 au. For each scenario, we computed the number of ejected outer embryos in every simulation. Then, we calculated the averaged percentages of ejections adding the number of ejected outer embryos in each simulation and normalizing it to the total amount of removed external em- bryos in all simulations corresponding to each scenario. In Fig. 4, we show these average percentages with their asso- ciated error bars. Our results suggest that the percentage of ejected outer embryos is an increasing function of the gi- ant's mass. However, the reader can see that the averaged percentage of ejected outer embryos does not show signifi- cant changes for giant planets with masses between 1 Mjup and 3 Mjup. The rest of the removed outer embryos that were not ejected from the system ended being accreted by other bodies of the system or hitting with the central star. According to that presented in Fig. 4, the ejection of embryos from the system is a very efficient process in those work scenarios that host a giant planet more massive than 1 Mjup. To a better understanding, we calculate the he- liocentric escape velocity of a small body located at 3 au. We realized that less massive giant planets (0.5 Msat and 1 Msat) can not disperse small bodies with greater veloc- ity than the heliocentric escape velocity at 3 au, while more massive giant planets are able to disperse small bodies with greater velocities as well, which favors their ejection from the system. We considered that our systems of study may sig- nificantly contribute to the production of free-floating rocky planets. A detailed study about it is out of the scope of the present research. An interesting analysis is that concerning the percent- age of outer embryos that moved inward and ended in the inner region with a semi-major axis a ≤ 2.5 au in each work scenario. It is worth remarking that if the planet-planet scat- tering of outer embryos takes place in a given simulation, a single outer embryo moves inward reaching the inner region with a ≤ 2.5 au. In this context, we added those migrated embryos over the total amount of simulations per scenario. In Figure 5, we showed the averaged percentage of outer embryos that moves inward with the corresponding error bars for each work scenario. It could be noticed that giant planets with masses less than 1 Mjup allow the passage of outer embryos to the inner system up to 25%, while giant planets of 2 Mjup and 3 Mjup allow less than 10% of them. Moreover, we could see that a giant planet of 1 Mjup repre- sents a limit from which the amount of outer embryos which move inward the inner system, starts to decrease. This re- sult is important because it shows the behavior of massive giant planets as barriers to the passage of water-rich ma- terial into the inner system. It might be of astrobiological interest, since it could allow us to set restrictions in observa- tions, discarding those systems hosting a giant planet more massive than a few Jupiter masses to act as a barrier, de- creasing the probabilities to find planets with water content in the inner regions of the system. In order to analyze the resulting system structure, we studied the semi-major axis distribution of the surviving em- bryos for each scenario. In Fig. 6 we represent the cumulative percentage of those surviving embryos as a function of the semi-major axis for three different scenarios: 3 Mjup, 1 Mjup and 1 Msat. We note from the figure that systems with less massive giant planets present the most extended systems. In fact, on the one hand, systems with a giant of 3 Mjup presents more than 95% of embryos inside a ≤ 30 au, while systems with a giant of 1 Msat have less than 80% of embryos up to that distance, reaching values of a = 50 au. From the study of the evolution of planetesimals, we could say that in all of our scenarios, more than 90% of them were removed. Furthermore, we found that most of the removed ones were ejected of the system leaving a small amount of available planetesimals to be accreted by other bodies of the system. 5.2 Survival of planets in the HZ After studying how a single giant planet located around the snow line affects the embryo evolution in a given system once the gas has dissipated, we analyze how many surviving em- MNRAS 000, 1 -- 10 (2018) 0 20 40 60 80 100 0 0.5 1 1.5 2 2.5 3Removed external embryos (%)Giant planet mass (Mjup) 0 20 40 60 80 100 0 0.5 1 1.5 2 2.5 3Ejected external embryos (%)Giant planet mass (Mjup) 7 represents the conservative HZ, which is locked in between constant perihelion and aphelion curves. The change bars of the semi-major axis and eccentricity allow us to infer that the planet's orbit is not fully contained in the conservative HZ since the perihelion distance is slightly smaller than it is the inner limit during almost the whole integration. How- ever, planet's orbit is fully contained within the region de- limited by the minimum and maximum allowed stellar flux curves (thicker black curves). According to this, the planet under study is considered habitable in the present research from the averaged flux criterion. Taking into account all the simulations made for all our work scenarios, we found that 37% of the potentially habitable planets have their orbits completely contained into the limits of the HZ, while the remaining 63% have been selected by the averaged flux cri- terion. We must mention that all of those planets presents eccentricities e < 0.6, in order to apply the averaged flux cri- terion on them. A planet with an eccentricity higher than 0.6 cannot sustain surface liquid water during the whole orbital period (Bolmont et al. 2016). It is worth remarking that the six work scenarios of the present research produced planets in the HZ. In each scenario, some of the N-body simulations formed only one HZ planet, while the other ones did not produce anyone. To get a better understanding about the formation efficiency of potentially habitable planets in each work scenario, we com- puted the averaged percentages of HZ terrestrial-like planets over the total of simulations carried out per scenario, which can be seen in the top panel of Fig. 8. From this, the percent- age of potentially habitable planets formed in the six work scenarios ranges from 25% to 75%. Notice that the scenario 4 represents the most efficient one, producing 20 potentially habitable planets over 30 N-body simulations. In general terms, our simulations produced two different classes of planets in the HZ: dry worlds and water worlds. On the one hand, the dry worlds evolved from accretion seeds1 that were located inside the snow line at the end of the gas phase and did not accrete any water during their collisional history. On the other hand, the water worlds survived in the HZ, which evolved from accretion seeds initially located in a water-rich region beyond the snow line, present between 28% and 50% of water by mass. It is worth noting that 12% of water worlds that survived in the HZ evolved from accretion seeds initially located inside the snow line but received an impact from a water-rich outer embryo, which allows us to understand their very high water contents at the end of the simulations, between 22% and 28% of water in mass. The percentage of dry and water worlds formed in each of our work scenarios is illustrated in the middle and bot- tom panels of Fig. 8, respectively. From this, it is possible to observe that the percentage of dry (water) worlds that sur- vive in the HZ in each work scenario increases (decreases) as a function of the giant planet's mass. In this context, the scenarios 5 and 6, represent extreme cases since all planets that survived in the HZ at the end of the integration are dry worlds. In fact, these scenarios did not form water worlds in the total sample of simulations. The scenarios 1 and 2 also represent particular cases. 1 Following Raymond et al. (2009), we define a planet accretion seed as the largest embryo involved in its collisional history. Figure 5. Average percentages of the amount of migrated outer embryos in each work scenario with their associated error. Figure 6. Cumulative percentages of the amount of surviving embryos in scenarios 2, 3 and 6, as a function of the semi-major axis. bryos became planets in the HZ. In particular, we are inter- ested in describing the physical properties of the planets that survive in the HZ in each system of work, focusing in their masses and water contents. To do this, we use the definition of HZ proposed in Sect. 4. As we have already mentioned, we consider that a planet is potentially habitable if its semi- major axis a, and its perihelion q and aphelion Q distances are within the limits of the HZ. However, if the planet's or- bit is not fully contained in the HZ but the perihelion or the aphelion is near the limits of such region, we use the aver- aged flux criterion proposed by Williams & Pollard (2002) in order to analyze its habitability. Figure 7 shows both of such examples. On the one hand, in the left panel we can see the temporal evolution of the aphelion distance, semi- major axis and perihelion distance of a planet from scenario 6, respectively. From the beginning, such a planet evolves showing significant changes in its semi-major axis and ec- centricity. However, after a few Myr, the planet survives evolving with its orbit fully contained within the HZ con- servative limits up to the end of the integration, which are illustrated by the horizontal black lines. On the other hand, the right panel shows the averaged values of the semi-major axis and eccentricity, together with their respective variation bars, of a given planet from scenario 6. The gray shaded area MNRAS 000, 1 -- 10 (2018) 0 10 20 30 40 50 0 0.5 1 1.5 2 2.5 3Migratory external embryos (%)Giant planet mass (Mjup) 0 20 40 60 80 100 0 10 20 30 40 50 60Surviving embryos (%)Semi-major axis (au)3 Mjup1 Mjup1 Msat 8 M. B. S´anchez et al. Figure 7. On the left, we could see, from top to bottom, the temporal evolution of aphelion Q distance, semi-major axis a, and perihelion q distance, which are completely contained into the limits of the HZ, of a planet from scenario 6. On the right, we present the averaged values of semi-major axis and eccentricity, with their associated change bars, of a given planet from scenario 6. The gray shaded area represents the conservative HZ, locked in between constant perihelion and aphelion curves. Thicker black curves represent maximum and minimum averaged flux. Figure 8. Average percentages of the total amount of planets found in the HZ in each work scenario (top panel), dry planets (middle panel) and water worlds (bottom planets). The last two computed over the total amount of planets found in the HZ. On the one hand, they did not form dry worlds in the HZ and a high percentage of the planets that survived in the HZ are water worlds. On the other hand, it is important to mention that the scenario 2 represents the only one that pro- duced planets in the HZ with small percentages of water by mass, because they received, at least, one impact from plan- etesimals rich in water, in addition to dry or water worlds. They represent the 37% of potentially habitable planets and present less than 2.5% of water by mass. Moreover, the sce- nario 1 is the only one that presents the own giant into the HZ at the end of 50% of the simulations. The remaining 50% of planets in the HZ correspond to water worlds. As we described above, in all scenarios, different kind of planets were formed in the HZ with respect to their amount of water by mass. From these results, we can also say that they differ in the value of their masses. In scenarios MNRAS 000, 1 -- 10 (2018) 0 0.5 1 1.5 2 2.5 3 1000 10000 100000 1e+06 1e+07 1e+08Q,a,q [au]Time [years]Internal limit (HZ)External limit (HZ) 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.8 1 1.2 1.4 1.6 1.8 2 2.2 2.4EccentricitySemi-major axis (au)Maximum fluxMinimum fluxHZ 0 25 50 75 100Planets in the HZ 0 25 50 75 100Planets (%)Dry planets 0 25 50 75 100 0 0.5 1 1.5 2 2.5 3Giant mass (Mjup)Worlds of water 1 and 2 no sub-Earth planet was formed. Planets in the HZ reached masses up to 8 M⊕ (without counting the giant itself which migrated in scenario 1). All the others work scenar- ios, formed such as sub-Earth planets as super-Earth plan- ets. The Scenario 3 formed the most massive super-Earth planet of 11.16 M⊕ which is also a water world. scenarios 5 and 6 only reached masses up to 5 M⊕. They could also form Earth-like planets of a mass of 0.98 M⊕, which were dry worlds. The bottom panel of Fig. 8 allows us to infer a very in- terest result concerning the role of a giant planet as dynam- ical barrier in the evolution of terrestrial-like planets and water delivery in the HZ of the system. In fact, our results suggest that a single gaseous giant of 1 Mjup located around the snow line seems to represent a limit mass above which the efficiency of formation of water worlds in the HZ signifi- cantly decreases. This result is relevant since it allows us to define a selection criterion for the search of potentially hab- itable exoplanets in systems that host a single giant planet close to the snow line around solar-type stars. 6 CONCLUSIONS AND DISCUSSIONS In the present work, we analyzed how a single giant planet located around the snow line affects the dynamical evolution of terrestrial-like planets and water delivery in the HZ after the gas dissipation in solar-type star systems. Our study showed a statistical analysis based on results obtained from N-body simulations of planetary accretion. In order to an- alyze the sensitivity of our analysis regarding the mass of giant planets, we carried out N-body simulations for six dif- ferent work scenarios, in which the mass of the giant planet was varied between 0.5 Msat and 3 Mjup. Our results suggest that a Jupiter-mass planet could represent a limit mass above which the amount of water-rich embryos that moves inward from beyond the snow line starts to decrease. From this, a giant planet more massive than one Jupiter-mass might results to be an efficient dynamical barrier to inward-migrating water-rich embryos. This result has relevant implications concerning the sur- vival of water-rich terrestrial planets in the HZ of a given system. In fact, while the six work scenarios of our research produced planets in the HZ, the percentage of dry (water) worlds that survive in the HZ increases (decreases) as a function of the giant's mass. In this context, a Jupiter-mass planet located around the snow line seems to represent a limit mass above which the number of water worlds in the HZ significantly decreases. In fact, those scenarios that host a perturbing of 2 Mjup and 3 Mjup around the snow line rep- resent extreme cases, which did not produce water worlds in the HZ in any simulation. It is important to remark that the results previously described should be interpreted in the context of the nu- merical model used to carry out the N-body simulations. In fact, the MERCURY code used in the present study treats all collisions as inelastic mergers, which conserve the total mass and the water content of the interacting bodies. Thus, the masses and water contents of all planets formed in HZ should be interpreted as upper limits. Recent investigations based on hydro dynamical sim- ulations have shown that collisions are not always perfect MNRAS 000, 1 -- 10 (2018) 9 mergers. In fact, studies such as those developed by Lein- hardt & Stewart (2012) and Genda et al. (2012) analyze the limits of the different collisional regimes and describe the size and velocity distribution of the post-collision bodies. Later, Chambers (2013) used the results of those works to carry out N-body simulations of terrestrial planet formation incorpo- rating fragmentation and hit-and-run collisions. In such a work, the author compared those N-body simulations with other ones previously developed assuming all collisions as perfect mergers. The general results derived by Chambers (2013) suggested that the final planetary systems produced in the two numerical models were similar. However, the au- thor observed that planets that result in a given system have somewhat smaller masses and eccentricities when a more re- alistic treatment is included in the model. Recently, Quin- tana et al. (2016) studied giant impacts on Earth-like planets in the last stage of the evolution of a planetary system, us- ing N-body simulations, which included fragmentation and hit-and-run collisions. On the other hand, Dvorak et al. (2015) developed hy- drodynamic simulations to infer the amount of water in frag- ments after a collision for different velocities and impact angles. They found that most of the water is retained by the survivor body for impact angles α <(cid:39) 20◦ and velocities ν <(cid:39) 1.3νesc, with νesc their escape velocity. As a last work, Mustill et al. (2017) explored the effects of implementing a more realistic collision treatment on in-situ formation of planets which radial distances of few tenth of au. Taking those results into account, we consider that is important to include a more realistic treatment of the collisions and the evolution of water in the N-body code, in order to refine our percentages of water in the final potentially habitable plan- ets found in all the work scenarios and verify if the dry plan- ets that we found were totally dry or if they could present a small amount of water in mass. One last thing to take into account, is the fact that we fix the snow line in 2.7 au. We are aware of the evolution of the snow line with time and its profile according to a Sun-like star (Ciesla et al. 2015). However, we consider a fix snow line as a good approximation during our integration time and a distance of separation between dry and water rich material at the beginning of our simulations, as it is assumed by different authors such as Raymond et al. (2004), O'Brien et al. (2006), Raymond et al. (2009), Ronco & de El´ıa (2014), Zain et al. (2018), among others, who worked with N-body simulations in the last stage of the formation of a planetary system around a solar-mass star, once the gas has been already dissipated from the disk. Even though, we consider that it would be a good experiment to move the snow line inward as Ciesla et al. (2015) in their simulations in order to test the sensitivity of our results with respect to an inner separation between dry and water-rich material. This could have important consequences with respect to the final amount of water in mass of the resulting final planets in the HZ. However, this analysis is out of the scope of this work. We consider that the present work allows us to get a better understanding of the role of giant planets in the for- mation of terrestrial planets around a Sun-like star. We in- fer that our results could give a selection criteria for future searches of potentially habitable exoplanets. 10 M. B. S´anchez et al. ACKNOWLEDGMENTS Williams D. M., Pollard D., 2002, International Journal of Astro- This work was partially financed by CONICET through the PIP 0436/13. Moreover, the authors acknowledge the finan- cial support by FCAGLP for extensive use of its computing facilities. biology, 1, 61 Zain P. S., de El´ıa G. C., Ronco M. P., Guilera O. M., 2018, A&A, 609, A76 This paper has been typeset from a TEX/LATEX file prepared by the author. REFERENCES Andrews S. M., Wilner D. J., Hughes A. M., Qi C., Dullemond C. P., 2010, ApJ, 723, 1241 Bodenheimer P., Pollack J. B., 1986, Icarus, 67, 391 Bolmont E., Libert A.-S., Leconte J., Selsis F., 2016, A&A, 591, A106 Brunini A., Benvenuto O. G., 2008, Icarus, 194, 800 Chambers J. E., 1999, MNRAS, 304, 793 Chambers J. E., 2013, Icarus, 224, 43 Ciesla F. J., Mulders G. D., Pascucci I., Apai D., 2015, ApJ, 804, 9 Cumming A., Butler R. P., Marcy G. W., Vogt S. S., Wright J. T., Fischer D. A., 2008, PASP, 120, 531 Dvorak R., Maindl T. I., Burger C., Schafer C., Speith R., 2015, Nonlinear Phenomena in Complex Systems, Vol.18, No.3, pp. 310-325, 18, 310 Genda H., Kokubo E., Ida S., 2012, ApJ, 744, 137 Haghighipour N., Raymond S. N., 2007, ApJ, 666, 436 Hartmann L., Calvet N., Gullbring E., D'Alessio P., 1998, ApJ, 495, 385 Howard A. W., 2013, Science, 340, 572 Ida S., Lin D. N. C., 2004, ApJ, 604, 388 Ida S., Lin D. N. C., Nagasawa M., 2013, ApJ, 775, 42 Izidoro A., Raymond S. N., Morbidelli A., Hersant F., Pierens A., 2015, ApJ, 800, L22 Jacobson S. A., Morbidelli A., Raymond S. N., O'Brien D. P., Walsh K. J., Rubie D. C., 2014, Nature, 508, 84 Kokubo E., Ida S., 2000, Icarus, 143, 15 Kopparapu R. K., et al., 2013a, ApJ, 765, 131 Kopparapu R. K., et al., 2013b, ApJ, 770, 82 Kopparapu R. K., Ramirez R. M., SchottelKotte J., Kasting J. F., Domagal-Goldman S., Eymet V., 2014, ApJ, 787, L29 Leinhardt Z. M., Stewart S. T., 2012, ApJ, 745, 79 Lodders K., Palme H., Gail H.-P., 2009, Landolt Bornstein, Lynden-Bell D., Pringle J. E., 1974, MNRAS, 168, 603 Mandell A. M., Raymond S. N., Sigurdsson S., 2007, ApJ, 660, 823 Martin H., Albar`ede F., Claeys P., Gargaud M., Marty B., Mor- bidelli A., Pinti D. L., 2006, Earth Moon and Planets, 98, 97 Mizuno H., 1980, Progress of Theoretical Physics, 64, 544 Mordasini C., Alibert Y., Benz W., 2009, A&A, 501, 1139 Mustill A. J., Davies M. B., Johansen A., 2017, preprint, (arXiv:1708.08939) O'Brien D. P., Morbidelli A., Levison H. F., 2006, Icarus, 184, 39 Pollack J. B., Hubickyj O., Bodenheimer P., Lissauer J. J., Podolak M., Greenzweig Y., 1996, Icarus, 124, 62 Quintana E. V., Lissauer J. J., 2014, ApJ, 786, 33 Quintana E. V., Barclay T., Borucki W. J., Rowe J. F., Chambers J. E., 2016, ApJ, 821, 126 Raymond S. N., Quinn T., Lunine J. I., 2004, Icarus, 168, 1 Raymond S. N., Quinn T., Lunine J. I., 2006, Icarus, 183, 265 Raymond S. N., O'Brien D. P., Morbidelli A., Kaib N. A., 2009, Icarus, 203, 644 Raymond S. N., et al., 2011, A&A, 530, A62 Ronco M. P., de El´ıa G. C., 2014, A&A, 567, A54 Ronco M. P., Guilera O. M., de El´ıa G. C., 2017, MNRAS, 471, 2753 MNRAS 000, 1 -- 10 (2018)
1107.3643
2
1107
2011-08-16T13:22:13
High precision astrometry mission for the detection and characterization of nearby habitable planetary systems with the Nearby Earth Astrometric Telescope (NEAT)
[ "astro-ph.EP", "astro-ph.IM" ]
(abridged) A complete census of planetary systems around a volume-limited sample of solar-type stars (FGK dwarfs) in the Solar neighborhood with uniform sensitivity down to Earth-mass planets within their Habitable Zones out to several AUs would be a major milestone in extrasolar planets astrophysics. This fundamental goal can be achieved with a mission concept such as NEAT - the Nearby Earth Astrometric Telescope. NEAT is designed to carry out space-borne extremely-high-precision astrometric measurements sufficient to detect dynamical effects due to orbiting planets of mass even lower than Earth's around the nearest stars. Such a survey mission would provide the actual planetary masses and the full orbital geometry for all the components of the detected planetary systems down to the Earth-mass limit. The NEAT performance limits can be achieved by carrying out differential astrometry between the targets and a set of suitable reference stars in the field. The NEAT instrument design consists of an off-axis parabola single-mirror telescope, a detector with a large field of view made of small movable CCDs located around a fixed central CCD, and an interferometric calibration system originating from metrology fibers located at the primary mirror. The proposed mission architecture relies on the use of two satellites operating at L2 for 5 years, flying in formation and offering a capability of more than 20,000 reconfigurations (alternative option uses deployable boom). The NEAT primary science program will encompass an astrometric survey of our 200 closest F-, G- and K-type stellar neighbors, with an average of 50 visits. The remaining time might be allocated to improve the characterization of the architecture of selected planetary systems around nearby targets of specific interest (low-mass stars, young stars, etc.) discovered by Gaia, ground-based high-precision radial-velocity surveys.
astro-ph.EP
astro-ph
Noname manuscript No. (will be inserted by the editor) 1 1 0 2 g u A 6 1 . ] P E h p - o r t s a [ 2 v 3 4 6 3 . 7 0 1 1 : v i X r a High precision astrometry mission for the detection and characterization of nearby habitable planetary systems with the Nearby Earth Astrometric Telescope (NEAT) Fabien Malbet · Alain L´eger · Michael Shao · Renaud Goullioud · Pierre-Olivier Lagage · Anthony G. A. Brown · Christophe Cara · Gilles Durand · Carlos Eiroa · Philippe Feautrier · Bjorn Jakobsson · Emmanuel Hinglais · Lisa Kaltenegger · Lucas Labadie · Anne-Marie Lagrange · Jacques Laskar · Ren´e Liseau · Jonathan Lunine · Jes´us Maldonado · Manuel Mercier · Christoph Mordasini · Didier Queloz · Andreas Quirrenbach · Alessandro Sozzetti · Wesley Traub · Olivier Absil · Yann Alibert · Alexandre Humberto Andrei · Fr´ed´eric Arenou · Charles Beichman · Alain Chelli · Charles S. Cockell · Gilles Duvert · Thierry Forveille · Paulo J.V. Garcia · David Hobbs · Alberto Krone-Martins · Helmut Lammer · Nad`ege Meunier · Stefano Minardi · Andr´e Moitinho de Almeida · Nicolas Rambaux · Sean Raymond · Huub J. A. Rottgering · Johannes Sahlmann · Peter A. Schuller · Damien S´egransan · Franck Selsis · Jean Surdej · Eva Villaver · Glenn J. White · Hans Zinnecker Received: date / Accepted: date Abstract A complete census of planetary systems around a volume-limited sample of solar-type stars (FGK dwarfs) in the Solar neighborhood (d ≤ 15 pc) with uniform sen- sitivity down to Earth-mass planets within their Hab- itable Zones out to several AUs would be a major mile- stone in extrasolar planets astrophysics. This funda- mental goal can be achieved with a mission concept such as NEAT -- the Nearby Earth Astrometric Tele- scope. NEAT is designed to carry out space-borne extremely- high-precision astrometric measurements at the 0.05 µas (1σ) accuracy level, sufficient to detect dynamical ef- fects due to orbiting planets of mass even lower than Earth's around the nearest stars. Such a survey mission would provide the actual planetary masses and the full orbital geometry for all the components of the detected planetary systems down to the Earth-mass limit. The NEAT performance limits can be achieved by carrying out differential astrometry between the targets and a set of suitable reference stars in the field. The NEAT in- strument design consists of an off-axis parabola single- mirror telescope (D = 1m), a detector with a large The complete affiliations are given at the end of the paper. The full list of members of the NEAT proposal is avialable at http://neat.obs.ujf-grenoble.fr. field of view located 40 m away from the telescope and made of 8 small movable CCDs located around a fixed central CCD, and an interferometric calibration sys- tem monitoring dynamical Young's fringes originating from metrology fibers located at the primary mirror. The mission profile is driven by the fact that the two main modules of the payload, the telescope and the fo- cal plane, must be located 40 m away leading to the choice of a formation flying option as the reference mis- sion, and of a deployable boom option as an alternative choice. The proposed mission architecture relies on the use of two satellites, of about 700 kg each, operating at L2 for 5 years, flying in formation and offering a ca- pability of more than 20,000 reconfigurations. The two satellites will be launched in a stacked configuration using a Soyuz ST launch vehicle. The NEAT primary science program will encom- pass an astrometric survey of our 200 closest F-, G- and K-type stellar neighbors, with an average of 50 vis- its each distributed over the nominal mission duration. The main survey operation will use approximately 70% of the mission lifetime. The remaining 30% of NEAT observing time might be allocated, for example, to im- prove the characterization of the architecture of selected planetary systems around nearby targets of specific in- terest (low-mass stars, young stars, etc.) discovered by 2 Malbet, L´eger, Shao, Goullioud, Lagage et al. Gaia, ground-based high-precision radial-velocity sur- veys, and other programs. With its exquisite, surgical astrometric precision, NEAT holds the promise to pro- vide the first thorough census for Earth-mass planets around stars in the immediate vicinity of our Sun. Keywords Exoplanets · Planetary systems · Planetary formation · Astrometry · Space Mission 1 Introduction Exoplanet research has grown explosively in the past decade, supported by improvements in observational techniques that have led to increasingly sensitive de- tection and characterization. Among many results, we have learned that planets are common, but their phys- ical and orbital properties are much more diverse than originally thought. A lasting challenge is the detection and characteri- zation of planetary systems consisting in a mixed cortege of telluric and giant planets, with a special regard to telluric planets orbiting in the habitable zone (HZ) of Sun-like stars. The accomplishment of this goal requires the development of a new generation of facilities, due to the intrinsic difficulty of detecting Earth-like planets with existing instruments. The proposed NEAT mission has been designed to enter a new phase in exoplanetary science by delivering an enhanced capability of detect- ing small planets at and beyond 1 AU. Astrometry is probably the oldest branch of astron- omy. Greeks developed it and noticed that the posi- tion of most stars were stable in the sky, but the few that were moving became known as planets (πλάνετες στέρες = moving stars), pointing to a major differ- ence in their nature. Thanks to the precise astromet- ric measurements of planet positions by Tycho Brahe in the 16th century, Johannes Kepler established that these objects were orbiting the Sun on elliptical or- bits, expanding the Copernican revolution. After Hip- parcos, Gaia will play an important role in finding many systems with giant planets in our Galaxy. We want to extend these revolutions with the NEAT mission, namely to discover and characterize Earth-mass planets in Earth-like orbits around stars like the Sun, by cap- turing infinitesimal displacements with unprecedented accuracy. In Sect. 2, we present the science objectives of NEAT, we describe the principle of the differential astrome- try technique and we give a list of potential targets. In Sect. 3, after listing the technical challenges, we present the instrumental concept. We explain how to reach the performance and we give a summarized description of the payload, the mission and the spacecraft. In Sect. 4, we discuss both astrophysical and technical issues. Recom- mandations by the community summarized in Sect. 5 is an incentive to pursue the development of this mission in the future. 2 NEAT Science 2.1 Science objectives The prime goal of NEAT is to detect and character- ize planetary systems orbiting bright stars in the solar neighborhood that have a planetary architecture like that of our Solar System or an alternative planetary system made of Earth mass planets. It will allow the detection around nearby stars of planets equivalent to Venus, Earth, (Mars), Jupiter, and Saturn, with orbits possibly similar to those in our Solar System. It will permit to detect and characterize the orbits and the masses of many alternate configurations, e.g. where the asteroid belt is occupied by another Earth mass planet and no Jupiter. The NEAT mission will answer the fol- lowing questions: -- What are the dynamical interactions between giant and telluric planets in a large variety of systems? -- What are the detailed processes involved in planet formation as revealed by their present configura- tion? -- What are the distributions of architectures of plan- etary systems in our neighborhood up to ≈ 15 pc? -- What are the masses, and orbital parameters, of tel- luric planets that are candidates for future direct de- tection and spectroscopic characterization missions? Special emphasis will be put on planets in the Habit- able Zone because this is a region of prime interest for astrobiology. Indeed orbital parameters obtained with NEAT will allow spectroscopic follow-up observations to be scheduled precisely when the configuration is the most favorable. 2.2 High-precision differential astrometry The principle of NEAT is to measure accurately the offset angles between a target and 6-8 distant reference stars with the aim of differentially detecting the reflex motion of the target star due to the presence of its planets. An example of a field that will be observed is shown in Fig. 1 and a simulation of what will be measured is displayed in Fig. 2. The output of the analysis is a comprehensive de- termination of the mass, orbit, and ephemeris of the different planets of the multiplanetary system (namely Nearby Earth Astrometric Telescope (NEAT) 3 Fig. 2 Simulation of astrometric detection of a planet with 50 NEAT measurements (RA and DEC) over 5 yrs. Parameters are: MP = 1.5 M⊕, a = 1.16 AU , M∗ = 1 M(cid:12), D = 10 pc, SNR ≥ 6. (a) Sky plot showing the astrometric orbit (solid brown curve) and the NEAT measurements with error bars (in blue); (b) and (c) same data but shown as time series of the RA and DEC astrometric signal; (d) Separated periodogram of RA (blue line) and DEC (brown line) measurements. (e) Joint periodogram for right ascensions and declinations simultaneously. Whereas the orbit cannot be determined from the astrometric signal without the time information, its period is reliably detected in the joint periodogram (1.25 yr) with a false-alarm probability below 1% (green line). Then, the planetary mass and orbit parameters can be determined by fitting the astrometric measurements. planet orbiting around with a semi-major axis a at a distance D from the Sun is (cid:17)(cid:18) M∗ (cid:19)−1(cid:18) D (cid:19)−1 (cid:18) MP (cid:19)(cid:16) a A = 3 1 M⊕ 1 AU 1 M(cid:12) 1 pc µas. (1) To detect such a planet, one needs to reach a precision σ = A/SNR with a typical signal-to-noise ratio1 SNR = 6. If σ0 is the precision that NEAT can reach in one single observation that lasts t0 (e.g. σ0 = 0.8 µas in t0 = 1 h), when observing the same source Nvisits times during Tvisit each visit requires (cid:18) A (cid:19)−2 SNR σ0 Tvisit = t0 (2Nvisits − m)1/2 (2) Fig. 1 0.3◦ stellar field around upsilon Andromedae, a pro- posed NEAT target. There are six possible reference stars in this field marked in red (five V < 11 stars and a V = 11.1 one). for a given Nvisits, and with m = 5 + 7p parameters where p is the number of planets in the system since there are 5 parameters characterizing the star astromet- ric motion and 7 parameters for each orbit. Nvisits ≈ 50 is sufficient to solve for the parameters of 3 to 5 planets per system, for a 5-yr duration of the mission. 2.3 Targets the 7 parameters MP , P , T , e, i, ω, Ω), down to a given limit depending on the star characteristics, e.g. 0.5, 1 or 5 M⊕. The astrometric amplitude, A, of a M∗ mass star due to the reflex motion in presence of a MP mass A possible target list is shown in Table 1 where we consider the list of the nearest F, G, K stars deduced from the Hipparcos 2007 catalogue (new data reduction 1 Simulations like the ones presented in Fig. 2 show that SNR = 5.8 results in a false alarm probability of 1%. 6 (cid:115)(cid:52)(cid:40)(cid:37)(cid:45)(cid:37)(cid:41)(cid:26)(cid:52)(cid:40)(cid:37)(cid:51)(cid:37)(cid:33)(cid:50)(cid:35)(cid:40)(cid:38)(cid:47)(cid:50)(cid:40)(cid:33)(cid:34)(cid:41)(cid:52)(cid:33)(cid:34)(cid:44)(cid:37)(cid:55)(cid:47)(cid:50)(cid:44)(cid:36)(cid:51)Figure 1-2. Simulation of the astrometric detection of a planet with SIM Lite. The simulation assumes 100 measurements in RA and 100 in DEC over a mission of five years duration. The planet has a mass of 1.5 M (cid:132) orbiting at 1.16 AU from a 1.0 M (cid:132)star at a distance of 10 pc from Earth. This example was chosen to il-lustrate a system close to the limit of detectability with SIM Lite. (a) Sky plot showing the astrometric orbit (solid curve) and the individual SIM mea-surements with error bars. (b), (c) The same data as in (a) but shown as time series along with the astrometric signal projected onto RA and DEC. (d) Periodograms of the data plotted in (b) and (c). (e) Joint periodogram of data from (b) and (c). The horizontal lines in (d) and (e) show the level above which the false-alarm probability is less than 1 percent. The peak near P = 1.2 years is the astrometric signal of the 1.5 M (cid:132)planet. Note that the planet is not detected in RA or DEC alone, but is detected with a false-alarm probability of well below 1 percent in the joint periodogram.For the benchmark case of an Earth-mass planet orbiting at 1 AU around a solar-mass star at 10 pc, roughly 500 measurements at 0.82 µas accuracy are needed to attain a signal-to-noise ratio (SNR) of 5.8, the approximate threshold for detection. In general, if (cid:83) is the one-axis RMS noise per differential measurement, N is the number of SIM Lite visits, and (cid:65)THRESH is the threshold astrometric amplitude detectable with a probability of 50 percent, we have:(cid:65)THRESH = SNR × (cid:83)/ N1/2 SIM Lite offers the astrometric precision and duration (five years), along with a noise floor below 0.1 µas, to detect Earth-mass planets around the ~60 nearest FGK stars.3(cid:68)DEC, µas210 -- 1 -- 2 -- 3 -- 3 -- 2 -- 10123(cid:68)RA, µasObservedTrue(cid:68)RA, µas210 -- 1 -- 2012345Time, yearsObserved RA DisplacementTrue RA Displacement(cid:68)DEC, µas210 -- 1 -- 23Observed DEC DisplacementTrue DEC Displacement012345Time, yearsPower, µas20.40.30.20.10RA PeriodogramDEC PeriodogramDetection Threshold for 1% FAP10010Period, years100Period, years101Joint Periodogram Power, µas20.60.50.40.30.20.70.10.0Joint Periodogram, True Period 1.25 yrsDetection Threshold for 1% FAPDetected Period 1.25 yrs(b)(c)(d)(e)(a)6 (cid:115)(cid:52)(cid:40)(cid:37)(cid:45)(cid:37)(cid:41)(cid:26)(cid:52)(cid:40)(cid:37)(cid:51)(cid:37)(cid:33)(cid:50)(cid:35)(cid:40)(cid:38)(cid:47)(cid:50)(cid:40)(cid:33)(cid:34)(cid:41)(cid:52)(cid:33)(cid:34)(cid:44)(cid:37)(cid:55)(cid:47)(cid:50)(cid:44)(cid:36)(cid:51)Figure 1-2. Simulation of the astrometric detection of a planet with SIM Lite. The simulation assumes 100 measurements in RA and 100 in DEC over a mission of five years duration. The planet has a mass of 1.5 M (cid:132) orbiting at 1.16 AU from a 1.0 M (cid:132)star at a distance of 10 pc from Earth. This example was chosen to il-lustrate a system close to the limit of detectability with SIM Lite. (a) Sky plot showing the astrometric orbit (solid curve) and the individual SIM mea-surements with error bars. (b), (c) The same data as in (a) but shown as time series along with the astrometric signal projected onto RA and DEC. (d) Periodograms of the data plotted in (b) and (c). (e) Joint periodogram of data from (b) and (c). The horizontal lines in (d) and (e) show the level above which the false-alarm probability is less than 1 percent. The peak near P = 1.2 years is the astrometric signal of the 1.5 M (cid:132)planet. Note that the planet is not detected in RA or DEC alone, but is detected with a false-alarm probability of well below 1 percent in the joint periodogram.For the benchmark case of an Earth-mass planet orbiting at 1 AU around a solar-mass star at 10 pc, roughly 500 measurements at 0.82 µas accuracy are needed to attain a signal-to-noise ratio (SNR) of 5.8, the approximate threshold for detection. In general, if (cid:83) is the one-axis RMS noise per differential measurement, N is the number of SIM Lite visits, and (cid:65)THRESH is the threshold astrometric amplitude detectable with a probability of 50 percent, we have:(cid:65)THRESH = SNR × (cid:83)/ N1/2 SIM Lite offers the astrometric precision and duration (five years), along with a noise floor below 0.1 µas, to detect Earth-mass planets around the ~60 nearest FGK stars.3(cid:68)DEC, µas210 -- 1 -- 2 -- 3 -- 3 -- 2 -- 10123(cid:68)RA, µasObservedTrue(cid:68)RA, µas210 -- 1 -- 2012345Time, yearsObserved RA DisplacementTrue RA Displacement(cid:68)DEC, µas210 -- 1 -- 23Observed DEC DisplacementTrue DEC Displacement012345Time, yearsPower, µas20.40.30.20.10RA PeriodogramDEC PeriodogramDetection Threshold for 1% FAP10010Period, years100Period, years101Joint Periodogram Power, µas20.60.50.40.30.20.70.10.0Joint Periodogram, True Period 1.25 yrsDetection Threshold for 1% FAPDetected Period 1.25 yrs(b)(c)(d)(e)(a)6 (cid:115)(cid:52)(cid:40)(cid:37)(cid:45)(cid:37)(cid:41)(cid:26)(cid:52)(cid:40)(cid:37)(cid:51)(cid:37)(cid:33)(cid:50)(cid:35)(cid:40)(cid:38)(cid:47)(cid:50)(cid:40)(cid:33)(cid:34)(cid:41)(cid:52)(cid:33)(cid:34)(cid:44)(cid:37)(cid:55)(cid:47)(cid:50)(cid:44)(cid:36)(cid:51)Figure 1-2. Simulation of the astrometric detection of a planet with SIM Lite. The simulation assumes 100 measurements in RA and 100 in DEC over a mission of five years duration. The planet has a mass of 1.5 M (cid:132) orbiting at 1.16 AU from a 1.0 M (cid:132)star at a distance of 10 pc from Earth. This example was chosen to il-lustrate a system close to the limit of detectability with SIM Lite. (a) Sky plot showing the astrometric orbit (solid curve) and the individual SIM mea-surements with error bars. (b), (c) The same data as in (a) but shown as time series along with the astrometric signal projected onto RA and DEC. (d) Periodograms of the data plotted in (b) and (c). (e) Joint periodogram of data from (b) and (c). The horizontal lines in (d) and (e) show the level above which the false-alarm probability is less than 1 percent. The peak near P = 1.2 years is the astrometric signal of the 1.5 M (cid:132)planet. Note that the planet is not detected in RA or DEC alone, but is detected with a false-alarm probability of well below 1 percent in the joint periodogram.For the benchmark case of an Earth-mass planet orbiting at 1 AU around a solar-mass star at 10 pc, roughly 500 measurements at 0.82 µas accuracy are needed to attain a signal-to-noise ratio (SNR) of 5.8, the approximate threshold for detection. In general, if (cid:83) is the one-axis RMS noise per differential measurement, N is the number of SIM Lite visits, and (cid:65)THRESH is the threshold astrometric amplitude detectable with a probability of 50 percent, we have:(cid:65)THRESH = SNR × (cid:83)/ N1/2 SIM Lite offers the astrometric precision and duration (five years), along with a noise floor below 0.1 µas, to detect Earth-mass planets around the ~60 nearest FGK stars.3(cid:68)DEC, µas210 -- 1 -- 2 -- 3 -- 3 -- 2 -- 10123(cid:68)RA, µasObservedTrue(cid:68)RA, µas210 -- 1 -- 2012345Time, yearsObserved RA DisplacementTrue RA Displacement(cid:68)DEC, µas210 -- 1 -- 23Observed DEC DisplacementTrue DEC Displacement012345Time, yearsPower, µas20.40.30.20.10RA PeriodogramDEC PeriodogramDetection Threshold for 1% FAP10010Period, years100Period, years101Joint Periodogram Power, µas20.60.50.40.30.20.70.10.0Joint Periodogram, True Period 1.25 yrsDetection Threshold for 1% FAPDetected Period 1.25 yrs(b)(c)(d)(e)(a)6 (cid:115)(cid:52)(cid:40)(cid:37)(cid:45)(cid:37)(cid:41)(cid:26)(cid:52)(cid:40)(cid:37)(cid:51)(cid:37)(cid:33)(cid:50)(cid:35)(cid:40)(cid:38)(cid:47)(cid:50)(cid:40)(cid:33)(cid:34)(cid:41)(cid:52)(cid:33)(cid:34)(cid:44)(cid:37)(cid:55)(cid:47)(cid:50)(cid:44)(cid:36)(cid:51)Figure 1-2. Simulation of the astrometric detection of a planet with SIM Lite. The simulation assumes 100 measurements in RA and 100 in DEC over a mission of five years duration. The planet has a mass of 1.5 M (cid:132) orbiting at 1.16 AU from a 1.0 M (cid:132)star at a distance of 10 pc from Earth. This example was chosen to il-lustrate a system close to the limit of detectability with SIM Lite. (a) Sky plot showing the astrometric orbit (solid curve) and the individual SIM mea-surements with error bars. (b), (c) The same data as in (a) but shown as time series along with the astrometric signal projected onto RA and DEC. (d) Periodograms of the data plotted in (b) and (c). (e) Joint periodogram of data from (b) and (c). The horizontal lines in (d) and (e) show the level above which the false-alarm probability is less than 1 percent. The peak near P = 1.2 years is the astrometric signal of the 1.5 M (cid:132)planet. Note that the planet is not detected in RA or DEC alone, but is detected with a false-alarm probability of well below 1 percent in the joint periodogram.For the benchmark case of an Earth-mass planet orbiting at 1 AU around a solar-mass star at 10 pc, roughly 500 measurements at 0.82 µas accuracy are needed to attain a signal-to-noise ratio (SNR) of 5.8, the approximate threshold for detection. In general, if (cid:83) is the one-axis RMS noise per differential measurement, N is the number of SIM Lite visits, and (cid:65)THRESH is the threshold astrometric amplitude detectable with a probability of 50 percent, we have:(cid:65)THRESH = SNR × (cid:83)/ N1/2 SIM Lite offers the astrometric precision and duration (five years), along with a noise floor below 0.1 µas, to detect Earth-mass planets around the ~60 nearest FGK stars.3(cid:68)DEC, µas210 -- 1 -- 2 -- 3 -- 3 -- 2 -- 10123(cid:68)RA, µasObservedTrue(cid:68)RA, µas210 -- 1 -- 2012345Time, yearsObserved RA DisplacementTrue RA Displacement(cid:68)DEC, µas210 -- 1 -- 23Observed DEC DisplacementTrue DEC Displacement012345Time, yearsPower, µas20.40.30.20.10RA PeriodogramDEC PeriodogramDetection Threshold for 1% FAP10010Period, years100Period, years101Joint Periodogram Power, µas20.60.50.40.30.20.70.10.0Joint Periodogram, True Period 1.25 yrsDetection Threshold for 1% FAPDetected Period 1.25 yrs(b)(c)(d)(e)(a)6 (cid:115)(cid:52)(cid:40)(cid:37)(cid:45)(cid:37)(cid:41)(cid:26)(cid:52)(cid:40)(cid:37)(cid:51)(cid:37)(cid:33)(cid:50)(cid:35)(cid:40)(cid:38)(cid:47)(cid:50)(cid:40)(cid:33)(cid:34)(cid:41)(cid:52)(cid:33)(cid:34)(cid:44)(cid:37)(cid:55)(cid:47)(cid:50)(cid:44)(cid:36)(cid:51)Figure 1-2. Simulation of the astrometric detection of a planet with SIM Lite. The simulation assumes 100 measurements in RA and 100 in DEC over a mission of five years duration. The planet has a mass of 1.5 M (cid:132) orbiting at 1.16 AU from a 1.0 M (cid:132)star at a distance of 10 pc from Earth. This example was chosen to il-lustrate a system close to the limit of detectability with SIM Lite. (a) Sky plot showing the astrometric orbit (solid curve) and the individual SIM mea-surements with error bars. (b), (c) The same data as in (a) but shown as time series along with the astrometric signal projected onto RA and DEC. (d) Periodograms of the data plotted in (b) and (c). (e) Joint periodogram of data from (b) and (c). The horizontal lines in (d) and (e) show the level above which the false-alarm probability is less than 1 percent. The peak near P = 1.2 years is the astrometric signal of the 1.5 M (cid:132)planet. Note that the planet is not detected in RA or DEC alone, but is detected with a false-alarm probability of well below 1 percent in the joint periodogram.For the benchmark case of an Earth-mass planet orbiting at 1 AU around a solar-mass star at 10 pc, roughly 500 measurements at 0.82 µas accuracy are needed to attain a signal-to-noise ratio (SNR) of 5.8, the approximate threshold for detection. In general, if (cid:83) is the one-axis RMS noise per differential measurement, N is the number of SIM Lite visits, and (cid:65)THRESH is the threshold astrometric amplitude detectable with a probability of 50 percent, we have:(cid:65)THRESH = SNR × (cid:83)/ N1/2 SIM Lite offers the astrometric precision and duration (five years), along with a noise floor below 0.1 µas, to detect Earth-mass planets around the ~60 nearest FGK stars.3(cid:68)DEC, µas210 -- 1 -- 2 -- 3 -- 3 -- 2 -- 10123(cid:68)RA, µasObservedTrue(cid:68)RA, µas210 -- 1 -- 2012345Time, yearsObserved RA DisplacementTrue RA Displacement(cid:68)DEC, µas210 -- 1 -- 23Observed DEC DisplacementTrue DEC Displacement012345Time, yearsPower, µas20.40.30.20.10RA PeriodogramDEC PeriodogramDetection Threshold for 1% FAP10010Period, years100Period, years101Joint Periodogram Power, µas20.60.50.40.30.20.70.10.0Joint Periodogram, True Period 1.25 yrsDetection Threshold for 1% FAPDetected Period 1.25 yrs(b)(c)(d)(e)(a)Time (years)Time (years)Period (years)Period (years)δRA, μasδDEC, μasPower, μas2Joint Power, μas2(a)(b)(c)(d)(e)upsilon And (4.6)TYC2 2067(10.4)TYC2 2176 (12.8)TYC2 1302 (12.0)TYC2 2188 (11.3)TYC2 478 (11.7)TYC2 1396 (12.0)TYC2 1969 (12.1)TYC2 1474 (11.6)TYC2 352 (12.7)TYC2 2010 (10.0)TYC2 707(12.2)TYC2 454(9.8)TYC2 602(10.1)TYC2 560 (11.6)TYC2 286 (11.1)TYC2 1424 (12.0)TYC2 1182 (12.0)TYC2 1246 (11.5)TYC2 1542 (10.2)TYC2 1867 (10.8)TYC2 1880 (11.3)TYC2 1716 (10.6)TYC2 1668 (10.0)0.3°NE 4 Malbet, L´eger, Shao, Goullioud, Lagage et al. Table 1 Partial list of possible targets, the full list is available on the NEAT website (http: // neat. obs. ujf-grenoble. fr/ IMG/ xls/ Proposal_ targets_ total. xls ). Stars are ranked by decreasing astrometric signal for a planet in its habitable zone (HZ). This signal A(µas) is calculated for 0.5, 1 and 5 M⊕ planets around the 5, 70 and 200 first stars, respectively, assuming that the planet is located at the inner boundary of the HZ that secures its detection whenever the planet is in this zone. The corresponding integration time (tvisit in h) and cumulated times (ttot in h) are calculated for a detection with an equivalent SNR = 6. The total time corresponds to 70% of the available mission time with a 22% margin. Table 2 Left: summary of the main program capabilities and required resources. Right: Time and allocated maneu- vers for the different programs: (1) the Gaia Mission and its Exoplanet Science Potential; (2) NEAT follow-up program of Gaia detected planetary systems; (3) observations of young stars; and (4) characterizing planetary systems around some of the closest M stars. by van Leeuwen, 2007), disregarding spectroscopic bi- naries, and stars with an activity level 5 times greater than that of the Sun because of their astrometric noise (only 4% of the sample, Lagrange et al., 2011) and for which we compute the astrometric signal for a planet with given mass in the HZ of the stars (Kaltenegger et al., 2010). Conservatively, we select the inner part of the HZ in order to be able to detect the planet what- ever is its location in the HZ. The required number of visits and cumulative time to observe this list of target stars is summarized in Table 2. The list corresponds to an exhaustive search for 1 Earth mass planets (resp. 5 Earth mass planets) around K stars up to 6 pc (resp. 12 pc), G stars up to 10 pc (resp. 17 pc), and F stars up to 14 pc (resp. 19 pc) in the whole HZ of the star, exclud- ing spectroscopic binaries and very active stars. The spatial repartition of targets is shown in Fig. 3. 60% of the NEAT targets (118) are brighter than V = 6 and therefore will not be investigated by Gaia because of its bright limit. So, even if some of those sources do not harbor Earth-like planets, NEAT will be contributing Fig. 3 Representation of the NEAT targets in the 3D sphere of our neighborhood (D up to ≈ 15 pc). They correspond to a volume limited sample of all stars with spectral types between F and K. to the improvement of our knowledge about the neigh- borhood of our Solar System. In that respect, NEAT observations will not only be complementary to Gaia's ones, but NEAT data will also form a base to improve Gaia results. In addition to the survey for the NEAT main science program, we propose that 30% of NEAT time is allo- cated to study some objects of interest (planets around M dwarfs, young stars, multiple systems,... discovered by Gaia and others). The global required amount of time and number of maneuvers is listed in Table 2 (right part). 3 NEAT concept Our goal is to detect the signal corresponding to the reflex motion of a Sun-like star at 10 pc due to an Earth- mass planet in its HZ, with an equivalent final SNR of 6. RankStar_ident.Name1 HIP16537 eps Eri2 HIP8102 LHS 146(...)4 HIP104214 61 Cyg A5 HIP19849 40 Eri7 HIP99240 del Pav8 HIP96100 sig Dra(...)69 HIP64797 LTT 1385270 HIP23835 LHS 173671 HIP47592 LTT 355872 HIP26779 V538 Aur(...)199 HIP79537 LHS 413200 HIP18859 HD 25457Vmag SpType 3.72 K2V 3.49 G8V 5,20 K5V 4.43 K1V 3.55 G5IV 4.67 K0V 6,49 K2V4.91 G4V 4.93 G0V 6.21 K1V 7,53 K0V5.38 F5V D (pc)LogR'HKHZin (AU)3,2-4,6200,573,7-4,9580,583,5-4,7640,345,0-5,3800,626,1-5,0000,965,8-4,8320,6311,1-4,6300,5615,4-5,1271,2615,0-4,8621,1812,3-4,5500,6812,9-5,2050,4118,8-4,6001,02HZin (AU)ttot (h)Mlimit = 0,5 ME139337- 8041 0941 2651 357- 16 95717 34917 37717 405- 22 01322 058A (uas)tvisit (h)N_VisitsT_totalA (uas)tvisit (h)Mlimit = 0,5 MEMlimit = 1 MEM limit = 5 M E0,352,50,293,7- - 0,254,80,245,60,451,50,441,6- - 0,207,50,207,6T_totalA (uas)(visit (h)M limit = 5 M E1,000,311,000,31- - 0,700,640,690,65Number of starsMass threshold(M⊕)Cumulated time(h)Number of visits50.51,10050070115,6003,50020056,4006,000TotalTotal22,10010,000ProgramTime (h)ManeuversTransfert + com.3,650Main22,10010,000Add. 1-35,5002,000Add. 42,7501,000total margins9,800 (22%)7,000Total43,800 h (5 yrs)20,000 Nearby Earth Astrometric Telescope (NEAT) 5 That astrometric signal is 0.30 µas. The required noise floor is 0.05 µas, over 100 times lower than Gaia's best precision (7 µas). 3.1 Technical challenges Achieving sub-micro-arcsecond astrometric precision, e.g. 0.8 µas, in 1 hour and a noise floor under 0.05 µas with a telescope of diameter D requires mastering all effects that could impact the determination of the position of the point spread function. The typical diffraction lim- ited size of an unresolved star is about 1.2λ/D, which corresponds to 0.16 arcseconds for a 1-m telescope op- erating in the visible spectral region. The challenge is therefore to control these systematics effects to a level better than 1 part in 3 million (1 : 3×106). Even though differential astrometry of stars within the same field of view softens somewhat the requirement, this level of accuracy can only be obtained in an atmosphere-free space environment. Sub-micro-arcsecond level astrometry requires solu- tions to four challenges: -- Photon noise. Most targets are R ≤ 6 mag stars, but the required reference stars are R ≤ 11 mag so they dominate the photon noise. Using the mean stellar density in the sky, one finds that a field of view (FOV) as large as diam 0.6◦ is needed to get several (6 to 8) of theses references (see e.g. Fig 1). -- Beam walk. A classical three mirror anastigmat (TMA) telescope can also manage a 0.6◦ diffraction limited FOV. However the light coming from dif- ferent stars, and therefore from different directions, will hit the secondary and tertiary mirrors on differ- ent physical parts of the mirrors. The mirror defects will therefore produce different and prohibitive as- trometric errors between the images of the stars. Using a single mirror telescope solves this problem. To obtain sufficiently high angular resolution, a long focal length (≈ 40 m) for this mirror is needed, with no intermediate mirrors, a relatively unusual solu- tion in modern optical astronomy. -- Stability of the focal plane. Proper Nyquist sam- pling with typical detector pixels of the order of 10 µm requires a focal plane at a focal length of 40 m. Such a focal plane covering a FOV of 0.6◦ diameter would yield a costly detector mosaic with 40, 000 × 40, 000 ≈ 109 pixels. Sub-microarcesc as- trometry over a 0.6◦-diameter FOV requires the ge- ometry of the focal plane to be stable to ≈ 1 : 2 × 10−10. Therefore thermal stability of the focal plane geometry will be a major challenge although it has to be investigated in details. Instead of building Fig. 4 Proposed concept for a very high precision astrometry mission. It consists in two separated modules, the first one carrying the primary mirror (upper right) and the second one the detector plane (bottom left). a gigapixel focal plane with unprecedented stability we plan to use 9 small 512× 512 CCDs (Fig. 5) and a laser metrology system to measure the position of every pixel to the required precision, once every 10 to 30 s. We do not rely on their positioning, but measure it accurately with a laser metrology based on dynamic interference fringes. -- Quantum efficiency (QE) variations. The dy- namic fringes also allow the measurement of the inter- and intra-pixel QE variations. We characterize each pixel response with six parameters such that the systematic errors are kept below 10−6. This is a process derived from the SIM studies. These effects have in the past hampered the perfor- mance of space missions like HST. 3.2 Instrumental concept The proposed mission is based on a concept recently proposed by M. Shao and his colleagues that results from the experience gained in working with many as- trometry concepts (SIM, SIM-Lite, corono-astrometry2). The concept is sketched in Fig. 4 and consists of a pri- mary mirror -- an off-axis parabolic 1-m mirror -- a fo- cal plane 40 m away, and metrology calibration sources. The large distance between the primary optical surface and the focal plane can be implemented as two space- craft flying in formation, or a long deployed boom. The focal plane with the detectors having a field of view of 0.6◦ is shown in Fig. 5. It has a geometrical extent of 0.4m × 0.4m. The focal plane is composed of eight 512 × 512 visible CCDs located each one on an XY translation stage while the central two CCDs are fixed in position. The CCD pixels are 10µm in size. The principle of the measurement is to point the spacecraft so that the target star, which is usually brighter (R ≤ 6) than the reference stars (R ≤ 11), is located on the axis of the telescope and at the center of the central 2 See Guyon et al. (2010) MetrologyOff-axis parabolic mirror (D~1m)Focal plane array (FOV~0.6°, ∅~0.4m)Telescope spacecraftDetector spacecraftFocal length (~40m)Sun shadesTelescope Axis laser source 6 Malbet, L´eger, Shao, Goullioud, Lagage et al. Fig. 6 Principle of the metrology and the axis tracker. Left panel: the metrology laser light (in yellow) is launched from fibers located at the edge of the mirrors. Right panel: the laser beams interfere over the detector plane. Only the fringes corresponding to a pair of fibers are represented on this figure and they are not to scale, since the fringe spacing is equal to the PSF width. The axis tracker (sketched in red on the left panel) is a laser beam launched in the center of the mirror that is monitored in the lower central CCD. CCDs are read at 50 Hz providing many frames that will yield high accuracy. With the proposed concept, it is possible to achieve all of the main technical requirements: -- Focal plane stability. Instead of maintaining a focal plane geometry stable at the 0.1 nm level for a 5-yr duration, which is impossible, we implement a metrology for every pixel at the sub-nanometer level, with an interferometric system that has been qualified by the SIM-Lite laboratory demonstrators. -- Reference frame. By measuring the fringes at the sub-nanometer level using the information from all the pixels of each CCD (SIM-Lite technology), it is possible to solve for the position of all reference stars compared to the central target with an accuracy of 0.8 µas per hour. The field of view of 0.6◦ allows us to have 6 to 8 reference stars brighter than V = 11 in most fields. -- Photon noise. The field of 0.6◦ provides about 6 to 8 stars of magnitude brighter than R = 11. The number of photons received by one 11-mag star on the system is ≈ 4.1×109 ph/hr. Since the FWHM of diffraction-limited stars is 1.2λ/D = 0.16 arcsecond, √ the photon noise limit in 1 h of integration due to 6N ≈ 0.5 µas. a set of 6 reference stars is (λ/2D)/ With more than 50×2 measurements of a few hours spread over 5 yrs, the equivalent precision is 0.05 µas in RA and Dec, corresponding to the detection of the 0.30 µas signal with a SNR ≈ 6. Fig. 5 Schematic layout of the focal plane. The field of view is divided in 3 × 3 sub-fields. Exterior subfields have visible arrays which can be moved in X and Y directions to image the reference stars. The central field has two fixed arrays, one for the target star and one for the telescope axis tracker. CCD. Then the 8 other CCDs are moved to center each of the reference stars on one of them. To measure the distance between the stars, we use a metrology calibra- tion system that is launched from the telescope space- craft and that feeds several optical fibers (4 or more) located at the edge of the mirror. The fibers illuminate the focal plane and form Young's fringes detected si- multaneously by each CCD (Fig. 6). The fringes have their optical wavelengths modulated by acoustic opti- cal modulators (AOMs) that are accurately shifted by 10 Hz, from one fiber to the other so that fringes move over the CCDs. These fringes allow us to solve for the XYZ position of each CCD. An additional benefit from the dynamic fringes on the CCDs is to measure the QE of the pixels (inter-, and intra-pixel dependence). The Focal plane0.6° FOV, ~0.4m8 movable CCDs (reference stars)1 fixed CCD (telescope axis tracker)1 fixed CCD (target star) Nearby Earth Astrometric Telescope (NEAT) 7 -- Large-scale calibration. The detector plane does not have to be fully covered by pixels, since the posi- tions of the reference stars are known from available catalogues (10-20 mas for Tycho 2, and about tens of µas for Gaia). For the target stars (R ≤ 6), we use the Hipparcos catalog (few mas accuracy). This cor- responds to < 1/10th of the PSF or the fringe width. The number of fringes between the target star and the reference stars is then known, only the positions of the star centroids relative to the interferometric fringe have to be measured accurately. The use of 10 small CCDs drastically reduces the cost of what would otherwise be a giga-pixel focal plane and also helps to control systematics. With such a concept, the mission performance would be similar to, and even more favorable for exoplanets, than what was proposed for SIM-Lite with 5 years of operation, but at the price of giving up all-sky astrometry and the corresponding science objectives. 3.3 Performance assessment and error budget Achieving a relative precision of 2 × 10−10 is slightly better than the precision achievable by only the com- bination of our metrology laser, the thermal expansion coefficient of the primary mirror and our expected tem- perature stability. Achieving our target precision relies on not only the metrology stability, but also on the precise knowledge of the positions of the multiple ref- erence stars used since the expected motions of the ref- erences cannot be considered as fixed (see discussion in Sect. 4.1). Our comprehensive error budget takes into account all sources of error, including instrumental ef- fects, photon noise and astrophysical errors in the ref- erence star positions. The biggest term is the brightness dependent error for the set of Reference stars. The half-width of the PSF for the coma-aberrated images of the reference stars is about 19 µm on the focal plane (or 100 mas on the sky). After 1 s of integration, 1.3 × 106 pho- toelectrons are detected for each of the 11-mag refer- ence star; their centroid location can be estimated to 0.016 µm rms (1.6 × 10−3 pixel or 0.08 mas). Since all the stars are measured simultaneously, the stars do not need to be kept centered on the detector at the sub- mas level, but only to a fraction of the PSF width to avoid spreading of the photon outside of the PSF and therefore cause the PSF effective width to be larger. A tenth of pixel (1 µm) stability over the one-second integration is sufficient. After 3400 s of integration, the statistical averaged position of the barycenter of the set of reference stars (R ≤ 11 mag) will be measured with a residual 0.126 nm (0.63 µas) uncertainty. Similarly, the position of the target star (R ≤ 16 mag) will be mea- sured with a residual 0.024 nm (0.12 µas) uncertainty. Although the spacecraft will have moved by several arc- seconds, the differential position between the target star and the barycenter of the set of reference stars will be determined to 0.64 µas. Similarly, the focal plane metrology system will have determined the differential motion of the target CCD relative to the barycenter of the set of reference CCDs with an error smaller than 0.16 µas after 60×1 s metrol- ogy measurements. NEAT will not be capable of measuring the absolute separation between the target and the set of reference stars to 0.8 µas. NEAT objectives will therefore be to measure the change in the relative position of those stars between successive observations spread over the mission life, with an error of 0.8 µas for each one hour visit. The six major errors terms are captured in the simplified version of the error budget shown in Fig. 7. If unmonitored, the displacement of the projected field aberrations on the focal plane would produce a 60 µas differential astrometric error per arcsecond of relative spacecraft motion. The telescope axis tracker will monitor the relative position of the focal plane rela- tive to the parabola axis simultaneously with the stellar observation with a 1 mas accuracy per hour. This will be sufficient to correct the observations during post- processing for the field-dependent aberration to better than 0.1 µas. Static figure errors of the primary mirror will pro- duce centroid offsets that are mostly common-mode across the entire field of view. Differential centroid off- sets are significantly smaller than the field-dependent coma and are in fact negligible. Similarly, changes in the primary mirror surface error, e.g. due to thermal dilata- tion3, meteorite impacts,... produce mostly common- mode centroid shifts and negligible differential centroid offsets. On the other hand, displacement and changes in the shape of the PSF would couple with the CCD re- sponse if the CCD response is not properly calibrated. This is continuously done by the metrology fringes. 3.4 Design of the payload subsystems Focal plane assembly. A proposition for implementa- tion of the focal plane is shown in Fig. 8. The detector is foreseen to be a CCD fabricated by the E2V tech- nologies company in UK. The target star, the reference 3 The coefficient of thermal expansion of the mirror is about 100 times smaller than those of the elements that compose the detector. The metrology parameters are constantly mon- itored. 8 Malbet, L´eger, Shao, Goullioud, Lagage et al. Fig. 7 Top-level error budget for NEAT. It shows how the 0.8 µas accuracy enables the detection of 0.3 µas signatures with a signal to noise of 6 after 360 h of observation. It also shows the major contributors to the astrometric error. Fig. 8 Views of the focal plane assembly. (a) Magnified view of one of the 8 XY translation stages of the focal plane. In yellow, the 512 × 512 CCD and its support. In blue and green the two translation stages. (b) The front part of the focal plane with its 8 movable CCDs and two fixed CCDs at the center. (c) The electronics racks. star and the telescope axis tracker will all use the same CCD that includes the capability to read windowed im- ages, typically 10 × 10 to 30 × 30 pixels. The 8 XY tables consist of two linear tables mounted on top of each other. Each table uses a piezo-reptation motor4, a linear ball bearing system and an optical incremen- tal encoder. These motors fulfill several requirements of simplicity: they are self-locked when they are not pow- 4 Such reptile motors have been qualified by the Swiss firm RUAG for the LISA GPRM experiment. ered; they can be used both for large displacements by stepping up to 100 mm × 100 mm and elementary ana- log motion down to 50 nm. Since 8 tables are used in parallel in the focal plane, the loss of one table is not a single point failure. An alternative implementation could be to drive the XY tables with ball screws and rotary motors. The limited resolution of such a motor stage (about 5 µm) could be supplemented by a second high-resolution piezo XY table5, mounted on top of the 5 such as the Cedrat XY25XS Nearby Earth Astrometric Telescope (NEAT) 9 substrate. Three of them are located around the edge of the mirror, and are used by pairs in order to conduct three redundant measurements of the relative location of the CCDs. The fourth fiber is located inside the clear aperture, and is used in combination with each of the three other fibers to produce three additional measure- ments during focal plane calibration and calibrate the distance mirror-focal plane. Pointing servo systems. The pointing of the tele- scope from one target to the next one is accomplished by the two spacecraft in formation flying. The target stars will be typically separated by 10◦. Re-pointing of the telescope will require rotation of the two spacecraft by several degrees using reaction wheels and translation of the telescope spacecraft by several meters using hy- drazine propulsion. Fine positioning of the focal plane relative to the mirror is done by cold gas propulsion system, and at the end of the maneuver, the telescope spacecraft will be oriented to better than 3 arcseconds from the target star line of sight using star trackers and the focal plane spacecraft will be positioned to bet- ter than 2 mm from the primary mirror focus. At that point, the spacecraft will maintain their relative posi- tion to better ±2 mm in shear and in separation for the duration of the observation. The separation does not require a servo-loop of the payload, because its effect is only a degradation in performance (when FWHM in- creases, final precision decreases in same proportion) and is managed in the error budget. During the observation, the instrument uses a tip- tilt stage behind the primary mirror to center the target star on the 32× 32 pixel sub-window on the target star CCD. Once in the 32 × 32 pixel sub-frame mode, the target star CCD is read at 500 Hz, and feedback control between the CCD and the tip-tilt stage can be used to keep the star centered on the detector to better than 5 milli-arc-second RMS (0.1 pixel RMS) for the dura- tion of the observation. This is the only active feedback loop in the instrument system working at 50 Hz; the other degrees of freedom (focal plane tip, tilt, clock- ing and focal-plane-to-mirror separation) are monitored but not corrected for in real-time. Prior to acquisition, the reference star CCDs will be pre-positioned to the expected location of the reference stars using the trans- lation stages. The XY translation stage fine motion of the reference star CCDs at a 0.2 µm precision enables centering of the reference stars on the detectors to bet- ter than a tenth of a pixel. Once the reference stars are acquired, the translations stages are locked for the duration of the observation. Fig. 9 Laser metrology of the telescope axis tracker. The launcher embedded in the primary mirror substrate forms a 600 µm (60 pixels) FWHM diffraction limited image of the fiber output on the Telescope Axis Tracker CCD. By sampling the entire 512 × 512 pixel detector, the parabola axis can be tracked over a 26 arcsec range, with a 0.3 mas estimation error at 10 Hz. first XY table. The main structure of the focal XY ta- bles consists of a large lightweight aluminum cylinder which is thermally controlled and in which pockets are machined for the fixation of the XY tables. Telescope. The primary mirror is an off-axis paraboloid, with a 1 m diameter clear aperture, an off-axis distance of 1 m and a focal length of 40 m. It would be fabri- cated in either Zerodur or ULE, 70% light-weighted and weight about 60 kg. The surface quality should be bet- ter than λ/4 peak-to-valley and would be coated with protected aluminum. A 50 mm hole at the center of the mirror accommodates the beam launcher for the tele- scope axis tracker. The three bipods on the back of the mirror support the mirror with minimum deflec- tion. The bipods interface with the tip-tilt stage made of 3 preloaded piezo stacks on parallel flexures that pro- vide the +/-6 arcsecond amplitude for two-axis artic- ulations. The entire primary mirror assembly interface to the telescope payload plate is a 34 kg hogged-out aluminum plate. This plate also hosts the metrology source, the telescope drive electronics, the telescope baf- fle and the interface to the spacecraft. The telescope axis tracker is used to estimate the location of the pri- mary mirror axis with respect to the focal plane in or- der to monitor it and then correct for the telescope field dependent errors. The sensor is the second fixed CCD located in the focal-plane. The launcher would consist of either an achromatic doublet or an aspherical sin- glet lens, embedded in the primary mirror substrate at its center and a single-mode fiber-coupled laser diode. Fig. 9 shows how the fiber tip is re-imaged onto tele- scope axis tracker CCD in the focal plane. Laser metrology. The focal-plane metrology sys- tem consists of the metrology source similar to the one developed for SIM (Erlig et al., 2010), the metrology fiber launchers and the focal plane detectors (CCDs) which alternatively measure the stellar signal (57 s ob- servations) and the metrology (1 s per axis every min- utes). The metrology fiber launchers consist of nomi- nally four optical fibers attached to the primary mirror ! 10 3.5 Mission requirements The objectives of the NEAT mission require to per- form acquisitions over a large number of targets during the mission timeline, associated to a 40 m focal length telescope satellite. The preliminary assessment of the NEAT mission requirements allows to identify the fol- lowing main spacecraft design drivers. Launch configuration and mission orbit. The L2 orbit is the preferred orbit, as it allows best forma- tion flying performance and is particularly smooth in terms of environment. The Soyuz launch, proposed as a reference for medium class missions, offers satisfying performance both in terms of mass and volume. Formation Flying and 40 m focal length. The mission relies on a 40 m focal length telescope, for which the preferred solution is to use two satellites in forma- tion flying. The performance to be provided by the two satellites in order to initialize the payload metrology systems are of the order of magnitude of ±2 mm in rel- ative motion, and of 3 arcseconds in relative pointing, which are typically compatible with Formation Flying Units and gyroless AOCS6 architecture. In addition, at L2, the solar pressure is the main disturbance for for- mation flying control. As a result, surface-to-mass ra- tio (S/M ratio) is the main satellite drift contributor and should be as close as possible for the two satellites. Although satellite design can cope with these require- ments, the S/M ratio of the satellites will evolve dur- ing the mission (because of fuel losses and sun angle). However, the preliminary mission assessment tends to demonstrate that the S/M difference between the two satellites can be reduced down to 20-30%, which is deemed compatible with mission formation flying re- quirements. Number of acquisitions and Mission ∆V . The mission aims at a complete survey of a large number of targets and the maximization of the number of ac- quisitions will be a main objective of the next mission phases. The mission objectives require a threshold of 20,000 acquisitions (see Sect. 3.6 for details). In addi- tion, the time allocation for these reconfiguration ma- neuvers is quite limited, in order to free more than 85% of mission duration for observations. As a result, the mission is characterized by a large ∆V (550 to 880 m/s) dedicated to reconfigurations, plus allocations for fine relative motion initialization and control using the µ- propulsion system. This large number of reconfigura- tions is also driving the number of thruster firing, which are qualified to typical numbers of up to 5,000 to 50,000 with cycling as required for NEAT. 6 Attitude and Orbit Control System Malbet, L´eger, Shao, Goullioud, Lagage et al. Baffles and Parasitic Light. The mission perfor- mance relies on the ability of the focal plane to receive only star flux reflected by the telescope satellite. A first requirement is to implement baffles on the two satel- lites, coupled by a diaphragm on the focal plane. In ad- dition, all parasitic light coming from telescope satellite reflections should be avoided, thus requiring all bus el- ements to be shielded by a black cover. Following this preliminary satellite requirement analysis, a first simple and robust mission concept has been identified. 3.6 Preliminary Spacecraft Design The preliminary NEAT mission assessment allowed to identify a safe and robust mission architecture (Fig. 10), relying on high technology-readiness-level (TRL) tech- nologies, and leaving safe margins and mission growth potential that demonstrates the mission feasibility within the medium class mission cost cap. System Functional Description. The proposed mission architecture relies on the use of two satellites in formation flying (FF). The two satellites are launched in a stacked configuration using a Soyuz ST launcher, and are deployed after launch in order to individu- ally cruise to their operational Lissajous orbit. Acqui- sition sequences will alternate with reconfigurations, during which the Telescope Satellite will use its large hydrazine propulsion system to move around the Fo- cal Plane Satellite and to point at any specified star. At the approach of the correct configuration, the Focal Plane Satellite will use a cold gas µ-propulsion system for fine relative motion acquisition. The Focal Plane Satellite will be considered as the chief satellite regard- ing command and control, communications and payload handling. Communications with the L2 ground station would typically happen on a daily basis through the Fo- cal Plane Satellite, with data relay for TC/TM7 from the Telescope Satellite using the FFRF8 units. This satellite will however be equipped with a similar com- munication subsystem, in order to support cruise and orbit acquisition, and to provide a secondary backup link. Formation Flying Architecture. The formation flying will have to ensure anti-collision and safeguard- ing of the flight configuration, based on the successful PRISMA flight heritage. In addition, the spacecraft will typically perform 12 to 20 daily reconfigurations of less than 10◦ of the system line of sight corresponding to 7 m of translation of one satellite compared to the other perpendicular to the line of sight. During these configu- 7 Telecommand / Telemetry 8 Formation Flying Radio Frequency Nearby Earth Astrometric Telescope (NEAT) 11 Fig. 10 Left: NEAT spacecraft in operation with the two satellites separated by 40 m. Right: closer external view of the two satellites. rations, the telescope satellite will perform translations -- supported by the FF-RF Units -- using its large hy- drazine tanks (250 kg) for a ∆V ≈ 605 m/s. When the two satellites will approach the required configuration, the telescope satellite will freeze, and the focal plane satellite will perform fine relative pointing control us- ing micro-propulsion system. As a result, the micro- propulsion will have to compensate for hydrazine con- trol inaccuracies, which will require large nitrogen gas tanks (92 kg for ∆V ≈ 75 m/s). Finally, 28 kg of hy- drazine carried by the FP satellite allows ∆V ≈ 55 m/s for station keeping and other operations. Satellite Design Description. The design of the two satellites is based on a 1194 mm central tube ar- chitecture, which will allow a low structural index for the stacked configuration and provides accommodation for payloads and large hydrazine tanks. Strong heritage does exist on the two satellites avionics and AOCS. In addition, they both require similar function which would allow to introduce synergies between the two satellites for design, procurement, assembly, integra- tion and tests. The proposed AOCS configuration is a gyroless architecture relying on reaction wheels and high-performance star trackers (Hydra Sodern), which is compatible with a 3 arcsec pointing accuracy (see end of Sect. 3.4 for payload control). The satellites com- munication subsystems use X-Band active pointing an- tenna, supported by large gain antenna for low Earth orbit positioning and cruise, coupled with a 50 W RF Transmitter. The active pointing medium gain antenna allows simultaneous data acquisition and downlink. A reference solution for the satellite on-board computer could rely on the Herschel-Planck avionics. Fig. 11 NEAT stowed configuration The two satellites would have custom mechanical- thermal-propulsion architectures. The telescope satel- lite features a dry mass of 724 kg and the focal plane satellite a dry mass of 656 kg. The focal plane satellite carries the stacked configuration. The payload (focal plane + baffle) are assembled inside a 1194 mm central tube, which will also ensure the stacked configuration structural stiffness. The spacecraft bus, and large cold gas tanks, will be assembled on a structural box carried by the central tube. The proposed architecture uses a large hydrazine tank inside the 1194 mm central tube which offers a capacity of up to 600 kg hydrazine, thus allowing both a low filling ratio and a large mission growth potential. The payload module -- with the pay- load mirror, rotating mechanisms and baffle -- is then assembled on the central tube. Proposed Procurement Approach. The NEAT mission is particularly adapted to offer a modular space- 12 Malbet, L´eger, Shao, Goullioud, Lagage et al. objective is of astrophysical nature, the impact of stellar activity. Spots and bright structures on the stellar sur- face induce astrometric, photometric and RV signals. Using the Sun as a proxy, Lagrange et al. (2011) have computed the astrometric, photometric and RV vari- ations that would be measured from an observer lo- cated 10 pc away. It appears that the astrometric vari- ations due to spots and bright structures are small com- pared to the signal of an Earth mass planet in the HZ (see also Meunier et al., 2010; Makarov et al., 2009, 2010). This remains true throughout the entire solar cycle. If we consider a star 5 times more active than the active Sun, an Earth-mass planet would still be de- tectable even during the highest activity phases. Such activity, or lower, translates in terms of activity index HK) ≤ −4.35. Consequently, in our target list, log(R(cid:48) we have kept only stars with such an index (only 4% were discarded ), for which their intrinsic activity should not prevent the detection of an Earth-mass planet, even during its high activity period. Perturbations from reference stars. The vast majority of the reference stars will be K giants at a dis- tance of ≈ 1 kpc. The important parameters in addition to the position are the proper motion with typical value of ≈ 1 mas/yr and the parallax whose typical value is ≈ 1 mas. They are to be compared to the accuracy of the cumulative measurements during a visit. An impor- tant value for NEAT accuracy is what is obtained for an R = 6 magnitude target: 0.8 µas/h. The ratios between that (required) accuracy and the expected motions of the references indicates clearly that the latter cannot be considered as fixed. Their positions are members of the set of parameters that have to be solved for. Because the reference stars are much more distant (≈ 1 kpc) than the target star (≈ 10 pc), we are 100 times less sensitive to their planetary perturbations. Only Saturn-Jupiter mass objects matter, and statistically, they are only present around ≈ 10% of stars. These massive plan- ets can be searched for by fitting first the reference star system (≈ 100Nref measurements for 5Nref parameters when there are no giant planets around the reference stars), possibly eliminate those with giant planets, and studying the target star with respect to that new ref- erence frame. Moreover, the largest disturbers will be detected from ground based radial velocity measure- ments, and the early release of Gaia data around 2016 will greatly improve the position accuracy of the refer- ence stars. For smaller planets at or below the threshold of detection, their impact on the target astrometry will be only at a level (cid:28) 1 M⊕ around it. Similarly the activ- ity of these K giants has been investigated and neither the stellar pulsations nor the stellar spots will disturb the signal at the expected accuracy. Fig. 12 Flight system concept for the deployable telescopic tube version craft approach, with simple interfaces between payload and spacecraft bus elements. For both satellites, the payload module is clearly identified and assembled in- side the structural 1194 mm central tube. In addition, a large number of satellite building blocks can be common to the two satellites, in order to ease mission procure- ment and tests. This configuration is particularly com- patible with the ESA procurement scheme. The pay- load is made of 3 subsystems: primary mirror and its dynamic support, the focal plane with its detectors and the metrology. Alternative mission concept. An alternative mis- sion concept would consist of a single spacecraft with an ADAM-like9 deployable boom (from ATK-Able en- gineering) that connects the telescope and the focal plane modules. The preliminary investigation made by CNES identified no show-stoppers for this option: no prohibitive oscillation modes during observation; dur- ing maneuvers, the boom oscillation modes can be ex- cited but they can be filtered by Kalman filters (SRTM10 demonstration). The use of dampers on the boom struc- ture allows damping at a level of 10% of the oscillations. The main worry concerns retargeting, which requires large reaction wheels or control momentum gyroscopes (CMGs) on the spacecraft due to the important inertia but propellers could be added at the boom end. A pos- sible implementation made by JPL is shown in Fig. 12. 4 Discussion 4.1 Astrophysical issues Stellar activity. If all instrumental problems are con- trolled then the next obstacle to achieve the scientific 9 ADAM: ABLE Deployable Articulated Mast 10 Shuttle Radar Topography Mission Nearby Earth Astrometric Telescope (NEAT) 13 Planetary system extraction from astromet- ric data. We recently carried out a major numerical simulation to test how well a space astrometry mission could detect planets in multi-planet systems (Traub et al., 2010). The simulation engaged 5 teams of the- orists who generated model systems, and 5 teams of double-blind "observers" who analyzed the simulated data with noise included. The parameters of the study were the same as for NEAT, viz., astrometric single- measurement uncertainty (0.80 µas noise, 0.05 µas floor, 5-year mission, plus RV observations with 1 m/s ac- curacy for 15 years). We found that terrestrial-mass, habitable-zone planets (≈ Earths) were detected with about the same efficiency whether they were alone in the system or if there were several other giant-mass, long-period planets (≈ Jupiters) present. The reason for this result is that signals with unique frequencies are well separated from each other, with little cross-talk. The number of planets per system ranged from 1 to 11, with a median of 3. The SNR value of 5.8 value was predicted by Scargle (1982) for a false alarm probability (FAP) of less than 1%, and verified in our simulations. The completeness and reliability to detect planets was better than 90% for all planets, where the comparison is with those planets that should have been detected ac- cording to a Cramer-Rao estimate (Gould et al., 2010) of the mission noise. The Cramer-Rao estimates of un- certainty in the parameters of mass, semi-major axis, inclination, and eccentricity were consistent with the observed estimates of each: 3% for planet mass, ≈ 4◦ for inclination and 0.02 for eccentricity. Radial velocity screening. To solve unambigu- ously for giant planets with periods longer than 5 yrs, it is necessary to have a ground RV survey for 15 yrs of the 200 selected target star, at the presently avail- able accuracy of 1 m/s. More than 80% of our targets are already being observed by RV, but the observations of the rest of them should start soon, well before the whole NEAT data is available. The capability of ground based RV surveys, despite their impressive near-term potential to obtain accuracies better than 1 m/s, is not sufficient to detect terrestrial planets in the HZ of F, G and K stars. Formally, an accuracy of 0.05 m/s is required to see an edge-on Earth mass planet at 1 AU from a solar-mass star with SNR=5 (semi-amplitude = 0.13 m/s), which might be achievable instrumentally, but is stopped in most cases by the impact of stellar activity on RV accuracy. It is necessary to find partic- ularly "quiet" stars, but they are a minority (few per- cents) and cannot provide a full sample. Furthermore, the ambiguity in physical mass associated with the sig- nal coming only from the radial component of the stellar reflex motion (sin i ambiguity) requires additional in- formation to determine the physical mass and relative inclination in complex planetary systems. In some, but not all cases, limits are possible, and one can argue sta- tistically that 90% of systems should be oriented such that the physical planet mass is within a factor of two of the mass found in RV. However, for finding a small number of potential future targets for direct detection and spectroscopy, an absolute determination that the mass is Earth-like is required as well as an exhaustive inventory of the planets around stars in our neighbor- hood. Flexibility of objectives to upgrades / down- grades of the mission. One of the strengths of NEAT is its flexibility, the possibility to adjust the size of the instrument with impacts on the science that are not prohibitive. The size of the NEAT mission could be re- duced (or increased) with a direct impact on the ac- cessible number of targets but not in an abrupt way. For instance, for same amount of integration time and number of maneuvers, the options listed in Table 3 are possible, with impacts on the number of stars that can be investigated down to 0.5 and 1 Earth mass, and on the mass of the instrument, required fuel for maneu- vers, and therefore cost. The time necessary to achieve a given precision depends on the mass limit that we want to reach: going from 0.5 M⊕ to 1 M⊕ requires twice less precision and therefore 4 times less observing time allowing a smaller telescope. There is room for adjust- ment keeping in mind that one wants to survey the neighborhood with the smallest mass limit possible and a typical number of targets of ≈ 200. 4.2 Technical issues Optical aberrations. NEAT uses a very simple tele- scope optical design. A 1-m diameter clear aperture off- axis parabola, with an off-axis distance of 1 m and a 40 m focal length. The focal plane is at the prime fo- cus. The telescope is diffraction limited at the center of the field, where the target stars will be observed, but coma produces some field dependent aberrations. At the mean position of the reference stars, 0.2◦ away from the center of the field, the coma produces a steady 23% increase of the point spread function (PSF) width and an 8 µm centroid offset. The impact remains low since we are looking at differential effects. Centroid measurements. They consist of two steps: the determination of the stellar centroid on each CCD during 57 s and then the calibration of the relative po- sition of the CCDs during 3 s thanks to the metrology. The metrology determines also the response map of the detectors. As in the normal approach to precision as- trometry with CCDs, we perform a least-square fit of 14 Malbet, L´eger, Shao, Goullioud, Lagage et al. Table 3 Science impact of NEAT scaling. The nominal mission is highlighted in yellow. a template PSF to the pixelated data. PSF knowledge error leads to systematic errors in the conventional cen- troid estimation. We have developed an accurate cen- troid estimation algorithm by reconstructing the PSF from well sampled (above Nyquist frequency) pixelated images. In the limit of an ideal focal plane array whose pixels have identical response function (no inter-pixel variation), this method can estimate centroid displace- ment between two 32x32 images to sub-micropixel accu- racy. Inter-pixel response variations exist in real CCDs, which we calibrate by measuring the pixel response of each pixel in Fourier space11. Capturing inter-pixel vari- ations of pixel response to the third order terms in the power series expansion, we have shown with simulated data that the centroid displacement estimation is accu- rate to a few micro-pixels. Stability of the primary mirror. The primary optic will be made of zerodur/ULE with a temperature coefficient better than 10−8/K with an optics thickness ≈ 10 cm and the effective temperature and temperature gradients are kept stable to ≈ 0.1 K over the mirror, the optic is then stable to ≈ 0.1 nm (λ/6000) during the 5 yr mission. We have simulated two images, one at the center of the field that is a perfect Airy func- tion and one at the edge of the field that has a λ/20 coma. We added also wavefront errors with a conser- vative rms value of λ/1000. With the new wavefronts, we calculated the change in the differential astrome- try bias caused by both pixelation and changing wave- fronts. While the wavefront deviations to optimal shape caused a centroid shift of ≈ 6−10 µas (10−4 pixels), dif- ferential errors remained less than ≈ 0.3 µas (3 × 10−6 pixels). CCD damage in L2 environment. CCDs, like most semiconductors, suffer damage in radiation envi- ronments such as encountered by space missions. One particular performance parameter, Charge Transfer Ef- ficiency (CTE), degrades with known consequences on 11 They are determined by calculating the first 6 coefficients of the Taylor series expansion in powers of wave numbers of the detector response map Fourier components. the efficiency of science missions like Gaia12. The re- duced CTE is caused generally by prompt particle events (PPE), including solar protons and cosmic rays, collid- ing with the CCD silicon lattice and causing damages to the silicon lattice. This leads to the formation of so-called traps which can capture photo-electrons and release them again after some time. This results in sig- nal loss and distortion of the PSF shape. The latter leads to systematic errors in the image location due to a mismatch between the ideal PSF shape and the actual image shape. For Gaia, this effect of radiation damage is a major contributor to the error budget and extensive research and laboratory tests have been done in order to understand better the radiation damage effects and to develop approaches in both hardware and data pro- cessing to mitigate the negative impact. However, there are a number of important differences between NEAT and Gaia which justify the assumption that radiation damage effects will play a much smaller role: i) NEAT looks for extended periods at very bright stars com- pared to Gaia in which the stars continuously move on the CCD. Also, unlike Gaia, NEAT will not be oper- ated in time-delayed integration mode. In addition the CCDs are regularly illuminated by the laser light from the metrology system. This means that in general the signal level in the CCD pixels is high which will keep the traps with long (≥ 60 s) release time constants filled and effectively inactive. ii) NEAT also does not suf- fer from the varying CCD illumination history that a scanning mission like Gaia necessarily encounters. This illumination history is in fact one of the major compli- cating factors for Gaia. Finally, iii) NEAT uses much smaller CCDs than Gaia and in addition has four read- out nodes, thus reducing the number of charge transfer steps and mitigating the effects of radiation damage. The one concern for the NEAT case is the presence of traps with release time constants that are of the order of several times the charge transfer period between pix- els. In the case of NEAT the transfer period averages 12 The Gaia community (http://www.rssd.esa.int/gaia) speaks of the complementary quantity, charge transfer ineffi- ciency (CTI), in order to emphasize its detrimental effects. !Mission name Mirror diameter Focal length Field of view diameter Focal Plane size Ref. star mean magnitude DMA in 1h # targets for a given mass limit (m) (m) (deg) (cm) (R mag) (µas) 0.5M! 1 M! 5 M! NEAT plus 1.2 50 0.45 40 11.5 0.7 7 100 200 NEAT 1.0 40 0.56 40 11 0.8 5 70 200 NEAT light 0.8 30 0.71 35 10.5 1.0 4 50 200 EXAM 0.6 20 0.85 30 10.1 1.4 2 35 200 DMA = Differential astrometric Measurement Accuracy (rms) !! Nearby Earth Astrometric Telescope (NEAT) 15 tens of µs and from laboratory tests with E2V CCDs, carried out in the context of the Gaia project, traps with time constants of 10 -- 100 µs are known to exist. If these traps dominate at the operating temperature of the NEAT CCDs they could lead to subtle PSF image shape distortions and thus image location biases. From the Gaia experience, it is known that such image shape distortions can be handled in the post-processing by a careful modeling of the effects of radiation damage on the PSF image. A similar strategy, building on the Gaia heritage, can be employed for NEAT. CCD/metrology tests in the lab. In the ab- sence of optical errors, the major error sources are as- sociated with the focal plane: (1) motions of the CCD pixels, which have to be monitored to 3×10−6 pixels ev- ery 60 s, i.e. 0.03 nm; (2) measurements of the centroid of the star images with 5 × 10−6 pixel accuracy. We have set up technology testbeds to demonstrate that we can achieve these objectives. The technology objec- tive for (1) has almost been reached and the technol- ogy demonstration for (2) is underway and should be completed soon. Latest results with no metrology nor QE 6-parameter calibration have been obtained from the CCD / metrology test bench (Fig. 13). Allen devi- ation of the centroid location for one artificial star "A" projected on the CCD and Allen deviation of the differ- ential centroid location for two artificial stars A and B projected on the CCD are plotted in Fig. 13 (left). One can see that star A moves on the CCD by a few hun- dred micro-pixels at time scale greater than 1 second, but that the differential position of the two stars is bet- ter 4 × 10−5 pixel at 100s integration time. On Fig. 13 (right), the Allen deviation of the differential centroid location for two artificial stars projected on the CCD is plotted, concatenating data from 38 runs, a minimum of ≈ 20 µ-pixels at about 10 min before differential drift dominated. This data shows that we are only a factor 10 from the final goal and that differential metrology at intervals of minutes is required to reach it. 5 Perspectives In the Cosmic Vision plan for 2015-2025, the commu- nity has identified in Theme 1 the question: "What are the conditions for planet formation?", and the recom- mendation in Sect. 1.2: "Search for planets around stars other than the Sun..." ultra high precise astrometry as a key technique to explore our solar-like neighbors. "On a longer timescale, a complete census of all Earth-sized planets within 100 pc of the Sun would be highly desirable. Building on Gaia's ex- pected contribution on larger planets, this could be achieved with a high-precision terrestrial planet astrometric surveyor." We have designed NEAT to be this astrometric sur- veyor. In Europe, as discussed in detail in the conclu- sions of the conference Pathways to Habitable Planets (Coud´e Du Foresto et al., 2010) and in the Blue Dot Team report, the exoplanet community recognizes the importance of astrometric searches for terrestrial plan- ets and has prioritized this search as a key question in the mid-term, i.e. in the time frame 2015-2022. The Ex- oPlanet Task Force (ExoPTF) in the US made a simi- lar statement. Finally the ESA dedicated ExoPlanetary Roadmap Advisory Team (EPRAT) prioritizes Astro- metric Searches for Terrestrial Planets in the mid term, i.e. in the time frame 2015-2022. Although the Decadal Survey of Astronomy and Astrophysics for 2010-2020 ranked down the SIM-Lite proposal, but placed as num- ber one priority a program "to lay the technical and sci- entific foundation for a future mission to study nearby Earth-like planets". Because of these recommandations by the commu- nity, we believe that there is a place for a mission like NEAT in future space programs, that is to say, a mis- sion that is capable of detecting and characterizing plan- etary systems orbiting bright stars in the solar neigh- borhood that have a planetary architecture like that of our Solar System or an alternative planetary system partly composed of Earth-mass planets. These stars vis- ible with the naked eye or simple binoculars, if found to host Earth-mass planets, will change humanity's view of the night sky. Acknowledgements This work has benefited support from the Centre National des ´Etudes Spatiales (CNES), the Jet Propulsion Laboratory (JPL), Thales Alenia Space (TAS) and Swedish Space Corporation (SSC). References Coud´e du Foresto V., Gelino D.M., Ribas I., Pathways Towards Habitable Planets, ASPC Series 430 (2010) Erlig H, Qiu Y., Poberezhskiy I., Meras P., Wu J., Reliable optical pump architecture for highly coher- ent lasers used in space metrology applications, SPIE 7734, E71 (2010) Gould A., Dong S., Gaudi B.S., Udalski A., Bond I.A., Greenhill J., Street R.A., Dominik M., Sumi T., Szyma´nski M.K., Han C., et al. Frequency of Solar-like Systems and of Ice and Gas Giants Beyond the Snow Line from High-magnification Microlensing Events in 2005-2008, ApJ 720, 1073 (2010) Guyon O., Shaklan S., Levine M., Cahoy K., Tenerelli D., Belikov R., Kern B., The pupil mapping exo- 16 Malbet, L´eger, Shao, Goullioud, Lagage et al. Fig. 13 Latest results obtained from the CCD/metrology test bench. No metrology nor quantum efficiency 6-parameter calibration have been performed yet. Left: Allen deviation of the centroid location for one artificial star "A" projected on the CCD and Allen deviation of the differential centroid location for two artificial stars A and B projected on the CCD. One can see that star A moves on the CCD by a few hundred micro-pixels at time scale greater than 1 second, but that the differential position of the two stars is better 4 10−5 pixel at 100 s integration time. Right: Allen deviation of the differential centroid location for two artificial stars projected on the CCD. Concatenating data from 38 runs, a minimum of ≈ 20 µpixels at about 10 min before differential drift dominated. This data shows that differential metrology at intervals of minutes is needed. planet coronagraphic observer (PECO), SPIE 7731, 68 (2010) Kaltenegger L., Eiroa C., Ribas I., Paresce F., Leitzinger M., Odert P., Hanslmeier A., et al. Stellar Aspects of Habitability - Characterizing Target Stars for Terrestrial Planet-Finding Missions, Astrobiology 10, 103 (2010) Lagrange A., Meunier N., Desort M., Malbet F. Us- ing the Sun to estimate Earth-like planets detection capabilities . III. Impact of spots and plages on as- trometric detection, A&A 528, L9 (2011) Makarov V.V., Beichman C.A., Catanzarite J.H., Fis- cher D.A., Lebreton J., Malbet F., Shao M. Starspot Jitter in Photometry, Astrometry, and Radial Veloc- ity Measurements, ApJL 707, L73 (2009) Makarov V.V., Parker D., Ulrich R.K. Astrometric Jit- ter of the Sun as a Star, ApJ 717, 1202 (2010) Meunier N., Lagrange A., Desort M. Reconstructing the solar integrated radial velocity using MDI/SOHO, A&A 519, A66 (2010) Scargle J.D. Studies in astronomical time series anal- ysis. II - Statistical aspects of spectral analysis of unevenly spaced data. ApJ 263, 835 (1982) Traub W.A., Beichman C., Boden A.F., Boss A.P., Casertano S., Catanzarite J., Fischer D., et al. De- tectability of Earth-like Planets in Multi-Planet Sys- tems: Preliminary Report. EAS Pub Series 42, 191 (2010) van Leeuwen F. Validation of the new Hipparcos reduc- tion. A&A 474, 653 (2007) Author affiliations F. Malbet -- Ph. Feautrier -- A.-M. Lagrange -- A. Chelli -- G. Duvert -- T. Forveille -- N. Meunier : UJF- Grenoble 1 / CNRS-INSU, Institut de Plan´etologie et d'Astrophysique de Grenoble (IPAG), UMR 5274, BP 53, F-38041 Grenoble cedex 9, France. E-mail: [email protected] A. L´eger : Universit´e Paris Sud / CNRS-INSU, In- stitut d'Astrophysique Spatiale (IAS) UMR 8617, Bat. 120-121, F-91405 Orsay cedex, France M. Shao -- R. Goullioud -- W. Traub : Jet Propulsion Laboratory (JPL), California Institute of Technol- ogy, 4800 Oak Grove Drive, Pasadena CA 91109, USA P.-O. Lagage -- C. Cara -- G. Durand : Labora- toire AIM, CEA-IRFU / CNRS-INSU / Universit´e Paris Diderot, CEA Saclay, Bat 709, F-91191 Gif- sur-Yvette Cedex, France A.G.A. Brown -- H.J.A. Rottgering : Leiden Obser- vatory, Leiden University, P.O. Box 9513, NL-2300 RA Leiden, The Netherlands C. Eiroa -- J. Maldonado -- E. Villaver : Universidad Aut´onoma de Madrid (UAM), Dpto. F´ısica Te´orica, M´odulo 15, Facultad de Ciencias, Campus de Can- toblanco, E-28049 Madrid, Spain ! Nearby Earth Astrometric Telescope (NEAT) 17 B. Jakobsson : Swedish Space Corporation (SSC), P.O. Box 4207, SE-17104 Solna, Sweden E. Hinglais : Centre National d'Etudes Spatiales (CNES), Centre spatial de Toulouse, 18 avenue Edouard Belin, F-31401 Toulouse cedex 9, France L. Kaltenegger -- C. Mordasini : Max-Planck-Institut fur Astronomie, Konigstuhl 17, D-69117 Heidelberg, Germany L. Labadie : I. Physikalisches Institut der Univer- sitat zu Koln, Zulpicher Str. 77, D-50937 Koln, Ger- many J. Laskar -- N. Rambaux : UPMC-Paris 6 / Observa- toire de Paris / CNRS-INSU, Institut de m´ecanique c´eleste et de calcul des ´eph´em´erides (IMCCE) UMR 8028, 77 avenue Denfert-Rochereau, F-75014 Paris, France R. Liseau : Chalmers University of Technology, SE- 41296 Gothenburg, Sweden J. Lunine : Dipartimento di Fisica, University of Rome Tor Vergata, Via della Ricerca Scientifica 1, Roma I-00133, Italy M. Mercier : Thales Alenia Space, 100 boulevard du Midi, F-06150 Cannes, France D. Queloz -- J. Sahlmann -- D. S´egransan : Obser- vatoire Astronomique de l'Universit´e de Geneve, 51 Chemin des Maillettes, CH-1290 Sauverny, Switzer- land A. Quirrenbach : Universitat Heidelberg, Landesstern- warte, Konigstuhl 12, D-69117 Heidelberg, Germany A. Sozzetti -- A.H. Andrei : INAF - Osservatorio Astronomico di Torino, Strada Osservatorio 20, I- 10025 Pino Torinese, Italy O. Absil -- J. Surdej : Universit´e de Li`ege, D´epartement d'Astrophysique, G´eophysique et Oc´eanographie, 17 All´ee du Six Aout, B-4000 Sart Tilman, Belgium Y. Alibert : Physikalisches Institut, University of Bern, Sidlerstrasse 5, 3012, Bern, Switzerland Y. Alibert : Universit´e de Besan¸con / Observatoire de Besan¸con / CNRS-INSU UMR 6213, Institut UTI- NAM, 41 bis Avenue de l'Observatoire, BP 1615, F-25010 Besan¸con cedex, France A.H. Andrei : Observatorio Nacional - Minist´erio da Ciencia e Tecnologia, Rua General Cristino 77, S¯ao Crist´ov¯ao 20921-400, Rio de Janeiro, Brazil F. Arenou : Universit´e Paris 7 Diderot / Observa- toire de Paris / CNRS-INSU, UMR 8111, "Galaxie Etoile Physique Instrumentation" (GEPI), 5 Place Jules Janssen, F-92190 Meudon cedex, France C. Beichman : California Institute of Technology, NASA Exoplanet Science Institute / IPAC, 770 South Wilson Ave, Pasadena CA 91125, USA C.S. Cockell : The Open University, Dept. of Physics & Astronomy, Planetary and Space Sciences Re- search Institute, Milton Keynes MK7 6AA, UK P.J.V. Garcia : Universidade do Porto, Faculdade de Engenharia, Departamento de Engenharia Fsica, Laboratrio SIM, Rua Dr. Roberto Frias, P-4200-465 Porto, Portugal D. Hobbs : Lund Observatory, Lund University, Box 43, SE-22100 Lund, Sweden A. Krone-Martins : Instituto de Astronomia, Geof´ısica e Ciencias Atmosf´ericas, Universidade de Sao Paulo, Rua do Matao, 1226, Cidade Universit´aria, 05508- 900 Sao Paulo-SP, Brazil A. Krone-Martins -- S. Raymond -- F. Selsis : Uni- versit´e de Bordeaux 1 / Observatoire Aquitain des Sciences de l'Univers / CNRS-INSU, UMR 5804, Laboratoire d'Astrophysique de Bordeaux, BP 89, F-33271 Floirac Cedex, France H. Lammer : Space Research Institute, Austrian Academy of Sciences, Schmiedlstr. 6, A-8042 Graz, Austria S. Minardi : Institute of Applied Physics, Friedrich Schiller, University Jena, Max-Wien-Platz 1, D-07743 Jena, Germany A. Moitinho de Almeida -- A. Krone-Martins : Sys- tems, Instrumentation and Modeling (SIM) - Fac- uldade de Ciencias da Universidade de Lisboa, Ed. C8, Campo Grande, 1749-016 Lisboa, Portugal P.A. Schuller : Universit´e Paris 7 Diderot / Univer- sit´e Pierre et Marie Curie / Observatoire de Paris / CNRS-INSU, UMR 8109, Laboratoire d'´etudes spa- tiales et d'instrumentation en astrophysique (LESIA), 5 Place Jules Janssen, F-92190 Meudon cedex, France 18 Malbet, L´eger, Shao, Goullioud, Lagage et al. G.J. White : The Open University, Dept. of Physics & Astronomy, Venables Building, Walton Hall, Mil- ton Keynes MK7 6AA, UK G.J. White : Rutherford Laboratory, Space Science & Technology Department, CCLRC Rutherford Ap- pleton Laboratory, Chilton, Didcot, Oxfordshire OX11 0QX, UK H. Zinnecker : Deutsches SOFIA Institut, Institut fur Raumfahrtsysteme, Universitat Stuttgart, Pfaf- fenwaldring 31, D-70569 Stuttgart, Germany H. Zinnecker : SOFIA Science Center, NASA-Ames, MS 211-3,Moffett Field, CA 94035, USA
1505.02221
1
1505
2015-05-09T01:16:52
Helium Atmospheres on Warm Neptune- and Sub-Neptune-Sized Exoplanets and Applications to GJ 436 b
[ "astro-ph.EP" ]
Warm Neptune- and sub-Neptune-sized exoplanets in orbits smaller than Mercury's are thought to have experienced extensive atmospheric evolution. Here we propose that a potential outcome of this atmospheric evolution is the formation of helium-dominated atmospheres. The hydrodynamic escape rates of Neptune- and sub-Neptune-sized exoplanets are comparable to the diffusion-limited escape rate of hydrogen, and therefore the escape is heavily affected by diffusive separation between hydrogen and helium. A helium atmosphere can thus be formed -- from a primordial hydrogen-helium atmosphere -- via atmospheric hydrodynamic escape from the planet. The helium atmosphere has very different abundances of major carbon and oxygen species from those of a hydrogen atmosphere, leading to distinctive transmission and thermal emission spectral features. In particular, the hypothesis of a helium-dominated atmosphere can explain the thermal emission spectrum of GJ 436 b, a warm Neptune-sized exoplanet, while also consistent with the transmission spectrum. This model atmosphere contains trace amounts of hydrogen, carbon, and oxygen, with the predominance of CO over CH4 as the main form of carbon. With our atmospheric evolution model, we find that if the mass of the initial atmosphere envelope is 1E-3 planetary mass, hydrodynamic escape can reduce the hydrogen abundance in the atmosphere by several orders of magnitude in ~10 billion years. Observations of exoplanet transits may thus detect signatures of helium atmospheres and probe the evolutionary history of small exoplanets.
astro-ph.EP
astro-ph
DRAFT VERSION JUNE 18, 2018 Preprint typeset using LATEX style emulateapj v. 5/2/11 HELIUM ATMOSPHERES ON WARM NEPTUNE- AND SUB-NEPTUNE-SIZED EXOPLANETS AND APPLICATIONS TO GJ 436 B RENYU HU1,2,3, SARA SEAGER4, YUK L. YUNG1,2 1Jet Propulsion Laboratory, California Institute of Technology, Pasadena, CA 91109, USA 2Division of Geological and Planetary Sciences, California Institute of Technology, Pasadena, CA 91125 3Hubble Fellow and 4Department of Earth, Atmospheric and Planetary Sciences, Massachusetts Institute of Technology, Cambridge, MA 02139, USA) Draft version June 18, 2018 ABSTRACT Warm Neptune- and sub-Neptune-sized exoplanets in orbits smaller than Mercury's are thought to have experienced extensive atmospheric evolution. Here we propose that a potential outcome of this atmospheric evolution is the formation of helium-dominated atmospheres. The hydrodynamic escape rates of Neptune- and sub-Neptune-sized exoplanets are comparable to the diffusion-limited escape rate of hydrogen, and therefore the escape is heavily affected by diffusive separation between hydrogen and helium. A helium atmosphere can thus be formed – from a primordial hydrogen-helium atmosphere – via atmospheric hydrodynamic escape from the planet. The helium atmosphere has very different abundances of major carbon and oxygen species from those of a hydrogen atmosphere, leading to distinctive transmission and thermal emission spectral features. In particular, the hypothesis of a helium-dominated atmosphere can explain the thermal emission spectrum of GJ 436 b, a warm Neptune-sized exoplanet, while also consistent with the transmission spectrum. This model atmosphere contains trace amounts of hydrogen, carbon, and oxygen, with the predominance of CO over CH4 as the main form of carbon. With our atmospheric evolution model, we find that if the mass of the initial atmosphere envelope is 10- 3 planetary mass, hydrodynamic escape can reduce the hydrogen abundance in the atmosphere by several orders of magnitude in ∼ 10 billion years. Observations of exoplanet transits may thus detect signatures of helium atmospheres and probe the evolutionary history of small exoplanets. Subject headings: radiative transfer - atmospheric effects - planetary systems - techniques: spectroscopic - planets and satellites: individual (GJ 436 b) 1. INTRODUCTION Recent exoplanet surveys have discovered warm Neptune- and sub-Neptune-sized planets in tightly bound orbits around their parent stars (Butler et al. 2004; Fressin et al. 2013; Howard 2013). Were these planets in our Solar System, their orbits would be interior to that of Mercury. The atmospheres of Neptune-sized exoplanets are assumed to be similar to those of giant planets in our Solar System, i.e., dominated by hydrogen and helium (Rogers & Seager 2010; Nettelmann et al. 2010; Madhusudhan & Seager 2011; Line et al. 2014). However, this assumption may not be valid for warm extrasolar Neptunes in close-in orbits. Due to stellar irradiation, these planets may have experienced significant at- mospheric evolution (Lopez et al. 2012; Owen & Wu 2013). Here we consider if some Neptune- and sub-Neptune-sized exoplanets can have their atmospheres depleted in hydro- gen but abundant in helium, due to extensive atmosphere loss. Figure 1 illustrates a potential evolution scenario for an exoplanet that leads to an atmosphere dominated by he- lium. The planet starts with a hydrogen-helium atmosphere accreted from the planet-forming nebula. This initial atmo- sphere may evaporate and reduce its mass by accretion heat- ing, impact erosion, stellar winds, and X-ray and extreme ul- traviolet (EUV) radiation from the parent star. Subsequently, the planet may continue to experience atmospheric loss due to stellar irradiation (Lammer et al. 2003; Erkaev et al. 2007; Baraffe et al. 2008; Lopez et al. 2012; Owen & Wu 2013; Inamdar & Schlichting 2014). If the escape rate is very high, the planet may lose both hydrogen and helium com- [email protected] 2015. All rights reserved. pletely. However, if the escape rate is comparable to the diffusion-limited escape rate of hydrogen, the escaping flow may be highly enriched in hydrogen that is 4 times lighter than helium. We show later in this paper that Neptune- and sub-Neptune-sized exoplanets undergoing transonic hydrody- namic escape would generally have an escape rate comparable to the diffusion-limited escape rate of hydrogen, and the frac- tion of hydrogen in their atmospheres would decrease as the planets evolve. Our model suggests that if the initial mass of the atmosphere is 10- 3 planetary mass, the remaining atmo- sphere would become helium-dominated within ∼ 10 billion years. This concept of fossil helium atmosphere can be important because sub-Neptune-sized planets are ubiquitous in our in- terstellar neighborhood (e.g. Howard 2013), and because he- lium atmospheres would have very different molecular com- positions from hydrogen atmospheres, leading to distinctive spectral features in transmission and thermal emission. The chemical compositions and spectral features of the atmo- spheres on super Earths and mini Neptunes are highly sen- sitive to the elemental abundances of the atmospheres, in par- ticular to the hydrogen abundance and the carbon to oxygen (C/O) ratio (Moses et al. 2013; Hu & Seager 2014). Com- pared to other non-H2-dominated exoplanet atmospheres, a helium atmosphere has a more extended scale height, and presents a better opportunity to characterize highly evolved exoplanet atmospheres via transmission spectroscopy. The paper is organized as follows. We describe our models in § 2. We develop an atmosphere evolution model to study the conditions to form helium atmospheres on Neptune- and sub-Neptune-sized exoplanets, and upgrade an existing atmo- 2 Hu et al. spheric chemistry and radiative transfer model for exoplanets (Hu et al. 2012, 2013; Hu & Seager 2014) to include the ca- pability of treating helium atmospheres. We present a gen- eral result of the fractionation between hydrogen and helium for exoplanets undergoing transonic hydrodynamic escape in § 3. We then apply our hypothesis to the Neptune-sized exo- planet GJ 436 b in § 4, and propose that the planet can have a helium atmosphere, which explains its puzzling emission features observed with Spitzer. We discuss how transmission spectroscopy of exoplanets can distinguish highly evolved, helium-dominated atmospheres from hydrogen atmospheres in § 5, and we conclude in § 6. 2. MODEL 2.1. Atmosphere Evolution Model We have created an atmosphere evolution model to inves- tigate whether a primordial hydrogen-helium atmosphere on short-period exoplanets can evolve to a helium-dominated one. We trace the atmosphere mass and elemental abundances from the present time to the past, focusing on the effects of at- mosphere escape. This way, our model can be used to test if a helium atmosphere is a plausible evolutionary outcome for an exoplanet. We truncate the model at an age of 0.1 bil- lion years, because the processes dominating the period less than 0.1 billion years after the planet formed (e.g., accretion heating, impact erosion, rapid photoevaporation) do not frac- tionate hydrogen versus helium. To simplify the calculation, we assume that the core to be unchanged in mass, size, and composition in the evolution history, and that the atmosphere is isothermal. The core could cool down and shrink by up to 1 R⊕ over the history (Baraffe et al. 2008). We have verified that the evolution history of the atmospheric hydrogen abun- dance is not qualitatively sensitive to these assumptions. For is escape of A key element of the evolution model irra- hydrogen and helium from the envelope. diated exoplanets, the escape is mainly driven by the stellar EUV radiation (e.g. Lammer et al. 2003; Yelle 2004; Lecavelier des Etangs et al. 2004; Tian et al. 2005; García Muñoz 2007; Murray-Clay et al. 2009; Guo 2013; Koskinen et al. 2013; Lammer et al. 2013), and the escape rate can be estimated by an energy-limited escape formula that relates the EUV irradiation to adiabatic expansion and cool- ing of the outflow, with an efficiency (η) that umbrellas the underlying complex energetic and hydrodynamic processes (e.g. Erkaev et al. 2007). Typical values for η vary from a few percent to a few tens percent and may change during the evolution depending on the composition and the ioniza- tion degree of the flow (Murray-Clay et al. 2009; Owen & Wu 2013). It has also been suggested that X-ray irradiation may drive escape very early in the history when the stellar X- ray luminosity is high (Cecchi-Pestellini et al. 2006, 2009; Owen & Jackson 2012). However, X-ray-driven escape is less important for our purpose than EUV-driven escape be- cause much of the X-ray energy deposition is reradiated away and does not drive escape (Owen & Jackson 2012), and be- cause the X-ray flux is one-order-of magnitude smaller than the EUV flux after 1 billion years since formation (see be- low). The energy-limited escape rate is ΦEL = LEUVηa2r3 p 4Kd2GMp , (1) in which LEUV is the luminosity of the star in EUV, η is the heating efficiency, or the fraction of absorbed EUV energy that drives escape, a is the ratio between the planet's EUV absorbing radius and the planetary radius, rp is the planetary radius, Mp is the planetary mass, d is the semi-major axis, G is the gravitational constant, and K is the potential energy reduction factor due to the Roche lobe effect (Erkaev et al. 2007). For convenience we define Q as the incident EUV heating rate on a planet, i.e. Q = LEUVr2 p 4d2 , (2) and Qnet = ηQ as the net EUV heating rate that drives escape. For LEUV, we can adopt the following empirical estimates for stellar EUV fluxes based on X-ray observations and coro- nal modeling of 80 stars with spectral types spanning from M to F (Sanz-Forcada et al. 2011) log(LEUV) = 22.12 - 1.24 log(τ), log(LX- ray) = 21.28 - 1.44 log(τ), (3) (4) where the luminosity has a unit of J s- 1 and the age (τ) has a unit of billion year. This estimate of the EUV luminosity has an uncertainty of an order of magnitude for any individ- ual star; however, we will see later that our result does not sensitively rely on the EUV luminosity. Equations (3) and (4) indicate that the X-ray luminosity can only be a minor contri- bution to the escape energy for the bulk part of the evolution. The planetary radius is defined to be the radius of the ho- mopause, i.e., the level above which the binary diffusion co- efficient between hydrogen and helium is greater than the eddy diffusion coefficient. Such a definition is only nomi- nal because significant hydrogen photochemistry may occur, and the mixing ratio of H is not a constant below the ho- mopause (e.g. Liang et al. 2003; Yelle 2004; Hu et al. 2012; Koskinen et al. 2013). For an eddy diffusion coefficient rang- ing in 109 ∼ 1011 cm2 s- 1 (Parmentier et al. 2013), the ho- mopause is at 10- 3 ∼ 10- 5 Pascal for a temperature of 600 K, and 10- 1 ∼ 10- 3 Pascal for a temperature of 104 K. Most of the EUV irradiation is deposited at the τ = 1 pres- sure level (Yung & Demore 1999), P1 = n1kT ∼ 1 σH kT ∼ mHg σ , (5) where H is the scale height, mH is the mass of hydrogen, g is the gravitational acceleration, and σ = 2 ∼ 6 ×10- 18 cm2 is the cross section of hydrogen for the EUV irradiation in 13.6-20 eV (Spitzer 1978). For GJ 436 b (g ∼ 10 m s- 2), the pressure level is 3 ∼ 8 × 10- 5 Pascal. Therefore, the EUV absorbing altitude is up to 9 scale heights above the homopause, which corresponds to less than 0.1 planetary radii. In other words, a < 1.1. We therefore assume a = 1 in this work, effectively assuming the uncertainty in a2 in Equation (1) is absorbed in a wide range of η. The parameter a being close to unity is also verified by full thermosphere models (e.g. Yelle 2004; Murray-Clay et al. 2009; Koskinen et al. 2013). We note that this assumption does not imply EUV heating occurs at the homopause, and the parameter a can be greater for smaller planets (e.g. Lammer et al. 2013). In order to apply Equation (1) to calculate the escape rate, we need to determine the regime for the hydrodynamic out- flow. Hydrodynamic outflow from planetary atmospheres powered by external heating has recently been calculated by Monte Carlo direct simulations, and the advantage of these Initial state H2 He CH4 H2O H2O CH4 H2-He H2O Helium Atmosphere on Exoplanet 3 After rapid evaporation and impact removal Asymptotic state H2 H2O CH4 H2-He H2O CO He H2O t = 0 t ~ 0.1 Gyr t ~ several Gyr Figure 1. An evolution scenario for an irradiated Neptune- or sub-Neptune-sized exoplanet that produces a helium-dominated atmosphere from a primordial hydrogen-helium atmosphere. Rapid evaporation could last for ∼ 0.1 Gyr, depending on the thermal history of the planet and the evolution of the stellar X-ray and EUV radiation, and the subsequent thermal escape modulated by diffusive separation between hydrogen and helium could last for multiple billion years to form a helium-dominated atmosphere. The required time depends on the mass of the initial atmosphere. simulations is that they no longer require outer boundary con- ditions and that they self-consistently determine the location of the Knudsen layer between the static and the hydrodynamic parts of the atmosphere (Tucker et al. 2012; Erwin et al. 2013; Johnson et al. 2013; Volkov & Johnson 2013). These sim- ulations find the energy-limited escape rate to be a good approximation of the escape rate when the escape flow is subsonic, but not when the escape flow becomes transonic. The minimum heating rate to drive a transonic outflow is (Johnson et al. 2013) Qnet > Qc ∼ 4πr∗ γ ccσcKnmr 2U(r∗) m U(rp), (6) where Qc is the critical heating rate, r∗ is the sonic point ra- dius, γ is the heat capacity ratio, ccσc is the collisional cross section, Knm is the maximum Knudsen number for the flow to be in the continuum regime, U is the gravitational energy of the escaping particle, and m is the mass of the particle. For an atomic flow, we adopt γ = 5/3, ccσc = 5 × 10- 20 m2, Knm ∼ 1 (valid when the heat is primarily absorbed over a broad range of radius below r∗), and r∗ ∼ r0. Note that r∗ can be greater than rp, but Qc only weakly depends on its specific value as r1/2 ∗ . In the transonic escape regime, the escape rate no longer increases with the energy put; instead, incremental energy input would be taken away as thermal and translational en- ergy (Johnson et al. 2013). Therefore, Equation (1) can sig- nificantly overestimate the escape rate for transonic flow. We instead estimate the escape rate as Φ = fr ΦEL, (7) where fr, adopted from Johnson et al. (2013), is the escape rate reduction factor due to the flow being transonic. fr is a function of Qnet/Qc: when Qnet/Qc < 1, fr ∼ 1 and we return to the conventional energy-limited escape; however, when Qnet/Qc > 1, fr < 1 and it decreases nearly as inverse- proportionally to a greater Qnet/Qc. Equation (7) allows us to trace the history of atmosphere loss in terms of the total mass loss flux. Next, in order to trace the effect of atmosphere loss on the atmosphere's elemental abundance, we need to partition the flux into the mass flux of escaping hydrogen and the mass flux of escaping helium, i.e., Φ = ΦH + ΦHe = 4πr2 p(φHmH + φHemHe), (8) where φH and φHe denote the number fluxes, and mH and mHe denote the masses of hydrogen and helium atom, respec- tively. The escaping flow should be made of atomic hydrogen and helium, because photochemistry models show that the dominant form of hydrogen in the upper atmosphere should be atomic hydrogen, due to photodissociation of hydrogen molecules catalyzed by water vapor (Liang et al. 2003; Yelle 2004; Moses et al. 2011; Hu et al. 2012). The physical reason for different ΦH and ΦHe is that the he- lium atom is 4 times more massive than hydrogen and that both fluids are subject to the gravity of the planet. When Φ is large, the coupling between the two components is strong and therefore the two fluxes are approximately proportional to the mixing ratios of the gases at the homopause, which re- sults in little fractionation. When Φ is small enough helium will not escape and hydrogen escape will be in the so-called "diffusion-limited" escape regime. The nature of mass frac- tionation in hydrodynamic escape has been investigated thor- oughly with hydrodynamic formulations (Zahnle & Kasting 1986; Hunten et al. 1987; Zahnle et al. 1990). Two complications arise as we apply the classical calcula- tion of mass fractionation to irradiated exoplanets. We have shown earlier that the homopause can be quite close to the EUV absorption radius, which means that (1) the tempera- ture of the homopause can be much higher than the planetary equilibrium temperature due to EUV heating; and (2) the out- flow can be partially ionized and the ion-neutral interactions should be considered in assessing whether escaping hydrogen can drag helium. The rate of momentum exchange between hydrogen and he- lium can be written as (Schunk & Nagy 1980) δMHe δt = nHe(uH - uHe)nH kT b + nHe(uH+ - uHe)nH+ mH+v nHe , (9) where n denotes the number density, u denotes the veloc- ity, b = 1.04 × 1018T 0.732 cm- 1 s- 1 is the binary diffusion coefficient between hydrogen and helium (Mason & Marrero 1970), and v is the ion-neutral momentum transfer collision frequency (Schunk & Nagy 1980). The first term is the mo- mentum exchange between neutral hydrogen and helium, and 10-41 10-42 4 ] 1 - s 3 m g k [ g n i l p u o C H+ - He H - H e 10-43 103 Temperature [K] 104 Figure 2. Comparison between the H-He coupling and the H+-He coupling. The H-He coupling (the term kT /b in Eq. 9) is calculated from the binary diffusion coefficient (Mason & Marrero 1970), and the H+-He coupling (the term mH+ v/nHe in Eq. 9) is calculated from the ion-neutral momentum trans- fer collision frequency (Schunk & Nagy 1980). the second term is the momentum exchange between ionized hydrogen and helium. Helium may also be partially ionized, but the ionization fraction of helium is low up to large radius (e.g. Koskinen et al. 2013). The relative efficiency between the two exchange processes is shown in Figure 2. We see that ionized hydrogen or heated hydrogen can better drag he- lium than cold neutral hydrogen, both by up to a factor of 2. Assuming uH+ = uH, the rate of momentum exchange can be written as δMHe δt = nHe(uH - uHe)(nH + nH+) kT b′ , (10) where b′ is the effective binary diffusion coefficient, defined as kT b′ ≡ (1 - x) kT b + x mH+v nHe , (11) where x is the ionization fraction. As such, the diffusion- limited escape rate is φDL = GMp(mHe - mH)b′ r2 pkT , (12) where T is the temperature of the homopause. The fact that the efficiencies of the neutral interaction and the ion-neutral interaction are similar and only weakly depend on temperature (Figure 2) allows us to choose a representa- tive temperature and ionization fraction. As a conservative estimate for the mass fractionation effect, we use a high tem- perature of 104 K, and an ionization fraction of 0.1, which yields b′ = 8.0 × 1020 cm- 1 s- 1 based on Equation (11) The diffusion-limited escape rate determines the difference between the two escaping components, as φHe XHe = φH XH - φDL. (13) This simple relation is derived assuming subsonic flow but is proved to be a close approximation to a transonic flow as well Hu et al. (Zahnle et al. 1990). This equation demonstrates that when φH ≫ XHφDL, fractionation between H and He would be min- imal, but when φH → XHφDL, φHe → 0, we return to the classi- cal formula for diffusion-limited escape. Near this limit, both H and He escape, so the flow is not exactly diffusion-limited escape, but significant fractionation between H and He can still occur. With Equations (8 and 13) we determine the es- cape flux of hydrogen and helium. The solution is If Φ ≤ φDLXHmH4πr2 p, ΦH = Φ, ΦHe = 0; If Φ > φDLXHmH4πr2 p, ΦH = ΦHe = ΦmHXH + φDLmHmHeXHXHe4πr2 mHXH + mHeXHe , ΦmHeXHe - φDLmHmHeXHXHe4πr2 p p mHXH + mHeXHe (14) . (15) For each time step, the mass of the atmosphere is updated by adding the escape fluxes of hydrogen and helium, and the composition and the mean molecular mass of the atmo- sphere are updated accordingly. Then, the new mass and mean molecular mass are used to determine the new thickness of the atmosphere and the new radius of the planet. 2.2. Atmosphere Chemistry Model a We use Atmospheric escape and evolution can have large effects on the atmospheric chemical composition and then a planet's transmission and thermal emission spectrum, by depleting hydrogen over a long period. The fact that vertical mixing (∼years) is much faster than atmospheric evolution (∼billion years) allows us to treat the atmospheric chemistry calcula- tions as "snapshots in the evolutionary history", with the in- gredients of the atmosphere (i.e., the elemental abundances) controlled by the evolution. one-dimensional photochemistry-thermochemistry photochemistry- thermochemistry kinetic-transport model (Hu et al. 2012, 2013; Hu & Seager 2014) to explore the chemical state and spectral characteristics of a potential helium atmosphere. The kinetic-transport atmosphere model computes the atmospheric composition and the opacities self-consistently for an exoplanet, with the atmospheric elemental abundance, the efficiency of vertical mixing, and the internal heat flux as the input parameters. The model also computes the temperature profiles from the opacities using the grey-atmosphere approximation. Our atmosphere chemistry model contains two steps of computation: the first step is to compute the atmospheric composition at thermochemical equilibrium self-consistently with the temperature profile; and the second step is to use the result of the first step as the initial condition to simulate the effects of vertical mixing and photochemical processes on the atmospheric composition, with the temperature-pressure profile adjusted accordingly. After the photochemistry- thermochemistry simulation converges to a steady state, we compute synthesized spectra of the modeled exoplanet's atmospheric transmission and thermal emission with a line-by-line method. A unique feature in our photochemistry- thermochemistry model is that the model does not require specification of the main component of the atmosphere (nor the mean molecular mass) and instead the model takes the elemental abundances as the input parameters, making our model ideal for exploring the composition of hydrogen- Helium Atmosphere on Exoplanet 5 versus helium-dominated atmospheres. Our general atmosphere chemistry model and its val- (Hu et al. 2012, idation have been described in detail 2013; Hu & Seager 2014). To properly treat a helium- dominated atmosphere, we have made the following up- grade to the general model. We have included the H2- He collision-induced absorption (Borysow et al. 1988) in our calculations of the opacities, and used the heat capacity of helium recommend by the NIST Chemistry Webbook (http://webbook.nist.gov/chemistry/) for calculation of the adiabatic lapse rate of helium-dominated atmospheres. We have treated helium as a chemically inert gas in our model, and included helium in calculation of the mean molecular mass. 3. FRACTIONATION BETWEEN HYDROGEN AND HELIUM We study whether hydrodynamic escape of hydrogen can fractionate hydrogen and helium in an exoplanet atmosphere. First, most of the detected exoplanets receive enough EUV radiation to have transonic hydrodynamic escape. Figure 3 shows that a majority of the detected exoplanets with their masses and radii measured would undergo transonic hydrody- namic escape for a moderate heating efficiency of ∼ 10% at an age of 10 billion years. When they are younger and the stellar EUV luminosities are higher, the planets are even more likely to experience transonic escape. This is reasonable because most of the planets have short orbital periods and locate close to their parent stars. Therefore, the transonic regime must be considered when estimating the hydrodynamic escape rate of the short-period exoplanets. Second, hydrodynamic escape from many Neptune- and sub-Neptune-sized exoplanets significantly fractionates hy- drogen versus helium in their atmospheres. Assuming tran- sonic hydrodynamic escape, we can replace Qnet by Qc in the energy-limited escape rate calculation for an estimate of the upper-limited of the escape mass flux, because increas- ing the stellar EUV flux would no longer increase the escape mass flux for transonic outflow (Johnson et al. 2013). Note that this upper limit is independent from the stellar irradia- tion, but depends on the mass and the radius of the planet, as well as the composition of the atmosphere. We then partition this mass flux to the escape fluxes of hydrogen and helium, using Equations (14 and 15), and calculate the fractionation factor between hydrogen and helium, x2, which is defined as Transonic escape possible 70 60 50 40 30 20 10 l s t e n a p f o r e b m u N 0 -4 10 Q c 0 10 1 10 2 10 /Q, i.e. the mininum heating efficiency to drive transonic escape -1 10 -3 10 -2 10 Figure 3. The minimum heating efficiency to drive transonic hydrodynamic escape from detected exoplanets, shown as a histogram. The exoplanets hav- ing their masses and radii measured more precisely than 3 - σ are counted, and the total number of planets are 254 (http://exoplanet.eu). We use Equa- tion (2) to estimate the incident EUV heating rate, and assume a conservative 10-billion-year-old stellar EUV luminosity from Equation (3). We use Equa- tion (6) to estimate the critical heating rate for transonic escape, assuming that the sonic point radius is close to the planetary radius, and that the mean molecular mass is 2 (i.e. 67% H and 33% He). principle remove hydrogen in the atmosphere to an arbitrarily low abundance. Therefore, fractionation between hydrogen and helium is significant for hydrodynamic escape from many exoplanets. This does not imply these planets have helium atmospheres, however, as additional conditions have to be met. Our at- mosphere evolution model serves to explore the conditions to form helium atmospheres for specific planets. Neptune- and sub-Neptune-sized planets can have small hydrogen and helium envelopes, and the fractionation factors for them are substantially smaller than unity. Note that the fractionation factors for them are not zero, so they are technically not in the regime of diffusion-limited escape. Rather, hydrodynamic escape from these planets is highly affected by the diffusion separation between hydrogen and helium. We suggest that fractionation between hydrogen and helium is an important process for hydrogen-helium atmospheres of short-period ex- oplanets undergoing hydrodynamic escape. The effect of this fractionation, as later discussed for the cases of GJ 436 b, de- pends on the initial mass of the atmosphere, and is likely to be significant for Neptune-sized and smaller exoplanets. x2 = φHe/φH XHe/XH . (16) 4. A HELIUM ATMOSPHERE MODEL FOR GJ 436 B x2 depends on the instantaneous volume mixing ratio between hydrogen and helium, and takes a value between 0 and 1. x2 = 1 means no fractionation, and x2 = 0 means the highest degree of fractionation (i.e., no helium escaping with hydrogen). We show the results in Figure 4. Most of the Neptune- and sub-Neptune-sized planets have a small fractionation factor between hydrogen and helium. This means that hydrodynamic escape from their atmospheres would deplete hydrogen and enrich helium. Since the as- sumed escape rate is an upper limit, we are conservative in estimating this fractionation effect. For a smaller escape rate, the fractionation effect would be even more significant. We also note that the fractionation effect becomes milder as the hydrogen mixing ratio decreases. However, the asymptotic value of x2 is less than unity (Figure 4). This implies that with unlimited time and no hydrogen source, the process can in The observational characterization of atmospheric com- positions of Neptune- and sub-Neptune-sized exoplan- ets is ongoing (e.g. Seager & Deming 2010), and pro- vides the opportunity to search for highly evolved exo- planet atmospheres and test our hypothesis of the forma- tion of helium atmospheres. One of the most observed Neptune-sized exoplanets is GJ 436 b (Gillon et al. 2007; Deming et al. 2007; Demory et al. 2007; Alonso et al. 2008; Pont et al. 2009; Cáceres et al. 2009; Ballard et al. 2010; Stevenson et al. 2010; Beaulieu et al. 2011; Knutson et al. 2011, 2014a; Lanotte et al. 2014; Morello et al. 2015). This planet orbits a nearby M dwarf star and has a mass of 23.2 M⊕ and a radius of 4.22 R⊕ (Torres et al. 2008). The bulk density of the planet implies that the planet should have a thick atmo- sphere or envelope made of hydrogen and/or helium, which itself has a mass of at least 10- 4 ∼ 10- 3 of the planet's mass (Nettelmann et al. 2010). 6 Hu et al. processes (Line et al. 2011), atmospheric chemistry models uniformly predict most carbon should be in the form of CH4–not CO–in a solar-abundance hydrogen and helium background atmosphere. A solar-abundance hydrogen and helium atmosphere on GJ 436 b cannot produce a spectrum consistent with either dataset of the spectra (Stevenson et al. 2010; Lanotte et al. 2014), because the chemical equilibrium models would pre- dict most of carbon to be in the form of CH4, but the observations indicate otherwise. Including disequilibrium processes like strong vertical transport and efficient pho- tochemical reactions, or exploration of temperature vari- ation in reasonable ranges, cannot address the "missing- methane" problem (Madhusudhan & Seager 2011; Line et al. the emission spec- 2011; Agúndez et al. 2014). trum of Stevenson et al. (2010), the mixing ratio of CH4 between 1 and 0.01 bar would need to below ∼ 1 ppm based on retrieval studies allowing the compositions and the temperature profiles to vary freely and independently (Madhusudhan & Seager 2011; Moses et al. 2013). The retrieval results of fitting to the more recent dataset of Lanotte et al. (2014) are not available, but one could imagine that more CH4 could be allowed in the atmosphere because of the smaller 3.6-µm eclipse depth. To fit Moses et al. (2013) presented a series of models for the at- mospheric compositions and emission spectra of GJ 436 b, using the PHOENIX atmospheric model (Barman et al. 2005) for the temperature-pressure profiles. To obtain a mixing ratio of methane less than 1 ppm, a metallicity 10,000 times greater than the solar atmosphere would be required (Moses et al. 2013, Figure 12). Such a 10,000x solar atmosphere would be CO2-, CO-, N2- and H2O-dominated, and would have a very high mean molecular mass. However, this hypothetical world cannot be GJ 436 b, for such an extravagant metallicity violates the constraints on the bulk density of the planet in question. The ra- dius of an equally massive planet with a CO-, CO2- and H2O-dominated atmosphere would be much smaller than the measured radius of GJ 436 b. Detailed models of the interior structures of Neptune-sized exoplanets have also ruled out such a dense outer envelope (Rogers & Seager 2010; Nettelmann et al. 2010). In particular, Nettelmann et al. (2010) calculated a high-temperature endmember scenario, in which the temperature profile was assumed to be adiabatic and the temperature at 1 bar was assumed to be 1300 K (se- ries G). This scenario had a temperature as high as 5500 K at 1000 bar, but the minimum mass of hydrogen and helium for this model was still at least 0.01% of the planetary mass. A 10,000x solar outer envelope would not be permitted even when a fairly high internal temperature is considered. In addition to raising the metallicities, Moses et al. (2013) then assumed an ad hoc temperature profile, and proposed that an H2- and He-dominated atmosphere 300 times more metal-rich than the solar atmosphere could be consistent with the emission spectrum reported by Stevenson et al. (2010). To produce a sufficiently low mixing ratio of CH4, the tem- perature of such an atmosphere at 0.1 bar would have to be ∼ 1300 K, greater than radiative-convective model predic- tions by at least 200 K. In other words, the 300x solar model of Moses et al. (2013) likely assumed an unphysically large tem- perature at 0.1 bar. The tidal heat due to the planet's orbital eccentricity (e=0.16) is deposited at a much deeper pressure and cannot raise the temperature at 0.1 bar (Agúndez et al. 2014). When using reasonable temperature profiles, raising Figure 4. Fractionation factor between hydrogen and helium for Neptune- and sub-Neptune-sized exoplanets undergoing transonic hydrodynamic es- cape. The color contour shows the fractionation factor (x2), for which a smaller value means a more significant fractionation between hydrogen and helium. Neptune- and sub-Neptune-sized exoplanets having their masses and radii measured more precisely than 3 - σ are shown. 4.1. Puzzling Emission Features of GJ 436 b The composition of GJ 436 b's atmosphere has been a puzzle. A broad-band emission spectrum of the planet in 3-30 µm was obtained from Spitzer (Stevenson et al. 2010; Lanotte et al. 2014), which indicated that the planet's atmosphere is poor in CH4 and rich in CO (Madhusudhan & Seager 2011; Line et al. 2014). The key spectral feature is a detected emission flux in the 3.6 µm band (CH4 absorption band) and non-detection in the 4.5 µm band (CO absorption band). We note that the reported 3.6 µm emission flux of Lanotte et al. (2014) was significantly lower than that of Stevenson et al. (2010), though both values were derived from the same observation (see Appendix A for discussion). In addition, a recent transmission spectrum of the planet in 1-2 µm was obtained from HST, and the spectrum lacks H2O or CH4 features (Knutson et al. 2014a). These observations and their interpretation lead to a theo- retical challenge: even when accounting for disequilibrium Helium Atmosphere on Exoplanet 3 2.5 ] r a B [ e r u s s e r P 2 ] 3 − 0 1 [ F / r a t s 1.5 t e n a p l F 1 −4 10 −3 10 −2 10 −1 10 0 10 1 10 2 10 500 1000 Temperature [K] 1500 2000 0.5 0 2 3 5 7 9 11 Wavelength [microns] the metallicities to the extent still allowed by the planet's bulk density cannot reduce the mixing ratio of CH4 in the observ- able part of the atmosphere to less than 1 ppm. We also note that some of the self-consistent models for 300x and 1000x solar abundances in Moses et al. (2013), without assuming an ad hoc temperature profile, appear to be consistent with the more recent dataset of the emission spec- trum (Lanotte et al. 2014). Alternatively, one could explain GJ 436 b's emission spec- trum by the aforementioned scenario of atmospheric evolution culminating in a helium-dominated atmosphere. Once such a state is achieved, the scarcity of atmospheric hydrogen will cause the main molecular carrier of carbon to be CO rather than CH4, explaining the emission spectrum. 4.2. Chemistry and Spectrum Using the atmosphere chemistry model, we explore a wide range of parameters that define the atmospheric composition of GJ 436 b, including the hydrogen elemental abundance of the atmosphere (by number, denoted as XH), the metal- licity of the atmosphere (denoted as XM), the carbon to oxy- gen ratio of the atmosphere (denoted as XC/XO), the intrin- sic temperature (to specify the internal heat flux, denoted as Tint), and the eddy diffusivity (to specify the efficiency of vertical mixing, denoted as Kzz). In this work we consider only H, He, C, and O species, so that XHe = 1 - XH - XM and XM = XC + XO. Nitrogen, sulfur, and other metal species may also contribute to the spectrum. Our selection suffices because the key observational feature to explain is lack of methane ab- sorption. For each atmosphere scenario that we consider, we use the thermochemistry-photochemistry model to compute the steady-state molecular composition from 103 bar to 10- 8 bar. We verify that the lower boundary at 103 bar is sufficient to maintain thermochemical equilibrium at the lowest atmo- sphere layer for each simulation. For the stellar input spec- trum, we use the latest HST measurement of the UV spec- trum of GJ 436 (France et al. 2013) and the NextGen simu- lated visible-wavelength spectrum (Allard et al. 1997) of an M star having parameters closest to those of GJ 436 (i.e., ef- fective temperature of 3600 K, surface gravity log(g)=5.0, and solar metallicity). We find that a helium-dominated atmosphere with a low hydrogen abundance and close-to-solar carbon and oxygen abundances can adequately fit both datasets of GJ 436 b's emission spectrum (Stevenson et al. 2010; Lanotte et al. 2014). At the same time a very different model also generally fits the more recent dataset (Lanotte et al. 2014), a hydrogen- rich atmosphere with highly super-solar carbon and oxygen abundances (Figure 5 and Figure 6). There is a strong resid- ual discrepancy between the data and both models at 16 µm, and it can be further reduced if the temperature profile is al- lowed to adjust freely (Appendix B). Both the helium atmosphere (i.e., the H2-poor model) and the H2-rich metal-rich model ensure the predominance of CO over CH4 as the main form of carbon, the key indication of the planet's emission spectrum. Our chemical models indi- cate that the atmosphere would be depleted in methane when XH < XM ∼ XC + XO (Figure 7). Under this condition, methane would have to compete with water for the limited resource of hydrogen. This condition generalizes the result of a previ- ous atmospheric chemistry model of this planet (Moses et al. 2013); instead of increasing the metallicity XM, we suggest decreasing XH could also make this condition satisfied. In 7 1100 K 900 K 700 K 500 K 14 17 21 25 30 CO CH4 CO CH4 CO2 H2O CH4 H2O CO2 H2O Figure 5. Comparison of model spectra of the dayside emission of GJ 436 b with observations. The reported values for the eclipse depths are shown in black (Stevenson et al. 2010) and orange (Lanotte et al. 2014). The filled circles are the model-predicted occultation depths with the Spitzer band- pass incorporated. The wavelength ranges of major molecular absorption bands are shown below the horizontal axis. The spectra are based on at- mosphere models with parameters tabulated in Table 1, and the correspond- ing temperature-pressure profiles are shown in the inserted panel. The thin dashed lines show the eclipse depths if the planet emits as a black body with temperatures of 500-1100 K. The red line shows the best-fit model to the data of Stevenson et al. (2010), which has XH = 10- 4 and an internal heat flux. The magenta line shows a model with the same composition but with- out the internal heat flux. The blue line shows the best-fit model to the data of Lanotte et al. (2014), and the green line shows the a model of H2-rich metal- rich atmospheres (XH ≥ 0.3, XM ≥ 0.1). The thermal emission measurement at the 3.6 µm band constrains on the potential planetary scenarios: the value reported by Stevenson et al. (2010) suggests a planet having an H2-poor at- mosphere and internal heating; but the value reported by Lanotte et al. (2014) allows both the scenarios of an H2-poor atmosphere and the scenarios of an H2-rich metal-rich atmosphere. addition, the weak absorption features at the 5.8-µm and 8.0- µm bands seen in both datasets suggest that the atmosphere should still have H2O and CH4, and lead to a lower bound of XH (Figure 7). Both the He atmosphere and the H2-rich atmosphere sce- narios require an aerosol layer at ∼ 1 mbar pressure to gener- ate the observed featureless transmission spectrum (Figure 8). This is because a hydrogen and/or helium atmosphere, with a mean molecular mass of 2 - 4, would produce large absorp- tion features in transmission if it does not have aerosols. We find that a thin aerosol layer having a vertical optical depth less than 0.1 at mid-infrared wavelengths could produce the featureless transmission spectrum observed at 1-2 µm, and also consistent with the broadband transit depths at 3.6, 4.5, and 8.0 µm (Figure 8). Such an aerosol layer would have little impact to the emission spectra. A thicker aerosol layer is possible and consistent with the transmission spectrum. We have assumed that the atmosphere to be free of thick clouds in the calculation of thermal emission, because the atmosphere would be too hot to form water clouds and too cold to form silicate clouds (Sudarsky et al. 2003). Even if thick clouds were present, the detected emission flux in the 3.6 µm band and non-detection in the 4.5 µm band still require the main molecular carrier of carbon to be CO and not CH4. The large emission flux at 3.6 µm reported by 8 Hu et al. The parameters and the modeled compositions for the atmospheric models shown in Figure 5. Table 1 Molecular Composition at 0.01-1 bar He 0.9951 0.9995 0.9990 0.76 H2 4.8 × 10- 5 1.4 × 10- 5 4.8 × 10- 4 0.17 H2O 2.5 × 10- 6 1.3 × 10- 6 2.8 × 10- 6 1.8 × 10- 2 CH4 3.1 × 10- 10 8.8 × 10- 9 6.5 × 10- 6 4.1 × 10- 4 CO 4.6 × 10- 3 4.2 × 10- 4 4.9 × 10- 4 4.7 × 10- 2 CO2 2.5 × 10- 4 5.1 × 10- 5 3.9 × 10- 6 5.1 × 10- 3 2/n χ Color 2.5a 7.1a 2.7b 3.1b r m b g Parameter XH XM Tint (K) XC/XO 0.95 0.90 0.9997 0.70 1 × 10- 2 1 × 10- 3 1 × 10- 3 1 × 10- 1 1 × 10- 4 3 × 10- 5 1 × 10- 3 3 × 10- 1 aComparison to the data of Stevenson et al. (2010) bComparison to the data of Lanotte et al. (2014) 60 20 60 60 H H O 2 CH 4 CO 2 CO H 2 He r m b g −1 −2 ) H −3 X ( g o l −4 −5 −6 −1 −2 ) 10-9 10-8 10-7 10-6 10-5 10-4 10-3 10-2 10-1 100 Mixing Ratio l ] r a B [ e r u s s e r P ] r a B [ e r u s s e r P ] r a B [ e r u s s e r P ] r a B [ e r u s s e r P 10-5 10-4 10-3 10-2 10-1 100 101 102 103 10-5 10-4 10-3 10-2 10-1 100 101 102 103 10-5 10-4 10-3 10-2 10-1 100 101 102 103 10-5 10-4 10-3 10-2 10-1 100 101 102 103 10-10 Figure 6. Modeled compositions of a helium atmosphere on GJ 436 b. The mixing ratios of important species are shown as a function of pressure. Each color corresponds to one species, and each panel corresponds to a model shown in Figure 5 and summarized in Table 1. The panel label indicates the color of the synthesized spectrum in Figure 5. In the third panel ("b"), H2 and CO lines overlap because their mixing ratios are very close. Stevenson et al. (2010) corresponds to a brightness tem- perature greater than 1100 K. With our model atmosphere, this thermal emission emerges from the deep atmosphere (Figure 5), which appears to be only possible for a clear helium-dominated atmosphere highly depleted in hydrogen and having close-to-solar abundances of carbon and oxygen. For such an atmosphere, the mixing ratio of methane is extremely low, and there is essentially little opacity at 3.6 µm, thereby creating an infrared window into the deep atmo- sphere. An internal heat flux ∼ 10 times Earth's geothermal flux may thus drive a deep adiabatic temperature profile and lead to a high emission flux in this window region. In our best-fit model, the emission of a clear helium-dominated atmosphere having a very low amount of methane can show high brightness temperatures between 3.6 and 4.0 µm, ranging from 1100 to 1400 K (Figure 5). Comparing with the temperature-pressure profile, this corresponds to emitting from as deep as 1-100 bars at these wavelengths. Probing such deep atmospheres is probably one of the unique features of the proposed helium-dominated atmosphere. For this reason, a small energy flux from the interior that drives a deep adiabatic temperature profile could lead to appreciable changes at specific wavelengths of the emerging infrared spectrum. We note that the grey-atmosphere approximation does not Fitting to Stevenson et al. (2010) χ2/n 9 Fitting to Lanotte et al. (2014) 8 7 6 5 4 3 9 8 7 6 5 4 3 H X ( g o −3 −4 −5 −6 −3 −2.5 −2 log(X C+X −1.5 O) −1 −3 −2.5 −2 −1.5 −1 O) C/X log(1−X −0.5 Figure 7. Constraints on the composition of the atmosphere on GJ 436 b. The color contours show the goodness of fit by comparing atmospheric mod- els to the dayside emission (Stevenson et al. 2010; Lanotte et al. 2014). The 2/n, where n = 6 is the number of obser- goodness of fit is defined as the χ vations. The axes define the mixing ratios of hydrogen, carbon, and oxygen with respect to the total number of atoms, and the mixing ratio of helium is XHe = 1 - XH - XC - XO. The axis for the carbon to oxygen ratio is shown in terms of log(1 - XC/XO) to emphasize the sensitivity when the carbon to oxygen ratio approaches 1. The upper two panels compare with the reported values of Stevenson et al. (2010) and the bottom two panels compare with the reported values of Lanotte et al. (2014). The parameters not shown are marginalized, including the eddy diffusion coefficient ranging from 106 to 109 cm2 s1, which are reasonable values for deep atmospheres according to the free-convection and mixing-length theories (Gierasch & Conrath 1985). Both datasets allow H2-poor scenarios, but the data of Lanotte et al. (2014) also allow H2-rich metal-rich scenarios. capture this infrared window, which makes our calculation of the temperature profile no longer consistent with the atmo- spheric composition. As a plausibility check, we calculate the total infrared emission flux from the planet, with and without the spikes at 3.6-4.0 µm. We find the case with a deep adiabat emits 6.6% more energy through this window than the case without a deep adiabat. This additional emission energy, if solely attributed to the internal heat flux, would correspond to an intrinsic temperature of 370 K, much higher than the 60 K assumed in the grey-atmosphere approximation. A non-grey radiative-convective model is required to accurately calculate the temperature profile and the internal heat flux indicated by the emission feature. Helium Atmosphere on Exoplanet 9 −3 x 10 7.6 7.4 7.2 7 6.8 6.6 6.4 6.2 h t p e D t i s n a r T 4 5 Wavelength [microns] 2 3 7 1012 15 20 6 0.3 0.5 0.7 1 −3 x 10 H2O CO2 7.4 7.3 7.2 7.1 7 6.9 6.8 6.7 t h p e D t i s n a r T 6.6 1 1.2 1.4 1.6 1.8 2 Wavelength [microns] Comparison of model Figure 8. transmission spectra of GJ 436 b with observations (Ballard et al. 2010; Pont et al. 2009; Alonso et al. 2008; Cáceres et al. 2009; Knutson et al. 2011, 2014a). The upper panel shows the transmission spectra from 0.3 to 20 µm and the lower panel provides a zoom- in view for the wavelength of 1-2 µm featuring recent HST observations (Knutson et al. 2014a). The model spectra are based on self-consistent at- mosphere models with parameters tabulated in Table 1, i.e., the red lines cor- responding to the best-fit model to Stevenson et al. (2010), and the blues lines corresponding to the best-fit model to Lanotte et al. (2014). The dashed lines show the transmission spectra of a clear atmosphere, dominated by molecular absorption features. The solid lines show the transmission spectra of an at- mosphere that has an aerosol layer at pressures between 1 and 100 mbar, with an aerosol opacity of 0.001 cm2 g- 1 at 3 µm. The aerosol optical property is computed for a material with a refractive index of 1.001, and a lognormal size distribution with a mean diameter of 0.1 µm and a size dispersion factor of 2, using the method detailed in Hu et al. (2013). The vertical optical depth produced by such an aerosol layer is maximally 0.1 for wavelengths longer than 3 µm, yielding no impact on the interpretation of the thermal emission spectrum. A helium-dominated atmosphere of GJ 436 b must have an aerosol layer to be consistent with the featureless transmission spectrum at 1-2 µm. One might ask whether probing such deep atmosphere is ever possible. We show that this is possible for a helium- dominated and hydrogen-poor atmosphere in terms of molec- ular absorption, because the common opacity sources at this wavelength range, including CH4 and H2-H2 collision- induced absorption, disappear in the helium-dominated atmo- sphere. H2-He collision-induced absorption is insignificant. The lack of absorbers thus provides an infrared window. One might also wonder, given that the model has a 10x solar metal- licity, whether clouds could form above ∼ 100 bar and block this emission window (Burrows & Sharp 1999). Comparing the modeled temperature profiles to the condensation temper- atures of the materials commonly suggested to form clouds in the atmospheres of irradiated exoplanets, including H2O, Mg2SiO4, MgSiO3, and Fe, we find that GJ 436 b is too hot to have water clouds, and too cold to have silicate or iron clouds above 100 bars (Sudarsky et al. 2003). A "cloud deck", if ex- isting, is then likely to be deeper than what we are concerned with. A cautionary note is that strong tidal heating may raise the temperature of the deep convective layer (Agúndez et al. 2014), and raise the altitude of this potential cloud deck, caus- ing the clouds to affect the infrared spectra. 4.3. Formation of a Helium Atmosphere 4.3.1. Current Escape GJ 436 b is undergoing transonic hydrodynamic escape. The total EUV flux received by GJ 436 b is 7.3 × 1015 W esti- mated from Equations (3), and this estimate is fully consistent with X-ray and Ly-α flux measurements of GJ 436 (Appendix C). Assuming γ = 5/3 for an atomic flow, ccσc = 5 × 10- 20 m2, Knm ∼ 1 (valid when the heat is primarily absorbed over a broad range of radius below r∗), and r∗ ∼ r0, we estimate the critical heating rate to be Qc = 6.5 × 1014 W for a helium- dominated atmosphere, and 2.3 × 1014 W for a hydrogen- dominated atmosphere from Equation (6). We find that as long as η > 0.09, the heating rate would be greater than the critical rate and the current hydrodynamic escape would be transonic. Since the earlier EUV fluxes have been greater than the current flux, hydrodynamic escape on GJ 436 b would have always been transonic. Our calculation of the escape rate is consistent with the re- cent HST observation of the GJ 436 system in the Lyman- α absorption (Kulow et al. 2014). The escape rate of H de- rived from the Ly-α transit light curve is 4 × 106 ∼ 2 × 107 g s- 1, assuming an ionization fraction of 0.1. In compari- son, the standard energy-limited escape rate of a hydrogen- dominated atmosphere calculated from Equation (1) would be 108 ∼ 1010 g s- 1 for an efficiency ranging from 0.01 to 1. Even for a very low efficiency, the standard formula overesti- mates the escape rate. In our model, the current escape rate is 7.3 × 104 ∼ 7.3 × 106 g s- 1 for H and 6.8 × 108 ∼ 6.9 × 108 g s- 1 for He, for XH ranging between 10- 4 and 10- 2, con- sistent with the observation. These calculations assume a heating efficiency of 10%, but varying the efficiency from a few % to 50% does not change the results significantly, because the escape rate is limited by Qc for the transonic flow. Here, the apparent agreement is only a proof of con- cept rather than observational confirmation, because the re- ported absorption depth is measured after the optical transit, and stellar variabilities could significantly affect the planetary signal (Loyd & France 2014; Llama & Shkolnik 2015). Also, the reported high velocity of the escaping hydrogen corona (60-120 km s- 1) does not indicate the outflow to be super- sonic by itself. The absorption at high Doppler shifts may very well be due to absorption in the wings of the line pro- file (Ben-Jaffel & Sona Hosseini 2010; Koskinen et al. 2010), or produced by charge exchange with stellar wind protons (Holmström et al. 2008; Kislyakova et al. 2014). 4.3.2. Formation Conditions We find that the fractionation factor between helium and hydrogen (x2) is ∼ 0.2 at present for a hydrogen abundance 10 (A) 0 10 −2 10 −4 10 0.25 0.2 0.15 0.1 0.05 0 −2 −2.5 −3 −3.5 H X 2 x (B) ) p M / m a t M y a d − n e s e r P t Hu et al. change of the hydrogen abundance is greater for a smaller ini- tial mass. Therefore, the timescale required to significantly reduce the hydrogen abundance from around the cosmic value to 10- 2 ∼ 10- 4 by hydrodynamic escape is sensitive to the mass of the atmosphere. For example, hydrodynamic escape for 10 billion years can fractionate an atmosphere that is 10- 3 planetary mass or less (Figure 9). Similarly, hydrodynamic escape for 5 billion years can fractionate an atmosphere that is 3 × 10- 4 planetary mass. The mass fractionation can be more efficient if the ho- mopause is well below the EUV absorption altitude and has a temperature close to the planet's equilibrium temperature. For a temperature of 600 K, additional calculations indicate a fractionation factor between helium and hydrogen of ∼ 0.1. In that case, hydrodynamic escape for 10 billion years can well fractionate an atmosphere that is 3 × 10- 3 planetary mass. The amount of hydrogen and helium accreted by the planet when it formed depends the disk dissipation time, the orbital location, and the mass of the planet's core (Rafikov 2006; Hansen & Murray 2012; Ikoma & Hori 2012; Inamdar & Schlichting 2014). If the planet formed at large orbital separation and migrated to close-in orbits, Rafikov (2006) shows that a 20-M⊕ core would only accrete ∼ 0.1 M⊕ of hydrogen and helium at 10 AU if the core has a high luminosity and the accretion is fast. This makes the ini- tial mass of the hydrogen-helium envelope 0.5% the plan- etary mass. Engulfing planetesimals or radiogenic heating would increase the core luminosity and reduce the amount of hydrogen and helium accreted. After accretion, the ini- tial atmosphere may also be reduced in mass by photoevap- oration (e.g. Lopez et al. 2012; Fortney et al. 2013). If the planet formed in situ (Hansen & Murray 2012), the amount of hydrogen and helium accreted would be on the orders of 10- 3 ∼ 10- 2 the core mass, and this initial envelope is typ- ically reduced in mass by giant impacts by 1 ∼ 2 orders of magnitudes (Inamdar & Schlichting 2014). Both formation scenarios could provide conditions to form the helium atmo- sphere. It is also important to check how long the proposed helium- dominated atmosphere would be stable against further tran- sonic hydrodynamic escape. Apparently the total mass loss for 10 billion years is on the order of 10- 3 planetary mass. Therefore, it is conceivable that a helium-dominated envelope of 10- 3 planetary mass would be stable for billions of years. Figure 9 also shows evolution scenarios in which the planet has already had a helium-dominated atmosphere ∼ 5 billion years ago, and the helium-dominated atmosphere persists to the present day. Therefore, although hydrodynamic escape might eventually deplete helium from the planet, the evolu- tionary phase when helium dominates the outer envelope can be fairly long. We therefore propose that exoplanet GJ 436 b may have a helium-dominated atmosphere that evolved from a primordial hydrogen and helium envelope. The proximity of the planet to its parent star provides the conditions to maintain transonic hydrodynamic escape throughout the evolution history. The mass and the size of the planet determine whether the escape rate has been close to the diffusion-limited escape rate of hy- drogen. As a result, the planet has experienced dispropor- tional loss of its primordial hydrogen. Some of the primordial helium has also been lost during this evolution, but some may have remained. 1 2 3 4 5 6 Age [Gyr] 7 8 9 10 log (Initial X ) Helium atmosphere possible −3 log (Present−day X −2.5 −2 ) H −1.5 −1 H −0.5 −1 −1.5 −2 −2.5 −3 −3.5 ( g o l −4 −4.5 −4 −3.5 Figure 9. Fractionation between helium and hydrogen via hydrodynamic escape of GJ 436 b, as a function of the present-day XH and the present- day mass of the atmosphere. Panel (A) shows the history of the hydrogen abundance of the atmosphere and the fractionation ratio between helium and hydrogen by escape (x2), for an atmosphere presently having a mass of 10- 3 (black lines) or 3 × 10- 4 (blue lines) planetary mass and various present-day XH distinguished by line types. The total mass loss is approximately 10- 3 planetary mass in these models. Panel (B) shows the calculated initial XH by color contours, assuming an escape efficiency of 10% and an age of 10 billion years. We have assumed the atmosphere to be isothermal with a temperature of 2000 K. Even a temperature of 1000 K yields only a very minor alteration, indicating the isothermal assumption is adequate. We caution that the inte- rior and the atmosphere may cool down and contract significantly during the first billion years, and our calculations may not capture this early phase of evolution. When the mass of the atmosphere is 10- 3 planetary mass or less, hydrodynamic escape can reduce the hydrogen abundance of the atmosphere by orders of magnitude. ranging from 0.1 to 10- 4, and has remained below ∼ 0.2 for the entire evolution history (Figure 9). The fractionation fac- tor is significantly smaller than unity, meaning that hydrody- namic escape disproportionally remove hydrogen from the at- mosphere. The cumulative effect of the selective loss of hydrogen via hydrodynamic escape depends on the mass of the atmosphere. If the initial mass of the atmosphere is 10- 3 planetary mass or less, hydrodynamic escape would decrease the hydrogen abundance of the atmosphere by more than one order of mag- nitude on a time scale of a few billion years (Figure 9). The 5. DISCUSSION: DETECTING EVOLVED EXOPLANET ATMOSPHERES VIA TRANSMISSION 2012). Helium Atmosphere on Exoplanet 11 The planetary evolution scenario for GJ 436 b can be a gen- eral process happening on many Neptune and sub-Neptune- sized exoplanets. Due to depletion of hydrogen, the abun- dances of major carbon- and oxygen-bearing species would change by orders of magnitude (Hu & Seager 2014). For ex- ample, the most important trace gases in an H2-dominated at- mosphere are CH4 and H2O at equilibrium temperatures lower than ∼ 1000 K (defined for zero albedo and full heat redis- tribution). When the planet loses hydrogen, the dominant carbon- and oxygen-bearing species change to CO, high-order hydrocarbons, CO2, and O2. With the recognition that primordial helium may persist as a planet loses hydrogen, we propose that the highly-evolved, helium-dominated atmospheres can be detected by measuring the transmission spectra of Neptune- and sub-Neptune-sized exoplanets. The helium atmosphere can be distinguished from a hydrogen atmosphere by its molecular compositions, i.e., the lack of CH4 and the dominance of CO, CO2 and 1 (see Figure 8). These gases have spectral features in O2 the visible and near-infrared wavelengths. The helium at- mosphere, with a mean molecular mass of ∼ 4, has a large scale height for the spectral features to manifest strongly in transmission spectra (Miller-Ricci et al. 2009). In fact, in comparison with recently acquired transmission spectra of Neptune- and sub-Neptune-sized exoplanets (e.g. Bean et al. 2010; Berta et al. 2012; Kreidberg et al. 2014; Fraine et al. 2014; Knutson et al. 2014a,b), observing helium atmospheres is within the reach of current observatories, amid complica- tion of clouds and aerosols masking the molecular features (e.g. Benneke & Seager 2013; Morley et al. 2013). In addition to transmission spectroscopy, could helium- dominated atmospheres be directly detected? We suggest the following three methods, taking advantage of the unique physical properties of helium. First, helium has a very low heat capacity compared to hydrogen, because a helium atom has only three degrees of freedom and a hydrogen molecule can have six. For example, at 1000 K helium has a specific heat capacity of 5.2 J g- 1 K- 1, compared to the hydrogen's of 15.1 J mol- 1 K- 1 (http://webbook.nist.gov/chemistry/). This would lead to a helium-dominated atmosphere having the greatest adiabatic temperature-pressure gradient among com- mon planetary atmospheres, given that the temperature gradi- ent of a convective atmosphere is inversely proportional to the specific heat capacity. High-resolution infrared spectroscopy in the future could potentially measure the temperature gra- dient by deep spectral features, and thereby verify the dom- inance of helium. Second, due to its low heat capacity, a helium-dominated atmosphere is more likely to have an un- shifted hot spot, i.e., a hottest zone at the substellar point, compared with a helium-dominated one. This will result in a thermal emission phase curve peaked at occultation. Third, helium is a very poor Rayleigh scatterer, with a Rayleigh scattering cross section 20 times smaller than that of H2 (Tarafdar & Vardya 1969). As a result, reflection of the stellar light by a helium-dominated atmosphere should be dominated by multiple scattering of clouds or aerosols in the atmosphere. This condition would lack the strong, broad linear polariza- tion peak at a 90◦ phase angle (Madhusudhan & Burrows 6. CONCLUSION This paper proposes a new concept that some Neptune- sized and sub-Neptune-sized exoplanets may have fossil helium-dominated atmospheres. This concept is important because sub-Neptune-sized planets are ubiquitous in our in- terstellar neighborhood, and because helium atmospheres have unique molecular compositions and can be detected via transmission spectroscopy due to an extended scale hight. The proposed helium atmospheres can be formed by atmo- spheric hydrodynamic escape driven by stellar irradiation. For hydrodynamic escape to produce a large impact on the overall atmospheric composition, the mass of the initial atmosphere must be 10- 3 planetary mass or less, and the escape rate must be close to the diffusion-limited escape rate. How restrictive the first condition is uncertain and depends on the formation processes of Neptune- and sub-Neptune-sized planets. We show that the second condition is likely to be met by many irradiated Neptune- and sub-Neptune-sized exoplanets. Applying the concept to the Neptune-sized planet GJ 436 b, we find that a helium atmosphere can be formed on a time scale of its age, if it has a small atmosphere envelope. He- lium atmosphere models appear to provide better consistency with the planet's spectra than hydrogen atmosphere models, naturally explaining the lack of CH4 absorption and the dom- inance of CO as the carbon-bearing species indicated by the thermal emission. The evolutionary scenarios presented in this paper are or- ders of magnitude in nature. Refinements are warranted for future work. In particular, we neglect the thermal evolution and the core contraction in the model, and we assume an isothermal atmospheric temperature profile for ease of cal- culation. These factors should be treated in full evolution- ary simulations (e.g. Lopez et al. 2012). We also neglect atmosphere-surface exchange. If the surface emits hydrogen at high fluxes, helium atmospheres cannot form. Therefore the evolution scenarios should also be coupled with geophys- ical calculations to fully map the potential outcomes of the evolution of short-period, low-mass exoplanets. Despite these caveats, we suggest that observations of Neptune- and sub- Neptune-sized exoplanets in the near future would test the at- mospheric evolution theories presented here, and provide im- portant clues on the origins of these alien worlds. We appreciate comments on the manuscript made by mem- bers of the Yuk Yung research group at the California Institute of Technology. R.H. thanks Robert Johnson, Jeffrey Linsky, and Edwin Kite for helpful discussion. This work has utilized the MUSCLES M-dwarf UV radiation database. Support for this work was provided by NASA through Hubble Fellowship grant #51332 awarded by the Space Telescope Science Insti- tute, which is operated by the Association of Universities for Research in Astronomy, Inc., for NASA, under contract NAS 5-26555. Y.L.Y. and R.H. (in the later phase of this work) were supported in part by an NAI Virtual Planetary Labora- tory grant NASA grant NNX09AB72G to the California In- stitute of Technology. Part of the research was carried out at the Jet Propulsion Laboratory, California Institute of Technol- ogy, under a contract with the National Aeronautics and Space Administration. 1 The transmission spectra of helium atmospheres can still have prominent H2O features because H2O absorption is so strong that a small mixing ratio can lead to large features. 12 Hu et al. APPENDIX A BROAD-BAND EMISSION SPECTRUM OF GJ 436 B OBTAINED BY SPITZER A broad-band emission spectrum of GJ 436 b in 3-30 µm was obtained by measuring the eclipse depths as the planet occults its parent star by Spitzer (Stevenson et al. 2010). Recently, the same observations were re-analyzed by an independent group, and new values of the eclipse depths were reported (Lanotte et al. 2014). The two datasets are consistent in the detection of an emission flux at 3.6 µm, and the non-detection at 4.5 µm. However, the specific values and uncertainties for the emission fluxes are different between the two datasets. The more recent dataset has a significantly smaller eclipse depth at 3.6 µm, and slightly smaller eclipse depths at 5.8 µm and 8.0 µm. One should exercise caution when interpreting the reported values, as the multiple-wavelength observations of Neptune-sized exoplanets have only started and our understandings of instrumental systematic errors and our data analysis technique are still improving. The fidelity of the data may be best established by repeating the measurements and verifying that the results are consistent within the quoted errors. For GJ 436 b, this has only been done for the 4.5-µm and 8.0-µm bands. The 8-µm emission flux has been re-observed multiple times and the results of these repeated observations are consistent with each other at 1σ level (Stevenson et al. 2010; Knutson et al. 2011). Unfortunately, the measurements of the eclipse depth at 3.6 µm have not converged. The more recent analysis claimed that the 3.6-µm light curve of GJ 436 of Stevenson et al. (2010) exhibits significant variability out-of-eclipse, which may be due to incomplete removal of systematic errors. The more recent analysis appears to be able to remove this variability, but derive a much smaller eclipse depth. However, this analysis used ordinary aperture photometry assuming small apertures (Lanotte et al. 2014). It is therefore unclear how well the intrapixel sensitivity variations have been corrected in their analysis (Stevenson et al. 2012). Repeating the measurements of the eclipse depth of GJ 436 b at 3.6 µm would be necessary to pinpoint its emission flux. MODEL-OBSERVATION DISCREPANCY AT 16-MICRON A remaining challenge for GJ 436 b is the high thermal emission flux of the planet detected at 16 µm, implying little CO2 absorption in the atmosphere. Our best-fit models differ from the observation at this wavelength by 2.5 σ, and this is the dominant source of χ2 for the best-fit models (Figure 5). We have made several attempts to improve the fit. First, making the carbon to oxygen ratio unity or very close to unity does not solve the problem. Although the amount of CO2 can be decreased, the amount of CH4 will be increased for a carbon to oxygen ratio closer to unity. As the 3.6-µm measurement for CH4 has a much smaller error bar than the 16-µm measurement for CO2, our fitting still favors a solution that primarily satisfies the 3.6-µm constraint. Here we note that further increasing the internal heat flux may suppress CH4 for large carbon to oxygen ratios, and such a large heat flux might be physically plausible if tidal heating is efficient (Agúndez et al. 2014). Second, we find that modifying the temperature profile will improve the apparent goodness of fit. The results shown in this paper are based on grey-atmosphere temperature profiles with the opacities consistent with the composition. The calculation has also assumed a dry adiabatic lapse rate of helium for the irradiation-induced convective layer that the thermal emission spectrum is probing. The dry adiabatic lapse rate of helium is quite large, which amplifies the size of spectral features including the CO2 absorption feature. As a test we have altered the temperature-pressure profile gradient for pressures smaller than 0.4 bar, mimicking an adiabatic lapse rate softened by condensable species, or an inefficient day-night heat recirculation (Lewis et al. 2010). We found that the fit to the 16-µm emission flux can be improved to within 2 σ without altering the composition. EVOLUTION HISTORY OF X-RAY AND EUV FLUXES The hydrodynamic escape is driven by of X-ray and EUV radiation from the star. Although the stellar X-ray flux can be measured, the EUV flux is not directly observable due to strong interstellar absorption. For GJ 436, its EUV flux may be estimated either from the X-ray flux or from the Ly-α flux. The X-ray flux of GJ 436 has been measured to be log(LX- ray) = 20.16 (Poppenhaeger et al. 2010) or log(LX- ray) = 18.96 (Sanz-Forcada et al. 2011), in the unit of J s- 1. Using a fitting relationship between the X-ray flux and the EUV flux (Sanz-Forcada et al. 2011), we estimate an EUV flux of log(LEUV) = 20.13, but this value could vary by 3 orders of magnitude due to uncertainties in the fitting relationship. Alternatively, one could estimate the EUV flux from the Ly-α flux. Reconstruction of the Ly-α flux gives an estimate of log(LLy- α) = 20.65 (Ehrenreich et al. 2011; France et al. 2013). Using fitting relationships between the fluxes of multiple EUV bands and the Ly-α flux (Linsky et al. 2014), we estimate log(LEUV) = 20.54, which is consistent with the estimate from the X-ray flux. Since we are concerned about the evolution of the atmosphere on GJ 436 b, an age-dependent evolutionary history of the X-ray and the EUV fluxes is necessary. We adopt the following empirical estimates for stellar EUV fluxes based on X-ray observations and coronal modeling of 80 stars with spectral types spanning from M to F (Sanz-Forcada et al. 2011) (C1) (C2) where the luminosity has a unit of J s- 1 and the age (τ) has a unit of billion year. As a fact check, this adopted evolution history gives current fluxes of log(LX- ray) = 19.84 and log(LEUV) = 20.88, and these values are well within the ranges permitted by the above-mentioned observations. The age dependency is also consistent with a recent study (Shkolnik & Barman 2014). log(LEUV) = 22.12 - 1.24 log(τ), log(LX- ray) = 21.28 - 1.44 log(τ), REFERENCES Helium Atmosphere on Exoplanet 13 Agúndez, M., Venot, O., Selsis, F., & Iro, N. 2014, Astrophysical Journal, 781, 68 Allard, F., Hauschildt, P. H., Alexander, D. R., & Starrfield, S. 1997, Annual Review of Astronomy & Astrophysics, 35, 137 Alonso, R., Barbieri, M., Rabus, M., et al. 2008, Astronomy & Astrophysics, 487, L5 Ballard, S., Christiansen, J. L., Charbonneau, D., et al. 2010, Astrophysical Journal, 716, 1047 Baraffe, I., Chabrier, G., & Barman, T. 2008, Astronomy & Astrophysics, 482, 315 Barman, T. S., Hauschildt, P. H., & Allard, F. 2005, Astrophysical Journal, 632, 1132 Bean, J. L., Miller-Ricci Kempton, E., & Homeier, D. 2010, Nature, 468, 669 Beaulieu, J.-P., Tinetti, G., Kipping, D. M., et al. 2011, Astrophysical Journal, 731, 16 Ben-Jaffel, L., & Sona Hosseini, S. 2010, The Astrophysical Journal, 709, 1284 Benneke, B., & Seager, S. 2013, The Astrophysical Journal, 778, 153 Berta, Z. K., Charbonneau, D., Désert, J.-M., et al. 2012, Astrophysical Journal, 747, 35 Borysow, J., Frommhold, L., & Birnbaum, G. 1988, Astrophysical Journal, 326, 509 Burrows, A., & Sharp, C. M. 1999, Astrophysical Journal, 512, 843 Butler, R. P., Vogt, S. S., Marcy, G. W., et al. 2004, Astrophysical Journal, 617, 580 Cáceres, C., Ivanov, V. D., Minniti, D., et al. 2009, Astronomy & Astrophysics, 507, 481 Cecchi-Pestellini, C., Ciaravella, A., & Micela, G. 2006, Astronomy & Astrophysics, 458, L13 Cecchi-Pestellini, C., Ciaravella, A., Micela, G., & Penz, T. 2009, Astronomy & Astrophysics, 496, 863 Deming, D., Harrington, J., Laughlin, G., et al. 2007, Astrophysical Journal Letters, 667, L199 Demory, B.-O., Gillon, M., Barman, T., et al. 2007, Astronomy & Astrophysics, 475, 1125 Ehrenreich, D., Lecavelier Des Etangs, A., & Delfosse, X. 2011, Astronomy & Astrophysics, 529, A80 Erkaev, N. V., Kulikov, Y. N., Lammer, H., et al. 2007, Astronomy & Astrophysics, 472, 329 Erwin, J., Tucker, O. J., & Johnson, R. E. 2013, Icarus, 226, 375 Fortney, J. J., Mordasini, C., Nettelmann, N., et al. 2013, Astrophysical Journal, 775, 80 Fraine, J., Deming, D., Benneke, B., et al. 2014, Nature, 513, 526 France, K., Froning, C. S., Linsky, J. L., et al. 2013, Astrophysical Journal, 763, 149 Fressin, F., Torres, G., Charbonneau, D., et al. 2013, Astrophysical Journal, 766, 81 García Muñoz, A. 2007, Planetary and Space Science, 55, 1426 Gierasch, P. J., & Conrath, B. J. 1985, Energy conversion processes in the outer planets (Cambridge University Press, Cambridge and New York), 121–146 Gillon, M., Pont, F., Demory, B.-O., et al. 2007, Astronomy & Astrophysics, 472, L13 Guo, J. H. 2013, The Astrophysical Journal, 766, 102 Hansen, B. M. S., & Murray, N. 2012, Astrophysical Journal, 751, 158 Holmström, M., Ekenbäck, A., Selsis, F., et al. 2008, Nature, 451, 970 Howard, A. W. 2013, Science, 340, 572 Hu, R., & Seager, S. 2014, Astrophysical Journal, 784, 63 Hu, R., Seager, S., & Bains, W. 2012, Astrophysical Journal, 761, 166 -. 2013, Astrophysical Journal, 769, 6 Hunten, D. M., Pepin, R. O., & Walker, J. C. G. 1987, Icarus, 69, 532 Ikoma, M., & Hori, Y. 2012, Astrophysical Journal, 753, 66 Inamdar, N. K., & Schlichting, H. E. 2014, ArXiv e-prints, arXiv:1412.4440 Johnson, R. E., Volkov, A. N., & Erwin, J. T. 2013, Astrophysical Journal Letters, 768, L4 Kislyakova, K. G., Holmström, M., Lammer, H., Odert, P., & Khodachenko, M. L. 2014, Science, 346, 981 Knutson, H. A., Benneke, B., Deming, D., & Homeier, D. 2014a, Nature, 505, 66 Knutson, H. A., Madhusudhan, N., Cowan, N. B., et al. 2011, Astrophysical Journal, 735, 27 Knutson, H. A., Dragomir, D., Kreidberg, L., et al. 2014b, The Astrophysical Journal, 794, 155 Koskinen, T., Harris, M., Yelle, R., & Lavvas, P. 2013, Icarus, 226, 1678 Koskinen, T. T., Yelle, R. V., Lavvas, P., & Lewis, N. K. 2010, The Astrophysical Journal, 723, 116 Kreidberg, L., Bean, J. L., Désert, J.-M., et al. 2014, Nature, 505, 69 Kulow, J. R., France, K., Linsky, J., & Loyd, R. O. P. 2014, Astrophysical Journal, 786, 132 Lammer, H., Erkaev, N. V., Odert, P., et al. 2013, Monthly Notices of the Royal Astronomical Society, 430, 1247 Lammer, H., Selsis, F., Ribas, I., et al. 2003, Astrophysical Journal Letters, 598, L121 Lanotte, A. A., Gillon, M., Demory, B.-O., et al. 2014, ArXiv e-prints, arXiv:1409.4038 Lecavelier des Etangs, A., Vidal-Madjar, A., McConnell, J. C., & Hébrard, G. 2004, Astronomy and Astrophysics, 418, L1 Lewis, N. K., Showman, A. P., Fortney, J. J., et al. 2010, Astrophysical Journal, 720, 344 Liang, M.-C., Parkinson, C. D., Lee, A. Y.-T., Yung, Y. L., & Seager, S. 2003, Astrophysical Journal Letters, 596, L247 Line, M. R., Knutson, H., Wolf, A. S., & Yung, Y. L. 2014, Astrophysical Journal, 783, 70 Line, M. R., Vasisht, G., Chen, P., Angerhausen, D., & Yung, Y. L. 2011, Astrophysical Journal, 738, 32 Linsky, J. L., Fontenla, J., & France, K. 2014, Astrophysical Journal, 780, 61 Llama, J., & Shkolnik, E. L. 2015, The Astrophysical Journal, 802, 41 Lopez, E. D., Fortney, J. J., & Miller, N. 2012, Astrophysical Journal, 761, 59 Loyd, R. O. P., & France, K. 2014, The Astrophysical Journal Supplement, 211, 9 Madhusudhan, N., & Burrows, A. 2012, Astrophysical Journal, 747, 25 Madhusudhan, N., & Seager, S. 2011, Astrophysical Journal, 729, 41 Mason, E. A., & Marrero, T. R. 1970, Advances in Atomic and Molecular Physics, 6, 155 Miller-Ricci, E., Seager, S., & Sasselov, D. 2009, Astrophysical Journal, 690, 1056 Morello, G., Waldmann, I. P., Tinetti, G., et al. 2015, ArXiv e-prints, arXiv:1501.05866 Morley, C. V., Fortney, J. J., Kempton, E. M.-R., et al. 2013, Astrophysical Journal, 775, 33 Moses, J. I., Visscher, C., Fortney, J. J., et al. 2011, Astrophysical Journal, 737, 15 Moses, J. I., Line, M. R., Visscher, C., et al. 2013, Astrophysical Journal, 777, 34 Murray-Clay, R. A., Chiang, E. I., & Murray, N. 2009, Astrophysical Journal, 693, 23 Nettelmann, N., Kramm, U., Redmer, R., & Neuhäuser, R. 2010, Astronomy & Astrophysics, 523, A26 Owen, J. E., & Jackson, A. P. 2012, Monthly Notices of the Royal Astronomical Society, 425, 2931 Owen, J. E., & Wu, Y. 2013, Astrophysical Journal, 775, 105 Parmentier, V., Showman, A. P., & Lian, Y. 2013, Astronomy & Astrophysics, 558, A91 Pont, F., Gilliland, R. L., Knutson, H., Holman, M., & Charbonneau, D. 2009, Monthly Notices of the Royal Astronomical Society, 393, L6 Poppenhaeger, K., Robrade, J., & Schmitt, J. H. M. M. 2010, Astronomy & Astrophysics, 515, A98 Rafikov, R. R. 2006, Astrophysical Journal, 648, 666 Rogers, L. A., & Seager, S. 2010, Astrophysical Journal, 712, 974 14 Hu et al. Sanz-Forcada, J., Micela, G., Ribas, I., et al. 2011, Astronomy & Astrophysics, 532, A6 Schunk, R. W., & Nagy, A. F. 1980, Reviews of Geophysics, 18, 813 Seager, S., & Deming, D. 2010, Annual Review of Astronomy & Astrophysics, 48, 631 Shkolnik, E. L., & Barman, T. S. 2014, ArXiv e-prints, arXiv:1407.1344 Spitzer, L. 1978, Physical processes in the interstellar medium (New York Wiley-Interscience) Stevenson, K. B., Harrington, J., Nymeyer, S., et al. 2010, Nature, 464, 1161 Stevenson, K. B., Harrington, J., Lust, N. B., et al. 2012, Astrophysical Journal, 755, 9 Sudarsky, D., Burrows, A., & Hubeny, I. 2003, Astrophysical Journal, 588, 1121 Tarafdar, S. P., & Vardya, M. S. 1969, Monthly Notices of the Royal Astronomical Society, 145, 171 Tian, F., Toon, O. B., Pavlov, A. A., & De Sterck, H. 2005, Astrophysical Journal, 621, 1049 Torres, G., Winn, J. N., & Holman, M. J. 2008, Astrophysical Journal, 677, 1324 Tucker, O. J., Erwin, J. T., Deighan, J. I., Volkov, A. N., & Johnson, R. E. 2012, Icarus, 217, 408 Volkov, A. N., & Johnson, R. E. 2013, Astrophysical Journal, 765, 90 Yelle, R. V. 2004, Icarus, 170, 167 Yung, Y. L., & Demore, W. B., eds. 1999, Photochemistry of planetary atmospheres Zahnle, K., Kasting, J. F., & Pollack, J. B. 1990, Icarus, 84, 502 Zahnle, K. J., & Kasting, J. F. 1986, Icarus, 68, 462
1808.05154
1
1808
2018-08-15T15:43:06
A differential algebra based importance sampling method for impact probability computation on Earth resonant returns of Near Earth Objects
[ "astro-ph.EP" ]
A differential algebra based importance sampling method for uncertainty propagation and impact probability computation on the first resonant returns of Near Earth Objects is presented in this paper. Starting from the results of an orbit determination process, we use a differential algebra based automatic domain pruning to estimate resonances and automatically propagate in time the regions of the initial uncertainty set that include the resonant return of interest. The result is a list of polynomial state vectors, each mapping specific regions of the uncertainty set from the observation epoch to the resonant return. Then, we employ a Monte Carlo importance sampling technique on the generated subsets for impact probability computation. We assess the performance of the proposed approach on the case of asteroid (99942) Apophis. A sensitivity analysis on the main parameters of the technique is carried out, providing guidelines for their selection. We finally compare the results of the proposed method to standard and advanced orbital sampling techniques.
astro-ph.EP
astro-ph
MNRAS 000, 1 -- 19 (2017) Preprint 16 August 2018 Compiled using MNRAS LATEX style file v3.0 A differential algebra based importance sampling method for impact probability computation on Earth resonant returns of Near Earth Objects Matteo Losacco,1(cid:63) Pierluigi Di Lizia,1 Roberto Armellin2 and Alexander Wittig3 1Department of Aerospace Science and Technology, Politecnico di Milano, Via G. La Masa 34, 20156 Milano, Italy 2Surrey Space Centre, University of Surrey, GU2 7XH Guildford, United Kingdom 3Aeronautics, Astronautics and Computational Engineering Unit, University of Southampton, SO17 1BJ, Southampton, United Kingdom Accepted XXX. Received YYY; in original form ZZZ ABSTRACT A differential algebra based importance sampling method for uncertainty propagation and impact probability computation on the first resonant returns of Near Earth Ob- jects is presented in this paper. Starting from the results of an orbit determination process, we use a differential algebra based automatic domain pruning to estimate resonances and automatically propagate in time the regions of the initial uncertainty set that include the resonant return of interest. The result is a list of polynomial state vectors, each mapping specific regions of the uncertainty set from the observation epoch to the resonant return. Then, we employ a Monte Carlo importance sampling technique on the generated subsets for impact probability computation. We assess the performance of the proposed approach on the case of asteroid (99942) Apophis. A sensitivity analysis on the main parameters of the technique is carried out, providing guidelines for their selection. We finally compare the results of the proposed method to standard and advanced orbital sampling techniques. Key words: individual: (99942) Apophis. celestial mechanics -- methods: statistical -- minor planets, asteroids: 1 INTRODUCTION Over the last thirty years, significant efforts have been de- voted to develop new tools for detection and prediction of planetary encounters and potential impacts by Near Earth Objects (NEO). The task introduces relevant challenges due to the imperative of early detection and accurate estimation and propagation of their state and associated uncertainty set (Chesley 2005). The problem is made more complicated by the fact that the dynamics describing the motion of these ob- jects is highly nonlinear, especially during close encounters with major bodies. Nonlinearities of the orbital dynamics tend to significantly stretch the initial uncertainty sets dur- ing the time propagation. Nonlinearities are not confined to object dynamics only: even simple conversions between co- ordinate systems introduce nonlinearities, thus affecting the accuracy of classical propagation techniques (Wittig et al. 2015). Present day approaches for robust detection and pre- diction of planetary encounters and potential impacts by NEO mainly refer to linearised models or full nonlinear or- bital sampling (Farnocchia et al. 2015). The impact prob- (cid:63) E-mail: [email protected] © 2017 The Authors ability computation by means of linear methods in the im- pact plane was introduced by Chodas (1993), whereas the introduction of the Monte Carlo technique to this problem was developed by Yeomans & Chodas (1994) and Chodas & Yeomans (1999), who suggested to apply the method to sample the linear six dimensional confidence region at the observation epoch and then numerically integrate over the time interval of investigation using fully nonlinear equations (Milani et al. 2002). Milani et al. (1999), Milani (1999) and Milani et al. (2000a,b) applied the multiple solutions ap- proach to sample the central Line of Variations (LOV) of the nonlinear confidence region at the initial epoch and then nu- merically integrate over the time span of interest in a similar way. Within the framework of the impact probability com- putation of resonant returns, a well-known approach relies on the concept of keyholes, small regions of the impact plane of a specific close encounter such that, if an asteroid passes through one of them, it will hit the Earth on subsequent return (Gronchi & Milani 2001; Milani et al. 2002; Valsecchi et al. 2003). The preferred approach to detecting potential impacts depends on the uncertainty in the estimated orbit, the inves- tigated time window and the dynamics between the obser- 2 M. Losacco et al. vation epoch and the epoch of the expected impact (Farnoc- chia et al. 2015). Linear methods are preferred when linear approximations are reliable for both the orbit determina- tion and uncertainty propagation. When these assumptions are not valid, one must resort to more computationally in- tensive techniques: among these, Monte Carlo methods are the most accurate but also the most computationally inten- sive, whereas the LOV method guarantees compute times 3-4 orders of magnitude lower than those required in MC simulations, though the LOV analysis may grow quite com- plex after it has been stretched and folded by multiple close planetary encounters, leaving open the possibility of missing some pathological cases (Farnocchia et al. 2015). Alternative approaches rely on the use of Differential Algebra (DA). Differential algebra supplies the tools to com- pute the derivatives of functions within a computer environ- ment, i.e. it provides the Taylor expansion of the flow of Ordinary Differential Equations (ODEs) by carrying out all the operations of any explicit integration scheme in the DA framework (Berz 1999; Wittig et al. 2015). DA has already proven its efficiency in the nonlinear propagation of uncer- tainties (Armellin et al. 2010b; Morselli et al. 2012; Valli et al. 2013). Nonetheless, the accuracy of the method dras- tically decreases in highly nonlinear dynamics. The propaga- tion of asteroids motion after a close encounter with a major body is a typical case. A DA based automatic domain splitting algorithm was presented by the authors in the past to overcome the limita- tions of simple DA propagation (Wittig et al. 2014a,b, 2015). The method can accurately propagate large sets of uncer- tainties in highly nonlinear dynamics and long term time spans. The propagation algorithm automatically splits the initial uncertainty domain into subsets when the polynomial expansions representing the current state do not meet pre- defined accuracy requirements. The performance of the algo- rithm was assessed on the case of asteroid (99942) Apophis, providing a description of the evolution of the uncertainty set to the epoch of predicted close encounters with Earth in 2036 and 2037 (Wittig et al. 2015). Though representing a significant improvement with respect to simple DA propaga- tion, the approach required a not negligible computational effort in propagating the whole set of generated subdomains. Moreover, no information about the impact probability for asteroid Apophis was provided, as the propagation of the uncertainty set was stopped before the close encounters. We present in this paper an evolution of the automatic domain splitting algorithm. The method, referred to as auto- matic domain pruning, automatically identifies possible res- onances after a close encounter with a major body. Then, as- suming no intervening close approaches with other celestial bodies in between, it optimizes the propagation to the first resonant returns, by limiting the propagation of the uncer- tainty set to the regions that generate a close encounter with that celestial body at the investigated epoch. The result is a list of polynomial state vectors, each mapping only specific subsets of the initial domain to the resonant return epoch. Taking advantage of the availability of the polynomial maps, a DA based Monte Carlo importance sampling technique is then used to generate samples in the propagated subsets and provide an estimate for the impact probability at the epoch of the selected resonant return. The proposed approach does not apply any simplification step on the uncertainty domain Figure 1. Analogy between the FP representation of real num- bers in computer environment (left) and the algebra of Taylor polynomials in DA framework (right)(Di Lizia 2008). associated with the orbit determination process. Thus, the method is proposed as an alternative approach with respect to equivalent techniques, such as a full Monte Carlo simu- lation or other six dimensional-based orbital sampling tech- niques, which will represent the main term of comparison for our analysis. The paper is organized as follows. First, we present a de- scription of the automatic domain pruning and importance sampling techniques, showing the application to the case of the first resonant return. Then, we apply the method to the critical case of asteroid (99942) Apophis, providing an estimate of the impact probability for the resonant return in 2036. Finally, we carry out a sensitivity analysis on the main parameters of the method, presenting a comparison with standard and advanced orbital sampling techniques. 2 DIFFERENTIAL ALGEBRA AND AUTOMATIC DOMAIN SPLITTING Differential algebra provides the tools to compute the deriva- tives of functions within a computer environment (Ritt 1932, 1948; Risch 1969, 1970; Kolchin 1973; Berz 1999). Histori- cally, the treatment of functions in numerics has been based on the treatment of numbers, and the classical numerical algorithms are based on the evaluation of functions at spe- cific points. The basic idea of DA is to bring the treatment of functions and the operations on them to the computer environment in a similar way as the treatment of real num- bers (Berz 1999). Real numbers, indeed, are approximated by floating point (FP) numbers with a finite number of dig- its. With reference to Fig. 1, let us consider two real num- bers a and b, and their FP counterpart ¯a and ¯b respectively: given any operation × in the set of real numbers, an ad- joint operation ⊗ is defined in the set of FP numbers so that the diagram in figure commutes. Consequently, trans- forming the real numbers a and b in their FP representation and operating on them in the set of FP numbers returns the same result as carrying out the operation in the set of real numbers and then transforming the achieved result in its FP representation. In a similar way, suppose two sufficiently regular functions f and g are given. In the framework of DA, these functions are converted into their Taylor series expan- sions, F and G respectively. In this way, the transformation of real numbers in their FP representation is now substi- tuted by the extraction of the Taylor expansions of f and g (see Fig. 1, right). For each operation in the function space, an adjoint operation in the space of Taylor polynomials is defined such that the corresponding diagram commutes. MNRAS 000, 1 -- 19 (2017) a,b∈Ra,b∈FPa×b×⊗a⊗bTTf,gf×g×⊗TTF,GF⊗G ADP -- IS method for impact probability computation 3 The implementation of DA in a computer environment provides the Taylor coefficients of any function of v variables up to a specific order n. More specifically, by substituting classical real algebra with the implementation of a new al- gebra of Taylor polynomials, any function f of v variables can be expanded into its Taylor expansion up to an arbi- trary order n, along with the function evaluation, with a limited amount of effort. The Taylor coefficients of order n for sum and product of functions, as well as scalar products with real numbers, can be directly computed from those of summands and factors. As a consequence, the set of equiva- lence classes of functions can be endowed with well-defined operations, leading to the so-called truncated power series algebra. In addition to basic algebraic operations, differenti- ation and integration can be easily introduced in the algebra, thus finalizing the definition of the differential algebra struc- ture of DA (Berz 1986, 1987). The DA used in this work is implemented in the DACE software (Rasotto et al. 2016). x(t0) = x0 A relevant application of DA is the automatic high or- der expansion of the solution of an ODE with respect to the initial conditions (Berz 1999; Di Lizia et al. 2008; Rasotto et al. 2016). This expansion can be achieved by considering that any integration scheme, explicit or implicit, is charac- terized by a finite number of algebraic operations, involving the evaluation of the ODE right hand side (RHS) at sev- eral integration points. Therefore, replacing the operations between real numbers with those on DA numbers, it yields to the nth order Taylor expansion of the flow of the ODE, φ(t; δx0, t0) = Mφ(δx0), at each integration time, assuming a perturbed initial condition x0 + δx0. Without loss of gener- ality, consider the scalar initial value problem: (cid:219)x(t) = f (t, x), (1) and the associated flow φ(t; δx0, t0). For simplicity, consider uncertain initial conditions only. Starting from the nth order DA representation of the initial condition, [x0] = x0 + δx0, which is a (n +1)-tuple of Taylor coefficients, and performing all the operations in the DA framework, we can propagate the Taylor expansion of the flow in x0 forward in time, up to the final time t f . Consider, for example, the forward Euler's scheme: xi = xi−1 + f (xi−1)∆t (2) and replace the initial value with the DA expression [x0] = x0 + δx0. The first time step yields [x1] = [x0] + f ([x0])∆t If the function f is evaluated in the DA framework, the output of the first step, [x1], is the nth order Taylor expansion of the flow φ(t; δx0, t0) in x0 for t = t1. Note that, as a result of the DA evaluation of f ([x0]), the (n + 1)-tuple [x1] may include several non zero coefficients corresponding to high order terms in δx0. The previous procedure can be repeated for the subsequent steps. The result at the final step is the nth order Taylor expansion of φ(t; δx0, t0) in x0 at the final time t f . Thus, the flow of a dynamical system is approximated, at each time step ti, with its nth order Taylor expansion in a fixed amount of effort. Any explicit ODE integration scheme can be rewritten as a DA scheme. For the numerical integrations presented in this paper, a DA version of a 7/8 Dormand Prince (8th order solution for (3) MNRAS 000, 1 -- 19 (2017) propagation, 7th order solution for step size control) Runge Kutta scheme is used. The main advantage of the DA based approach is that there is no need to write and integrate variational equations to obtain high order expansions of the flow. It is therefore independent on the RHS of the ODE and it is computation- ally efficient. Unfortunately, DA fails to accurately describe, with a single polynomial map, the evolution in time of an uncertainty set in case of highly nonlinear dynamics or long term propagation. The approximation error is strictly re- lated to the size of the domain the polynomial is defined in (Wittig et al. 2015). The approximation error between an n + 1 times differentiable function f ∈ Cn+1 and its Taylor expansion Pf of order n, without loss of generality taken around the origin, is given by Taylor's theorem: f (δx) − Pf (δx) (cid:54) C · δxn+1 (4) for some constant C > 0. Consider now the maximum error er of Pf on a domain Br of radius r > 0 around the expansion point. Considering equation (4), we obtain: f (δx) − Pf (δx) (cid:54) C · δxn+1 (cid:54) C · r n+1 = er (5) If the domain of Pf is reduced from Br to Br/2 of radius r/2, the maximum error of Pf over Br/2 will decrease by a factor 1/2n+1: f (δx) − Pf (δx) (cid:54) C · δxn+1 (cid:54) C ·(cid:16) r (cid:17) n+1 = er 2n+1 (6) 2 By subdividing the initial domain into smaller do- mains and computing the Taylor expansion around the cen- ter points of the new domains, the error greatly reduces, whereas the expansions still cover the entire initial set. Start- ing from these considerations, Automatic Domain Splitting (ADS) employs an automatic algorithm to determine at which time ti the flow expansion over the set of initial condi- tions is no longer able to describe the dynamics with enough accuracy (Wittig et al. 2015). Once this case has been de- tected, the domain of the original polynomial expansion is divided along one of the expansion variables into two do- mains of half their original size. By re-expanding the polyno- mials around the new centre points, two separate polynomial expansions are obtained. By defining with xi the splitting direction, both generated polynomial expansions P1 and P2 have terms of order n in xi smaller by a factor of 2n with respect to the original polynomial expansion P. Thus, the splitting procedure guarantees a more accurate description of the whole uncertainty set at the current time epoch ti. After such a split occurs, the integration process is resumed on both generated subsets, until new splits are required. A representation of the ADS procedure is shown in Fig. 2. The decision on the splitting epoch and, in case of multi- variate polynomials, the splitting direction relies on estimat- ing the size of the (n + 1)th order terms of the polynomial using an exponential fit of the size of all the known non- zero terms up to order n. If the size of the truncated order becomes too large, we decide to split the polynomial. This method allows us to consider all the information available in the polynomial expansion and to obtain an accurate esti- mate of the size of the n + 1 order term, the first discarded order. The exponential fit is chosen because, after reducing the domain with a sufficient number of splits, the coefficients of the resulting polynomial expansion decay exponentially as 4 M. Losacco et al. Figure 2. ADS algorithm schematic illustration (Wittig et al. 2015). of (cid:112)1 + x/2 via exponential fitting. Terms of order up to 9 are Figure 3. Truncation error estimation for the Taylor expansion used for the fitting (Wittig et al. 2015). a direct consequence of Taylor's theorem. A mathematical description is offered hereafter and follows the scheme pre- sented in Wittig et al. (2015). Consider a polynomial P of order n of the form aαx α (7) P(x) = Si =  α aα written using multi-index notation, the size Si of the terms of order i is computed as the sum of the absolute values of all coefficients of exact order i: (8) α=i We denote by I the set of indices i for which Si is non- zero. A least squares fit of the exponential function f (i) = AeBi (9) is used to determine the coefficients A and B such that f (i) = Si, i ∈ I, is approximated optimally in least squares sense. Then, the value of f (n+1) is used to estimate the size Sn+1 of the truncated order n +1 of P. An example of the application of the method is shown in Fig. 3, where the polynomial is the Taylor expansion of(cid:112)1 + x/2 up to order 9. The size Si of each order is shown as bars, whereas the resulting fitted function f is shown as a line. In the case of multivariate polynomials P(x) = P(x1, x2, . . . , xv), the split is performed in one component xi. We determine the splitting direction using a method simi- lar to the one adopted for the splitting decision. For each j = 1, . . . , v we begin by factoring the known coefficients of P of order up to n with respect to xj , i.e. P(x1, x2, . . . , xv) = xm j · qj,m(x1, . . . , xj−1, xj+1, . . . , xv) (10) n m=0 where the polynomials qj,m do not depend on xj . The size Sj,m of the polynomials qj,m is estimated by the sum of the absolute values of their coefficients. Then, the exponential fitting routine is applied to estimate the size Sj,n+1 of the truncated terms of order n + 1 in xj . Finally, the splitting direction i is chosen as the component xj with the largest truncation error Sj,n+1. In this way, all splits are performed in the direction of the variable that currently has the largest estimated contribution to the total truncation error of the polynomial P. The main parameters of the algorithm are the tolerance for the splitting procedure and the maximum number of al- lowed splits Nmax. The first parameter is selected according to the required precision of the polynomial expansions and determines the splitting epochs: when the estimated trun- cation error exceeds the imposed tolerance, the current do- main is split. As a direct consequence of the ADS procedure, the maximum error over the obtained set of polynomials de- creases with the selected splitting precision. However, the maximum error is always larger than the selected integra- tion precision. This difference is actually expected, as the splitting tolerance plays a similar role as the one-step er- ror set in the automatic step size control of the integra- tion scheme (Wittig et al. 2015). It is the maximum error that can accumulate at any time before the integrator takes action to reduce further error accumulation. However, the accumulated error at the time of the splitting cannot be undone as the splitting only re-expands the polynomial to prevent exponential growth in future integration steps. The ideal tolerance depends on both the dynamics and the inte- gration time, and it has to be chosen heuristically to ensure the final result satisfies the accuracy requirements of the application. A numerical example is shown in Section 7.1. The second parameter plays the role of limiting the number of generated subdomains by imposing a minimum size for the generated subsets: domain splitting is disabled on any set whose volume is less than 2−Nmax times that of the initial domain. That is, any set is split at most Nmax times. Then, instead of splitting a set further, integration is stopped at the attempt to perform the (Nmax +1)th split and the resulting polynomial expansion is saved as incomplete. Incomplete polynomials are later treated separately in the analysis of the results (Wittig et al. 2015). When each generated subset reaches either the final sim- ulation time or the minimum box size, the ADS propagation terminates, and the result is a list of polynomial expansions, each covering a specific subset of the domain of initial con- ditions. A more detailed description of the ADS algorithm can be found in Wittig et al. (2015). 3 AUTOMATIC DOMAIN PRUNING As described in Section 2, automatic domain splitting pro- vides an accurate description of the evolution in time of a given uncertainty set by splitting the domain in subsets when required. Unfortunately, this method may entail a not negligible computation effort, as all generated subsets are MNRAS 000, 1 -- 19 (2017) 𝑡𝑡=𝑡𝑡0𝑡𝑡=𝑡𝑡𝑖𝑖error < 𝜀𝜀𝑡𝑡=𝑡𝑡𝑖𝑖+1error > 𝜀𝜀error < 𝜀𝜀A.Wittigetal.Fig.4IllustrationoftheestimationofthetruncationerroroftheTaylorexpansionof√1+x/2viaexponentialfitting.Termsoforderupto9areusedforthefitting(hatched),whilethe10thorderterm(white)isshownforreference 1e-06 1e-05 0.0001 0.001 0.01 0.1 1 0 1 2 3 4 5 6 7 8 9 10Absolute SizeOrderExactEstimatewherethepolynomialsqj,mdonotdependonxj.ThenthesizeSj,mofthepolynomi-alsqj,misestimatedbythesumoftheabsolutevaluesoftheircoefficientsandthesameexponentialfittingroutineasdescribedaboveisappliedtoobtainanestimateofthesizeSj,n+1ofthetruncatedtermsofordern+1inxj.Finallythesplittingdirectioniischo-sentobethedirectioncorrespondingtothecomponentxjwithlargesttruncationerrorSj,n+1.Inthisway,allsplitsareperformedinthedirectionofthevariablethatcurrentlyhasthelargestestimatedcontributiontothetotaltruncationerrorofthepolynomialP,andthusthesplitshavethemaximalimpactonreducingtheapproximationerror.Thesplittingprocessdescribedhereingeneral,andtheselectionofthesplittingdirectioninparticular,arestronglydependentontheparametrizationoftheinitialcondition.Thedirectionofmax-imumexpansioningeneralisnotalignedwithasingledirectionoftheparametrization,inwhichcaseseveralvariableswillcontributetothetruncationerror.Inthiscase,splitsoccurautomaticallyalongallvariablesinvolved.However,theinitialconditioncanoftenbeparametrizedsuchthatexpansionhappensmainlyalongonlyafeworevenjustoneofthedirections.3.1Kepler'sdynamicsexample:domainsplittingillustrationBeforewepresentafullanalysisoftheeffectofthesplittingprecisionontheaccuracy,efficiencyandnumberoffinalsetsinthenextsubsection,wefirstdemonstratethedomainsplittingtechniquedescribedintheprevioussection.WeapplyittothesameproblemofpropagatingKepler'sdynamicsaspresentedintheSect.2.1.Computationsareperformedatorder14withthesameinitialconditionbox.Thesplittingprecisionissettoε=3×10−4,meaningthatwhentheestimatedtruncationerrorofanexpansionexceedsthislimitasplitistriggered.Thelimitwaschosenthishightoallowforabettervisualizationofthesplittingprocess,inactualapplicationsthelimitistypicallychosenmuchlower.Integratingthedynamicsfromtimet0=0totimetf=50(2.81nominalrevolutions),theentirecomputationtakesabout22sonthesamemachineusedfortheexampleintheprevioussection,andproduces23finalpolynomialexpansionscoveringtheinitialcondition.Figure5showstheresultingsetsatvarioustimesduringtheintegration.Upuntiltimeta=16day(0.90nominalrevolutions),theentiresetiswelldescribedbyasingleDAexpansion.Attimetb=17(0.96nominalrevolutions),justbeforecompletingthefirstrevolution2splitshaveoccurred,leadingtothreepolynomialpatches.Anothersplitisperformedattimetc=33(1.86nominalrevolutions).Figure5dshowsthe15DApatchesthatarenecessary123 ADP -- IS method for impact probability computation 5 propagated to the final simulation time or till the minimum box size is reached. While this approach is unavoidable when the behaviour of the whole uncertainty set is analysed, it be- comes a strong limitation when only a portion of the initial set is to be investigated. This is the case when the first res- onant return of a Near Earth Object is studied. Resonant returns occur when, during a close encounter, an asteroid is perturbed into an orbit with a period T(cid:48) ∼ k/h years. Thus, after h revolutions of the asteroid and k revolutions of the Earth, both celestial bodies are in the same region of the first close encounter and a second one may occur. Given the initial uncertainty set, only a portion of it may lead to the resonant return. It would be therefore interesting to limit the propagation to this region only. Starting from these considerations, the Automatic Do- main Pruning (ADP) we present in this paper combines the ADS algorithm with a pruning technique with the aim of limiting the number of propagated subsets. We make here the assumption of no close approaches with other celestial bodies between the first close encounter and the selected res- onant return. This assumption is easily checked right before the ADP propagation, as later explained in Section 7.2. The first phase of the algorithm consists in propagating the whole uncertainty set by means of ADS propagation up to the epoch of the first close encounter. The availability of the polynomial expansion of the state vector of the object with respect to the initial uncertainty provides the polyno- mial expansion of the orbital period of the object after the close encounter. By using a polynomial bounder, we can es- timate the range of all possible values of the orbital period after the close encounter and, thus, retrieve all possible reso- nances with the planet, i.e. all orbital periods included in the computed orbital period range leading to a resonant return with the planet. Once all resonances are computed, the analysis focuses on a single resonance, and the propagation is resumed. Every time a new subset is generated, the method automatically identifies if the set may lead to the investigated resonant return or not. By exploiting the knowledge of the DA state vector at the epoch of the first close encounter, indeed, we can assign a given orbital period range to each generated subset. This range, defined as ∆Tsub, is compared to a refer- ence range ∆Tref, centred in the resonance period T(cid:48) with a semi-amplitude ε, ∆Tref = [(1 − ε)T(cid:48), (1 + ε)T(cid:48)]. We select the reference range in order to consider small dynamical pertur- bations between the first close encounter and the resonant return. If ∆Tsub is at least partially included in the reference range, then the current subset is retained, and its propaga- tion is continued. If ∆Tsub is not included in the reference range, then the initial conditions included in the current subset do not lead to a resonant return at the investigated epoch, and so the subset is discarded. This way, subsets are dynamically pruned during the ADS propagation. An illus- tration of the ADP algorithm is shown in Fig. 4. The ADP algorithm, therefore, does not alter the se- quence of generated subdomains, but limits the propagation in time to those subsets that are involved in the investigated resonant return. This pruning action has a positive impact on the overall computational burden, since the computa- tional effort required by the propagation of all the discarded subsets is saved. As only subsets with close approaches to the Earth at the epoch of the investigated resonant return MNRAS 000, 1 -- 19 (2017) Figure 4. ADP algorithm illustration. Pruning is performed by comparing the estimated subset orbital period range ∆Tsub with the reference range ∆Tref. are maintained, the result at the end is a set of subdomains whose propagation stops slightly before the epoch of the in- vestigated resonant return for having reached their minimum box size. 4 IMPORTANCE SAMPLING METHOD The output of the ADP propagation is a list of subsets at epochs close to the investigated resonant return. Still, no value for the impact probability is available. We obtain an estimate for the impact probability by sampling the gen- erated subsets and propagating the samples till they reach their minimum geocentric distance. Among all possible sam- pling technique, we employ the Importance Sampling (IS) method (Zio 2013). The IS method amounts to replacing the original proba- bility density function (pdf) qx(x) with an Importance Sam- pling Distribution (ISD) qx(x) arbitrarily chosen by the an- alyst so as to generate a large number of samples in the importance region of the phase space F, the region of initial conditions leading to an impact with Earth at the epoch of the resonant return. In the case under study, we select the auxiliary distribution in order to limit as much as possible the generation of the samples to the subsets that get through the dynamic pruning. The IS algorithm is the following: Þ IF(x)qx(x) qx(x) qx(x)dx (i) Identify a proper qx(x). (ii) Express the impact probability p(F) as a function of IF(x)qx(x)dx = qx(x). p(F) = (11) where IF(x) : IRv → {0, 1} is an indicator function such that IF(x) = 1 if x ∈ F, 0 otherwise. (iii) Draw NT samples x k : k = 1, 2, . . . , NT from the im- portance sampling distribution qx(x). If a good choice for the auxiliary pdf is made, the generated samples concen- trate in the region F. (iv) Compute the estimate p(F) for the impact probability p(F) by resorting to equation (11): Þ NT k=1 p(F) = 1 NT IF(x k)qx(x k) qx(x k) (12) S1S2𝑡𝑡=𝑡𝑡0𝑡𝑡=𝑡𝑡𝑖𝑖S2S1error < 𝜀𝜀Store S1Discard S2∆𝑇𝑇𝑟𝑟𝑟𝑟𝑟𝑟∆𝑇𝑇𝑟𝑟𝑟𝑟𝑟𝑟∆𝑇𝑇𝑆𝑆𝑆∆𝑇𝑇𝑆𝑆𝑆𝑡𝑡=𝑡𝑡𝑖𝑖+𝑆error > 𝜀𝜀error < 𝜀𝜀𝜀𝜀𝜀𝜀S1 6 M. Losacco et al. (v) Compute the variance of the estimator p(F) as: σ2( p) = qx(x)dx − p2(F) x(x) F (cid:32)Þ I2 1 NT (x)q2 x(x) q2 (cid:33) (cid:16)(cid:92)p2(F) − p2(F)(cid:17) ≈ 1 NT (13) The selection of the ISD represents the most critical point for the method. Several techniques have been devel- oped in order to find the one giving small variance for the estimator (Zio 2013). In this paper, we shape the ISD ac- cording to the result of the ADP propagation. As described in Section 3, the ADP propagation provides a list of subsets whose propagation is stopped slightly before the resonant return. All subsets are identified as Potentially Hazardous Subdomains (PHS's), but no probability ranking is provided by the ADP propagation. Starting from these considerations, we define the ISD as a uniform probability density function including all the generated subsets over the whole domain. This selection allows us to increase the number of samples drawn in the PHS's and, eventually, in the impact-leading region. 5 AUTOMATIC DOMAIN PRUNING IMPORTANCE SAMPLING METHOD The combination of the methods presented in Sections 3 and 4 yields the ADP importance sampling method (ADP -- IS) for uncertainty propagation and impact probability compu- tation of the first resonant returns of NEO. The starting point is represented by the output of an orbit determination process of a given NEO at the observation epoch t0. This output can be expressed in terms of estimated state vector and related covariance matrix. Then, the steps of the ADP propagation phase are the following: (i) Consider the initial state vector and related pdf and perform an analysis to identify possible epochs of close en- counters and resonant returns. The analysis is carried out by propagating the uncertainty set using ADS up to the first close encounter, computing the semi-major axis dispersion over the set with a polynomial bounder and identifying the resonant frequencies. The validity of the resonances is then checked as explained in Section 7.2. (ii) Select a resonance and identify its epoch tres. (iii) Perform an ADP propagation till the epoch tres. Ev- ery time a split is required, compare the orbital period range of the current subset ∆Tsub with the reference range ∆Tref: ∆Tsub ∩ ∆Tref (cid:40)= 0 discard the current subset (14) include the current subset (cid:44) 0 f The method provides a set of nPHS PHS's and related ] at the truncation time ti f , with i = ] is a polynomial state vector, each DA state vectors [x i 1, . . . , nPHS. Vector [x i component being a function of the initial conditions x i 0. The IS phase is initialized by setting the value of the estimated impact probability pold and the number of itera- tions nit equal to zero. Then, the steps of the algorithm are the following: f (cid:40)0 (ii) Set nit = nit + 1 and draw one sample x it (iii) Check if the sample belongs to one of the PHS's: if it is out of the PHS's, go back to step (ii), otherwise identify the correct PHS i the sample belongs to. 0 from qx(x). (iv) Compute the algebraic state vector x it 0 at the truncation epoch ti f correspond- ing to the drawn sample x it f by performing a polynomial evaluation of the DA state vector = [x i [x i ](x it 0 ). ] at x it f (v) Propagate the state vector x it f (vi) Compute the minimum geocentric distance dit of the selected resonant return. 0 . That is, x it f to the epoch from ti f f res and evaluate the indicator Iit F NT = 1 Iit F if dit if dit res > R⊕ res < R⊕ = 0, go back to step (ii), otherwise evaluate the new impact probability pnew. By reformulating equa- tion (12), we obtain: (vii) If Iit F (15) IF(x k)qx(x k) = 1 nit pnew = qx(x k) ( I + qx(x it 0 )) 1 qx(x it 0 ) is the value of the original pdf in x it 0)qx(x k 1 0 ) (16) NT k=1 where qx(x it 0 , qx(x it 0 ) 0 , whereas the term I is the value of the auxiliary pdf in x it represents the summation of all terms IF(x k 0) of the previous iterations. The total number of samples considered for the estimation is nit , i.e. the number of drawn samples when the estimate is computed. Note that, since the ISD is uniform over the whole set of PHS's, it can be extracted from the summation. (viii) Compare pold and pnew: if the relative difference is larger than an imposed tolerance, go back to step (ii), otherwise stop. 6 NUMERICAL SIMULATIONS: THE CASE OF ASTEROID (99942) APOPHIS In this section, we assess the performance of the ADP -- IS method on the evaluation of the impact probability for the test case of asteroid (99942) Apophis. Table 1 shows the nominal initial state and associated uncertainties σ for Apophis on June 18, 2009 expressed in terms of equinoctial parameters p = (a, P1, P2, Q1, Q2, l), considering a diagonal covariance matrix. Data were obtained from the Near Earth Objects Dynamic Site1 in September 2009. We selected a diagonal covariance matrix in order to help distinguish the contribution of the six orbital param- eters and test our method in a scenario in which the un- certainty volume is maximized. In general, however, this se- lection may lead to quite inaccurate results as uncertainties may be highly correlated. Nevertheless,the method can be applied in the most general case of full covariance matrix ex- actly in the same way, with the only difference that the DA variables would be placed along the directions of the covari- ance eigenvectors to avoid artificially adding extra-volume in the initial domain definition. As previously stated, the starting point, not including (i) Define the ISD function qx(x) as a uniform pdf includ- ing all the generated PHS's. 1 http://newton.dm.unipi.it/neodys/ MNRAS 000, 1 -- 19 (2017) ADP -- IS method for impact probability computation 7 Table 1. Apophis equinoctial parameters and related uncertain- ties on June 18, 2009 00:00:00 (TDB). Nominal value σ a P1 P2 Q1 Q2 l 0.922438242375914 -0.093144699837425 0.166982492089134 -0.012032857685451 -0.026474053361345 88.3150906433494 2.29775 · 10−8 AU 3.26033 · 10−8 - 7.05132 · 10−8 - 5.39528 · 10−8 - 1.83533 · 10−8 - 6.39035 · 10−5 ◦ recent optical and radar observations performed from late 2011 onward, was selected in order to test the algorithm against the most critical scenario. Asteroid Apophis will have a close encounter with Earth on April 13, 2029 with a nominal distance of 3.8 · 104 km (Chesley 2005). Accord- ing to the selected initial conditions, though an impact in 2029 can be ruled out, the perturbations induced by the en- counter open the door to resonant returns in 2036 and 2037. The aim is therefore to apply the presented method to pro- vide an estimate for the impact probability at the epoch of the first resonant return, in 2036. The motion of Apophis in the Solar system is modelled according to the (N + 1)body problem, including relativis- tic corrections to the Newtonian forces (Seidelmann 1992; Wittig et al. 2015). Specifically, the full equation is (cid:40)  i γ (cid:219)r2 c2 + (cid:220)r = G mi(r i − r) (1 + γ) (cid:219)r i2 r3 i (cid:41) c2 + G 1 − 2(β + γ) c2 G − 2(1 + γ)  (cid:40) 3 + 4γ c2 mi c2ri i 2 1 2c2 (r i − r) · (cid:220)r i   − 2β − 1 (cid:20)(r − r i) · (cid:219)r i c2 G mj rj (cid:219)r · (cid:219)r i − 3 2c2 {[r − r i] · [(2 + 2γ)(cid:219)r mj ri j (cid:21)2 j(cid:44)i ri j (cid:220)r i + (1 + 2γ)(cid:219)r i]}((cid:219)r − (cid:219)r i) r2 i r2 i (17) where r is the position of Apophis in Solar System barycen- tric coordinates, G is the gravitational constant, mi and r i are the mass and the Solar System barycentric position of Solar System body i, ri = r i − r, c is the speed of light in vacuum, and β and γ are the parametrized post-Newtonian parameters measuring the nonlinearity in superposition of gravity and space curvature produced by unit rest mass (Sei- delmann 1992). The position and velocity vectors of all ce- lestial bodies are computed with NASA's SPICE library2. We used the planetary and lunar ephemeris DE432s. The N bodies include the Sun, the planets and the Moon. For plan- ets with moons, with the exception of the Earth, the centre of mass of the system is considered. The dynamical model is written in the J2000 ecliptic reference frame. Figure 6 shows the geocentric distance profile in time for one thousand samples from the initial Gaussian distri- bution. As expected, the uncertainties significantly increase after 2029 and pave the way to resonant returns in 2036 and 2037. 2 http://naif.jpl.nasa.gov/naif/toolkit.html MNRAS 000, 1 -- 19 (2017) The authors showed an analysis of the performance of the ADS algorithm for the propagation of the whole set up to the second resonant return in Wittig et al. (2015). The results are now limited to the first resonant return, and they will be used as a reference for the assessment of the perfor- mance of the ADP. All the results presented in this section are obtained considering an expansion order equal to 8, a tolerance for the splitting procedure equal to 10−10, a value of Nmax equal to 12 and an initial uncertainty set with 3σ boundaries, i.e. a 6-dimensional (6D) rectangle with 3σ boundaries. The initial uncertainty set should be properly selected, as the neglected part of the probability mass, i.e. the in- tegral of the pdf over the domain outside the considered box, could significantly alter the estimated impact probabil- ity. For the case under study, in which we are considering a 6-dimensional problem with uncorrelated variables, the se- lection of a 6D rectangular domain with 3σ boundaries cor- responds to considering the 98.4 per cent of the probability mass, and so the estimated impact probability may result underestimated. The accuracy of the estimate improves for larger initial uncertainty sets. A detailed sensitivity analysis on the uncertainty set size and all the other available param- eters is offered in Section 7. All computations are performed on a single core Intel i7-3770 CPU @3.4 GHz, 16 GB RAM processor. The number of subdomains obtained with ADS propa- gation without pruning is 653, while the computational time is 10 h 6 min. An analysis of the average number of splits per direction shows that most splits occur in the semi major axis (a) and true longitude (l) directions (Wittig et al. 2015). Thus, though the problem is six dimensional, the analysis on the dynamics can be focused on the projection onto the a− l plane of the initial conditions. Figure 5 shows the projection of the initial uncertainty box onto the a − l plane, along with the subdomains gen- erated during the ADS propagation. Colours refer to the truncation epoch of the related subset: white regions rep- resent subsets that were able to reach the final simulation time (May 31, 2036, after the expected resonant return), coloured regions represent subsets whose propagation was stopped earlier because they reached their minimum box size. Figure 5 can be exploited to easily identify the regions of the initial set that are involved in the resonant return in 2036. While all initial conditions lying within white regions have no risk to impact the Earth, coloured subdomains rep- resent sets of initial conditions that might lead to close en- counters with Earth at that epoch. That is, coloured regions represent PHS's. This behaviour is expected, as splits occur when the nonlinearities increase, which happens when tra- jectories get closer to Earth. It is evident, however, that a significant portion of the computational effort required by the ADS propagation is spent on regions of the initial set that are not involved in the first resonant return. Thus, the application of a selective pruning technique as the ADP aims at alleviating this inefficiency. We now investigate the performance of the ADP method. The first part of the analysis is represented by the propagation of the uncertainty set up to the epoch of the first close encounter in 2029. The DA propagation of the whole uncertainty set up to the close encounter in 2029 is performed with no splits. Therefore, the whole set can be + + − (cid:41) 8 M. Losacco et al. Figure 5. Projection of the generated subsets onto the a − l plane of the initial conditions (ADS propagation, order 8, tolerance 10−10, Nmax 12, 3σ domain). T(cid:48) = 7/6T⊕, ∆Tref is determined by setting a value of ε equal to 10−3. The value of ε is selected in order to take into ac- count small perturbations between the close encounter in 2029 and the resonant return in 2036. An analysis of the impact of ε on the results is carried out in Section 7.2. The propagation is then resumed as described in Section 5. Fig- ure 7 shows the results of the ADP propagation in terms of subdomains distribution on the a − l plane. A comparison with Fig. 5 clearly shows how the ADP restricts the propa- gation of the generated subdomains to a limited portion of the PHS's. That is, only subsets that are actually involved in the resonant return in 2036 are propagated till the end of the simulation. The pattern of subdomains is not altered by the introduction of the pruning. Simply, a large portion of the initial set is no longer investigated. This action has a strong impact on the number of propagated subdomains, that is now significantly lower (267). Consequently, the com- putational time required by the propagation reduces signif- icantly (4 h 6 min). The pattern of generated subdomains represents the starting point for the second phase, the application of the IS method for the computation of the impact probability in 2036. We initialize the method by defining a uniform pdf including all the generated PHS's as ISD. The boundaries of the ISD on the a − l plane are represented in blue in Fig. 7. Then, samples are drawn from the ISD and each sample is as- sociated with a PHS if possible. For samples belonging to the PHS's, the state vector corresponding to the drawn sample at the truncation epoch of the related PHS is reconstructed, and a pointwise propagation up to the epoch of minimum geocentric distance is performed. Figure 8 shows a focus of the resulting subsets, whereas Fig. 9 shows the pattern of generated samples projected onto the a − l plane. Samples belonging to the PHS's are represented in blue, whereas im- pacting samples are represented in yellow. Black dots repre- sent discarded samples. Not all samples belong to the PHS's, due to the shape of the selected ISD. A uniform ISD over a domain of regular shape enclosing all PHS's represents the MNRAS 000, 1 -- 19 (2017) Figure 6. Geocentric distance profile up to January 21, 2038 for one thousand samples from the initial uncertainty set. described with a single polynomial map at the epoch of the first close encounter. The availability of the DA state vector of the asteroid, then, provides the polynomial expansion of its perturbed orbital period immediately after the close en- counter with the Earth. This polynomial expansion allows us to estimate the asteroid orbital period range after the close encounter by means of a polynomial bounder: for the case under study, this range is equal to [415.02, 428.91] days. By looking at this range, we can identify the first resonances: Tres1 = 7/6T⊕ = 426.12 days (where T⊕ is the Earth or- bital period) is the first resonant orbital period included in the computed range, and it represents a resonant return in 2036. This value is expected, as shown in Fig. 6. We can also notice that the expected second resonant return (in 2037, resonance 8:7), is also included (Tres2 = 417.43 days). The a priori identification of the resonances and the application of the ADP -- IS method is strictly related to the assumption of no intervening close encounter with other ma- jor bodies in between. This assumption is checked imme- diately after the resonances computation, as explained in Section 7.2. For the case under study, the assumptions are verified. Therefore, we can now concentrate the analysis on the first resonant return, in 2036. Given a nominal value −6−4−20246−3−2−10123δa(AU)δl(rad) Date (TDB)(cid:10)(cid:10)(cid:10)(cid:10)(cid:10)(cid:10)(cid:10)x 10−8x 10−62036−04−03.02036−04−11.02036−04−21.02036−05−01.02036−05−11.02036−05−21.02036−05−31.000.511.522.5Geocentric distance (AU)Date (TDB)2021−11−26.02027−05−19.02038−01−21.02032−11−08.02010−12−14.02016−06−05.0 ADP -- IS method for impact probability computation 9 Figure 7. Projection of the generated subsets onto the a − l plane of the initial conditions (ADP propagation, order 8, tolerance 10−10, Nmax 12, 3σ domain). In blue, boundaries of the ISD. Table 2. Overall results for the ADP -- IS method (order 8, tolerance 10−10, Nmax 12, 3σ domain). nPHS tADP nsamples tIS tall 267 4 h 6 min 204293 26 min 4 h 32 min p σ( p) 1.17 · 10−5 2.93 · 10−6 Figure 8. Projection of the generated subdomains onto the a − l plane (detail of Fig. 7). Figure 9. Projection of the generated samples onto the a − l plane. In black, discarded samples. In blue, samples belonging to the PHS's. In yellow, impacting samples. easiest choice and can be applied regardless the complexity of the PHS's pattern. On the other side, this selection leads to the black dots shown in Fig. 9. These samples, however, have a minimal impact on the computational effort required by the method, as they are discarded as soon as they are identified. The selection of the IS method as sampling technique al- lows us to increase significantly the number of samples lying within the PHS's with respect to a standard Monte Carlo ap- proach, and this advantage is made possible by the pruning action of the ADP propagation. The analysis of the distri- bution of the impacting samples on the a− l plane, however, shows that these are confined to a limited region inside the PHS's. That is, not all PHS's actually give a contribution MNRAS 000, 1 -- 19 (2017) −6−4−20246−3−2−10123δa(AU)δl(rad) Date (TDB)(cid:10)(cid:10)(cid:10)(cid:10)(cid:10)(cid:10)(cid:10)(cid:10)(cid:10)(cid:10)(cid:10)(cid:10)(cid:10)(cid:10)x 10−6x 10−82036−04−03.02036−04−11.02036−04−21.02036−05−01.02036−05−11.02036−05−21.02036−05−31.0−6.5−6−5.5−5−3−2−10123δa(AU)δl(rad) Date (TDB)(cid:10)(cid:10)(cid:10)(cid:10)(cid:10)(cid:10)(cid:10)(cid:10)(cid:10)(cid:10)(cid:10)(cid:10)(cid:10)(cid:10)2036−04−11.02036−04−21.02036−05−01.02036−05−11.02036−05−21.02036−05−31.0x 10−6x 10−82036−04−03.0 10 M. Losacco et al. Figure 10. Estimated impact probability for Apophis resonant return 2036 as a function of the number of samples. In yellow, im- pacting samples. In grey, estimated Poisson statistics uncertainty (1σ). to the impact probability in 2036. This result is related to the selection of the amplitude of ∆Tref: the value was set in order to grant a conservative pruning action on the subsets. A more detailed analysis is offered in Section 7.2. The trend of the estimated impact probability with the number of drawn samples is represented in Fig. 10. Impact- ing samples are represented with yellow circles. The toler- ance for the stopping criterion was set equal to 0.01 per cent. After some initial significant oscillations, the impact prob- ability asymptotically converges to the value of 1.17 · 10−5. The estimate is of the same order of magnitude of the refer- ence value (2.2·10−5) obtained with a standard Monte Carlo analysis, though slightly lower (see Section 8). This differ- ence can be explained considering the size of the propagated uncertainty set, as later explained in Section 7.3. An overview of the main results of the simulation is shown in Table 2. Results are expressed in terms of num- ber of PHS's nPHS, computational time required by the ADP propagation tADP, number of generated samples for the IS method at convergence nsamples, computational time required by the IS method tIS, overall computational time tall, estimated impact probability value p and related Pois- son statistics uncertainty σ( p). 7 SENSITIVITY ANALYSIS The analysis presented in Section 6 was carried out starting from predefined values of expansion order, tolerance for the splitting routine, maximum number of splits and size of the uncertainty set. In this section, we investigate the role of the different parameters and provide some guidelines for the selection of the most appropriate set of parameters. The dis- cussion is carried out dividing parameters mostly affecting the ADP propagation (order, tolerance, minimum box size and reference orbital period range) and parameters affecting the estimated impact probability (uncertainty box size). 7.1 Selection of splitting tolerance, expansion order and Nmax As described in Section 2, the main parameters for the ADS propagation are the tolerance for the splitting procedure, Table 3. Average error in position as a function of the imposed tolerance for subsets at the epoch of the first resonant return (ADP propagation, order 8, Nmax 12, 3σ initial uncertainty set). Tolerance Error 10−8 10−9 10−10 4.83 · 10−6AU 1.24 · 10−6AU 5.95 · 10−7AU the expansion order and the maximum number of splits. The selection of the tolerance is strictly related to the accu- racy required in the description of the subsets at the end of the simulation. This concept is valid in both ADS and ADP propagation. Due to error accumulation during the integra- tion process, indeed, the actual accuracy of the ADP result tends to decrease with respect to the imposed accuracy. This effect becomes more significant as the nonlinearities of the dynamics increase, so that, in order to grant a specific ac- curacy, the imposed tolerance must be in some cases some orders of magnitude lower. Table 3 shows the average accuracy in position for the subsets at the epoch of the first resonant return consider- ing an expansion order equal to 8 and decreasing values of tolerance. We estimated the accuracy by comparing the re- sults of pointwise propagations and polynomial evaluations for random samples drawn in the generated subsets. The er- ror in position shown in Table 3 represents an average of the computed errors. As expected, there is a difference of around three orders of magnitude with respect to the imposed tol- erance. For the case under study, the error is strictly related to the intervening close encounter in 2029. As an example, if we perform an ADS propagation with order 8 and tolerance 10−10, and we stop the propagation three months before the close encounter in 2029, we obtain a position error of about 10−11 AU. This error expands to 10−8 AU six months later, i.e. three months after the close encounter. That is, the close encounter yields an increase of about 3 orders of magnitude in the position error. As the propagation continues, the error accumulates and reaches 5.95 · 10−7 AU at the epoch of the first resonant return. Therefore, the splitting tolerance is a critical parameter and its selection must account for all the above aspects. For our analysis, we selected a tolerance capable of granting a maximum error in position of 100 km. This requirement results into a splitting tolerance of at least 10−10. Expansion order and minimum box size, instead, play quite different roles in ADS and ADP propagation. During a DA propagation, a reduction of the expansion order causes a decrease in the accuracy of the results at a specific integra- tion epoch. This decreased accuracy yields an increase in the required number of splits during the ADS propagation and, overall, a larger number of generated subsets. The role of the minimum box size, instead, is to limit the number of splits, so that, overall, both parameters have a strong influence on the number of generated subdomains and, as a consequence, on the required computational effort. The role of the expan- sion order is twofold, since a decrease in the order causes the number of subdomains to increase, but reduces the compu- tational time required to perform a single integration step. MNRAS 000, 1 -- 19 (2017) 00.20.40.60.811.21.41.61.82Number of samples00.20.40.60.81x 105x 10-4 Table 4. Performance of the ADP -- IS method for different expansion orders (tolerance 10−10, Nmax 12, 3σ initial uncertainty set). ADP -- IS method for impact probability computation 11 Order nPHS tADP nsamples tIS tall 8 7 6 5 4 3 267 267 267 267 267 589 4 h 6 min 1 h 53 min 49 min 24 min 28 min 8 min 204293 204293 204293 204293 204293 273035 26 min 24 min 23 min 23 min 25 min 4 h 32 min 2 h 17 min 1 h 12 min 47 min 53 min 9 h 47 min 9 h 55 min p 1.17 · 10−5 1.17 · 10−5 1.17 · 10−5 1.17 · 10−5 1.17 · 10−5 8.60 · 10−6 σ( p) 2.93 · 10−6 2.93 · 10−6 2.93 · 10−6 2.93 · 10−6 2.93 · 10−6 2.20 · 10−6 Table 5. Performance of the ADP -- IS method for different values of Nmax (order 5, tolerance 10−10, 3σ initial uncertainty set). Nmax nPHS tADP nsamples tIS tall 12 11 10 9 267 148 84 47 24 min 17 min 12 min 9 min 204293 204293 204293 204293 23 min 26 min 29 min 34 min 47 min 43 min 41 min 43 min p 1.17 · 10−5 1.17 · 10−5 1.17 · 10−5 1.17 · 10−5 σ( p) 2.93 · 10−6 2.93 · 10−6 2.93 · 10−6 2.93 · 10−6 Table 6. Trend of the computational times tADP and tIS for increasing (↑) values of the expansion order and minimum box size. Order ↑ Nmax ↑ ADP tADP ↑↓ tADP ↑ IS tIS ↑↓ tIS ↓ Thus, it is reasonable to imagine that there exists a specific expansion order capable of minimizing the computational ef- fort required by the ADS propagation. This value, obviously, changes according to the specific case under study. The role of the minimum box size, instead, is univocal: by increasing the value of Nmax, the computational effort required by the ADS propagation increases. In the case of ADP propagation, the analysis is quite different. The role of the two parameters for the two phases is reported in Table 6. More specifically, the ADP propa- gation aims to select only subsets whose integration stops before the resonant return of interest having reached their minimum box size. A change in the expansion order mod- ifies the splitting history, which could, but not necessarily would, modify the overall number of splits. This behaviour has a direct impact on the required computational time, though the description of the role of the expansion order is not immediate. A decrease in the expansion order, indeed, may cause just earlier splits performed with the same split- ting sequence, or a complete change in the splitting history. In the first case, the role of the expansion order becomes uni- vocal: a reduction in the expansion order causes a decrease in the computational effort. In the second case, the changes in the splitting history and the number of generated sub- sets may be so relevant that what is gained in performing single integration steps may be lost in the longer propaga- tion of the generated subsets. Overall, the role of the order in not univocal, and it exists an order that minimizes the computational time required by the ADP propagation. The role of the minimum box size, instead, is the same as in the ADS propagation: a decrease in the value of Nmax MNRAS 000, 1 -- 19 (2017) causes an earlier stop of the propagation of the subsets, and a reduction of the computational effort. The whole procedure, however, includes both an ADP propagation and a sampling phase, and the role played by the two parameters during the sampling phase is different from the ADP phase. The role of the expansion order is, again, twofold: a reduction of the order causes longer point- wise propagations, but faster polynomial evaluations. The relative weight of the two effects essentially depends on the number of required samples. The role of the minimum box size, instead, is univocal and opposite with respect to the ADP propagation: a reduction of the value of Nmax implies longer pointwise propagations. The selection of the best combination of order and min- imum box size, therefore, relies on all these aspects. Starting from these considerations, we performed a sensitivity anal- ysis in order to quantify the impact of the two parameters on the performance of the ADP -- IS method for the case of the first resonant return of asteroid Apophis. The results of a sensitivity analysis on the expansion order are shown in Table 4, considering six different expan- sion orders. The comparison is performed by considering the same parameters of the analysis presented in Section 6. The second column shows the number of generated subdomains. The value is not affected by the expansion order till order 4, while for order 3 the value is more than doubled. This trend can be explained looking at the splitting history. For orders from 8 to 4, no split occurs before the close encounter in 2029, and the sequence of splits is exactly the same, though single splits are performed at different epochs. Things com- pletely change with order 3, with 4 splits occurring before the 2029 close encounter. This change has a direct impact on the number of generated subsets. A difference can be de- tected also by looking at the required number of samples or at the estimated impact probability and related Poisson statistics uncertainty. Assuming not to alter the sequence of generated samples, values obtained with orders 8 to 4 are identical, whereas values obtained with order 3 are slightly different. The expansion order has a significant impact on the computational effort required by the ADP and sampling 12 M. Losacco et al. Figure 11. Computational time versus expansion order (ADP -- IS method, tolerance 10−10, Nmax 12, 3σ uncertainty domain). Grey bars and white bars represent computational times required by the ADP propagation and IS phase respectively. phases (columns 3 and 5). As expected, the trends are not monotonic. Thus, it is possible to identify order 5 as the expansion order capable of minimizing the overall computa- tional time. Once again, it is interesting to see what happens with order 3: the early splits at the epoch of the 2029 close encounter completely change the splitting history, causing subsets to stop much earlier than what happens with larger orders. Unlike orders from 8 to 4, where the subsets are stopped few days before the expected resonant return in 2036, with order 3 the subsets are stopped around 2030, 6 years earlier. This earlier stop grants computational time saving for the ADP propagation, but has a tremendous back- lash for the sampling phase, with each sample propagated for years instead of days. For this reason, the computational time required for the sampling phase is much larger (column 5). The trend of the computational time for the different or- ders is shown in Fig. 11. We limited the analysis in Table 4 to order 3 as the minimum order, as early splits that appear with this order magnify with order 2, leading to 53 subsets generated before the 2029 close encounter. This behaviour exacerbates the limitations previously pointed out for low orders. Moreover, the error estimation procedure described in Section 2 does not work with linear approximation and tends to provide inaccurate estimates with order 2. We performed a similar analysis by considering the ef- fect of the minimum box size on the required computational effort and estimated impact probability. Table 5 shows the results of the analysis considering the optimal expansion or- der identified in Table 4 and the same values of tolerance and uncertainty box size of the previous simulations. As de- scribed before, the role of the minimum box size is univocal in the two phases, though opposite, and this trend is con- firmed by the analysis: a decrease in the value of Nmax causes a reduced computational effort required by the ADP prop- agation but longer pointwise propagations for all samples. For the case under study, Nmax equal to 10 allows us to minimize the required computational effort. As in the previous case, a change in the value of Nmax does not alter the estimated impact probability, though the pattern of generated subsets is now modified. This trend can be explained considering the fact that, a reduction of the value of Nmax generates larger subsets at earlier trunca- tion epochs, but with the same accuracy. Thus, if the drawn samples are fixed, their mapping to the epoch of the first resonant return is essentially the same. That is, the pattern of impacting samples is not altered. As described in the presented analysis, both expansion order and minimum box size influence the performance of the method. In particular, the expansion order plays a key role in the definition of the computational effort required by the method, while the minimum box size has a lower influ- ence. As the method is composed by two phases, we can say that the selection of the order must be done in order to min- imize the computational effort required by the heaviest one. In our method, the ADP propagation plays this role, so that, in order to decrease its impact on the overall computational time, the most effective way is to reduce the expansion order, still limiting as much as possible the number of generated subsets before the first close encounter. 7.2 Definition of ∆Tref As described in Section 3, the reference orbital period range ∆Tref represents the key parameter for the ADP propagation. The range, centred in the selected resonant return period T(cid:48), is defined to account both inaccuracies in the estimation of the orbital period range of the subdomains and small dynamical perturbations between the first close encounter and the predicted resonant return. The semi-amplitude of this range, ε, plays therefore a key role as it influences both the accuracy of the probability estimates and the required computational time. Table 7 shows the performance of the ADP -- IS method for different values of the reference range semi-amplitude ε. With respect to the previous analyses, two additional pa- rameters are shown: the number of generated subdomains nD and the number of subdomains that include impacting samples nimp. These two parameters provide, along with the number of PHS's nPHS, a clear picture of the pruning ac- tion performed during the ADP propagation. We performed the analysis considering four values of ε. In particular, it is interesting to analyse what happens considering the two limiting cases: ε = 0 and ε = 1. In the first case, the reference orbital period range col- lapses to the value of T(cid:48), i.e. subsets are maintained through- out the simulation only if their estimated orbital period range includes T(cid:48). In this case, the number of PHS's is lower than the one obtained with ε = 10−3, but the value of impact probability is exactly the same. This result can be explained considering that the pattern of impacting samples is not al- tered, as confirmed by the parameter nimp. That is, for the case under study, a less conservative selection of the param- eter ε would allow us to obtain the same results, though no evident savings in computational time would be obtained. The selection of ε = 0 corresponds to considering a Kep- lerian motion between the two encounters. As described in Valsecchi et al. (2003), this assumption may provide quite accurate results for the timing, and for the case under study, where no significant perturbations between the two encoun- ters exist, it can be considered acceptable. Let us now analyse the second limit case. If ε is set to 1, we are essentially selecting a very large reference range for the orbital period. That is, the ADP propagation becomes an ADS propagation, i.e. no pruning is performed and all subsets are propagated until the final simulation time or MNRAS 000, 1 -- 19 (2017) 345678Order0246810 Table 7. Performance of the ADP -- IS method for different values of ε (order 5, tolerance 10−10, Nmax 10, 3σ initial uncertainty set). ADP -- IS method for impact probability computation 13 ε nD nPHS nimp tADP nsamples tIS tall 0 10−5 10−3 1 (ADS) 71 71 84 470 71 71 84 121 43 43 43 41 11 min 11 min 12 min 1 h 31 min 204293 204293 204293 216299 29 min 29 min 29 min 31 min 40 min 40 min 41 min 2 h 02 min p 1.17 · 10−5 1.17 · 10−5 1.17 · 10−5 1.11 · 10−5 σ( p) 2.93 · 10−6 2.93 · 10−6 2.93 · 10−6 2.25 · 10−6 the maximum number of splits is reached. The sampling phase, instead, is not significantly altered: the ISD is de- fined including subsets whose propagation stops before the expected resonant return. Results are reported in the last row of Table 7. The value of impact probability is similar to the one obtained with the pruning action, but the number of generated subsets is much larger, which affects in turn the required computational time. That is, a very conservative selection of ε would yield a factor three increase in compu- tational time. v⊕ < 0.05 AU (18) The selection of the parameter ε is therefore crucial. Within the assumptions of our method, i.e. small perturba- tions between the two encounters, we can define an upper threshold for the ε value as hεT(cid:48) where h is the number of revolutions of the asteroid between the encounters, whereas v⊕ is the Earth heliocentric veloc- ity. The expression on the left hand side of the inequality represents the heliocentric arc covered by the Earth in the time range hεT(cid:48). When we define the reference range ∆Tref and we compare it with ∆Tsub of a given subset, therefore, we are verifying that the uncertainty in the position of the current subset with respect to the Earth position is lower than 0.05 AU, i.e. the current subset can be labelled as Po- tentially Hazardous. For the case under study, this value is about 10−3, the value selected for the analysis presented in the previous sections. As previously stated in the paper, the application of the ADP -- IS method is strictly related to the assumption of no intervening close approaches with other major bodies in the investigated time window. In case of expected close approaches, indeed, the situation drastically changes as the resonances estimated at the epoch of the first close encounter may lose their validity. In such cases, one must rely on the more conservative approach of ADS propagation for inves- tigating a selected propagation window, obtaining accurate results with unavoidable drawbacks in efficiency. The decision whether to perform an ADP propagation or disable pruning is made based on a preliminary analy- sis of the possible trajectories of the asteroid between the two encounters. For the case under study, the availability of the dispersion of Apophis' orbital parameters after the first close encounter allows us to estimate the minimum orbit in- tersection distance (MOID) dispersion between the asteroid and the other main bodies of the Solar System (see Armellin et al. (2010a)). This fast survey provides us with an overview of possible close approaches between the two encounters and drives our decision on the propagation method. For the case under study, the analysis required less than one minute and allowed us to exclude any significant encounter with other planets in between. MNRAS 000, 1 -- 19 (2017) 7.3 Effect of the size of the uncertainty domain The analysis presented in the previous sections was done considering different values of expansion order, tolerance for the splitting procedure and minimum box size, whereas we considered only one size for the initial uncertainty set, that is a 3σ 6-dimensional rectangle. This selection, in a 6- dimensional problem with uncorrelated variables, consists in considering the 98.4 per cent of the probability mass. In the following paragraphs, we present the impact of the size of the uncertainty set on the results of the ADP -- IS method. It is worth noting that, as previously mentioned in Section 6, in case of full covariance matrix, the initial uncertainty box would be defined in the eigenvector space in order to avoid wrapping effect and including very low probability solutions, so all the analyses and values presented in this section hold for the more general case of correlated variables. The ADP propagation and the IS phase are strongly in- fluenced by the selection of the size of the initial uncertainty set. The ADP propagation, indeed, limits the generation of the subsets within the boundaries of the considered uncer- tainty set, and this aspect influences also the shape of the ISD for the impact probability computation phase. Samples, indeed, are confined within the initial uncertainty set, im- pacting samples are found only within these limits and pos- sible impacting samples that lie out of the initial uncertainty set are discarded. This aspect distinguishes our sampling ap- proach from a standard Monte Carlo method, where samples are drawn directly from the original probability density func- tion, and the probability of drawing samples is determined by the pdf itself. In principle, samples could lie anywhere in the uncertainty region. Starting from these considerations, it is therefore inter- esting to study how the estimate for the impact probability changes with an increasing size of the initial uncertainty set. We initially performed the analysis by considering a size for the initial uncertainty set of 3σ (i.e. what was previously presented), 4σ and 5σ. Figure 12a shows the results for the ADP propagation considering order 5, tolerance 10−10, Nmax equal to 10 and an initial uncertainty set size of 4σ. The ISD boundaries on the a − l plane are represented in light blue. On the same plot, the 3σ boundaries are represented with black dashed lines, whereas the ISD boundaries for the 3σ case are repre- sented with dashed blue lines. The plot allows us to compare the sample regions in the two cases. In particular, a portion of the PHS's generated during the 4σ ADP propagation is not considered during the 3σ case. The 5σ case is shown in Fig. 12b, representing the 5σ ISD in cyan. Figure 13 shows a comparison of the results of the sam- pling phase for the 3σ, 4σ and 5σ cases. Blue points repre- sent drawn samples belonging to the 3σ case, with impacting 14 M. Losacco et al. (a) (b) Figure 12. Comparison between 4σ (a) and 5σ (b) cases. The figure box and the dashed black lines represent the selected boundaries (4σ in (a) and 5σ in (b)) and the 3σ boundaries respectively. In light blue and cyan, the ISD boundaries for the 4σ and 5σ respectively. In dashed blue, the 3σ ISD boundaries. Figure 13. Comparison of the samples distribution for the 3σ, 4σ and 5σ cases. In blue, light blue and cyan, drawn samples for the 3σ, 4σ and 5σ case respectively. In yellow, black and red, impacting samples for the 3σ, 4σ and 5σ case respectively. samples represented as yellow dots. Light blue points repre- sent drawn samples belonging to the 4σ case, with impact- ing samples represented as black dots. Finally, cyan points represent accepted samples for the 5σ case, with impacting samples represented with red dots. The analysis of the plot offers a clear picture of how the sampling region changes in the three cases. Moreover, it is possible to see how the in- creasing size of the initial uncertainty set allows us to include impacting samples out of the 3σ domain. While a 3σ domain appears as a too narrow selection, the 4σ and 5σ domains offer a better description of the impact region. Figures 14a and 14b show the distribution of the impacting samples for the three different simulations, with colors showing the con- tribution to the overall impact probability. Table 8 shows the results of the analysis, including the 6σ case. With reference to the previous analyses, we added the parameter pout, which represents the probability mass outside the selected uncertainty set, i.e. the complementary to 1 of the integral of the pdf over the considered domain. We remark that, because we use rectangular uncertainty sets, the values of pout shown in Table 8 are significantly smaller than those corresponding to the more commonly used ellip- soidal uncertainty regions, for which pout is equal to 0.17, 1.38 · 10−2, 3.41 · 10−4 and 2.76 · 10−6 for the 3σ, 4σ, 5σ and 6σ cases respectively. By increasing the size of the initial uncertainty set, the computational time required by the ADP propagation in- creases. This trend is essentially due to the larger com- putational effort required by a single integration step and the larger number of subsets. An increase in the number of generated PHS's can be detected passing from 3σ to 4σ, MNRAS 000, 1 -- 19 (2017) −8−6−4−202468−4−2024δa(AU)δl(rad) Date (TDB)(cid:10)(cid:10)(cid:10)(cid:10)(cid:10)(cid:10)(cid:10)2036−04−11.02036−04−21.02036−05−11.02036−05−21.02036−05−31.02036−05−01.02036−03−26.0x 10−8x 10−6−10−8−6−4−20246810−4−2024δa(AU)δl(rad) Date (TDB)(cid:10)(cid:10)(cid:10)(cid:10)(cid:10)(cid:10)(cid:10)2036−05−31.02036−03−17.02036−04−01.02036−05−01.02036−05−16.02036−04−16.02036−03−02.0x 10−6x 10−8 ADP -- IS method for impact probability computation 15 (a) (b) Figure 14. Impacting samples distribution for 4σ (a) and 5σ (b) cases. Colours are referred to the associated contribution to the impact probability. whereas this value remains essentially the same in the 5σ and 6σ cases. The analysis of the sampling phase shows some interest- ing results. As expected, an increase in the initial uncertainty size causes an increase in the estimated impact probability value. Essentially, regions of the uncertainty set that were not studied during the ADP propagation for the 3σ case are now considered, and impacting samples can be found also in these regions. As a result, the estimated impact prob- ability values for the 5σ and 6σ cases become very close to the reference value. The enlargement of the investigated region causes also an increase in the Poisson statistics un- certainty of the estimate. This result is expected too, as the variance is proportional to the sample region volume (see equation (13)). The analysis of the required number of sam- ples at convergence shows that this value increases for larger initial uncertainty sets, and this trend reflects back on the computational time required by the sampling phase. The size of the uncertainty set should be selected to achieve the desired resolution on the impact probability, which is directly expressed by the parameter pout. For the case under study, with an estimated impact probability of the order of 10−5, we selected a 6σ domain, which excludes only 1.18 · 10−8 of the probability mass. 8 COMPARISON WITH STANDARD AND ADVANCED ORBITAL SAMPLING TECHNIQUES The analysis presented in the previous sections showed how the ADP -- IS method represents a valuable tool for uncer- tainty propagation and impact probability computation for the first resonant return of a NEO. In order to assess the efficiency of the method with respect to other impact prob- ability computation tools, we present in this section a com- parison with standard and advanced orbital sampling tech- niques. In the first part, we compare our approach with MNRAS 000, 1 -- 19 (2017) Monte Carlo sampling techniques based on sample gener- ation on the whole uncertainty set. Finally, we present a general comparison with the most used technique for im- pact probability calculation, the LOV method. A first comparison can be done considering a standard Monte Carlo approach, where samples are drawn from the covariance matrix directly at the initial epoch (June 18, 2009). This method is probably the most straightforward approach but also the most expensive one, as the sampling is performed on the whole domain, and the propagation of each sample starts from the observation epoch. By perform- ing the propagation of one million samples, the estimated impact probability results into 2.2 · 10−5 , whereas the Pois- son statistics uncertainty is equal to 4.71 · 10−6. We selected the number of samples in order to detect a non null value of impact probability (Farnocchia et al. 2015). Unfortunately, if the Monte Carlo simulation is per- formed considering the same conditions of our method (i.e same dynamics, single core), the required computational time is much larger. The average computational time re- quired to perform a single pointwise propagation from the initial epoch to the epoch of the first resonant return is ≈ 1.2 s. As a result, within the computation time required by the ADP -- IS method for the 6σ case (see Table 8), about 4400 samples could be propagated, which is not enough to estimate the expected impact probability. All this would lead to an estimated computational time of around two weeks for propagating one million samples on a single core. This value is of course not realistic, as typically Monte Carlo analy- ses can be easily set up in a multi-thread environment, thus granting significant savings in computational time. It is in- teresting, however, to highlight the significant savings that our approach grants with respect to standard MC approach in the same conditions. The ADP -- IS method, indeed, em- ploys a lower number of samples, as samples are drawn just in a subset of the uncertainty set. Moreover, the propaga- tion of all samples starts immediately before the resonant −8−7−6−5−4−2024δa(AU)δl(rad) IF(xk)·qx(xk)/eqx(xk)3σ4σ2.8449e−203.9492x 10−8x 10−6−8−7−6−5−4−6−4−20246δa(AU)δl(rad) IF(xk)·qx(xk)/eqx(xk)(cid:10)3σ5σ2.8449e−203.9492x 10−8x 10−6 16 M. Losacco et al. Table 8. Performance of the ADP -- IS method for different values of initial uncertainty domain size (order 5, tolerance 10−10, Nmax equal to 10). Domain pout nPHS tADP nsamples tIS tall 3σ 4σ 5σ 6σ 1.61 · 10−2 3.80 · 10−4 3.44 · 10−6 1.18 · 10−8 84 107 104 107 12 min 21 min 29 min 46 min 204293 300643 341804 353056 29 min 39 min 41 min 42 min 41 min 1 h 1 h 10 min 1 h 28 min p 1.17 · 10−5 1.61 · 10−5 1.90 · 10−5 2.09 · 10−5 σ( p) 2.93 · 10−6 5.70 · 10−6 6.81 · 10−6 7.00 · 10−6 return, while in a standard Monte Carlo approach each sam- ple is propagated starting from June 18, 2009. Therefore, the computational effort required by the ADP propagation is largely repaid later by shorter pointwise propagations and a reduced number of samples. We show now a comparison with an advanced Monte Carlo technique called Subset Simulation (SS). The basic idea of SS is to compute small failure probabilities as the product of larger conditional probabilities (Au & Beck 2001; Zio & Pedroni 2009; Zuev et al. 2012). Given a target failure event F, let F1 ⊃ F2 ⊃ ... ⊃ Fn = F be a sequence of inter- k = 1, 2, ..., n. Considering a sequence of conditional probabilities, then the failure probability becomes: mediate failure events, so that Fk =k p(F) = p(F1) n−1 p(Fi+1Fi) i=1 Fi, (19) i=1 where P(Fi+1Fi) represents the probability of Fi+1 condi- tional to Fi. A detailed description of the algorithm can be found in Au & Beck (2001). In the problem under study, the failure F represents an impact with Earth, i.e. a geo- centric distance smaller than the Earth radius. The method is initialized using standard MC to generate samples at the so-called conditional level (CL) 0 starting from the avail- able nominal state vector and related uncertainty of the in- vestigated object at the observation epoch. The number of samples generated at this level is maintained for each gen- erated conditional level and it is referred to as N. Once the failure region F1 is identified, a Monte Carlo Markov Chain (MCMC) Metropolis Hastings algorithm is used to generate conditional samples in the identified intermediate failure re- gion. Another intermediate failure region is then located, and other samples are generated by means of MCMC. The procedure is repeated until the target failure region is iden- tified. An illustration of the method is shown in Fig. 15. The approach was originally developed for the identifi- cation of structural failures, but it was also used in different research areas in reliability such as the definition of fail- ure probabilities of thermo-hydraulic passive systems. The method was recently applied to the computation of space debris collisional probabilities by Morselli et al. (2014). In the presented approach, the intermediate failure re- gions are identified by assuming a fixed value of conditional probability p0 = p(Fi+1Fi). The identification of each con- ditional level, therefore, is strictly related to this value, and changes accordingly step by step, as explained in the follow- ings. The resulting SS algorithm follows the general descrip- tion presented in Morselli et al. (2014) and goes through the following steps: (i) Set i = 0 and generate N samples x 0,k 0 , k = 1 , ... , N at conditional level 0 by standard MC starting from the available state estimate of the investigated objects at the initial epoch t0. (ii) Propagate each sample up to the epoch of the first resonant return and compute its minimum geocentric dis- tance. Note that, as in the ADP -- IS method, the resonances can be easily determined by propagating the uncertainty set up to the epoch of the first close encounter by means of DA and evaluating the orbital period range. (iii) Sort the N samples in descending order according to the associated geocentric distance at the epoch of the first resonant return. (iv) Identify an intermediate threshold value Di+1 as the geocentric distance corresponding to the (1− p0)Nth element of the sample list. Define the (i + 1)th conditional level as Fi+1 = {d < Di+1}, where d represents the geocentric dis- tance. According to the definition of Di+1, the associated conditional probability p(Fi+1Fi) = p0. (v) If Di+1 < R⊕, i.e. the geocentric threshold distance is lower than the Earth radius, go the the last step, otherwise select the last p0N samples of the list x i, j j = 1, . . . , p0N. 0 , By definition, these samples belong to the (i + 1)th condi- tional level. (vi) Using MCMC, generate (1 − p0)N additional condi- tional samples starting from the previously selected seeds belonging to Fi+1. A sample is set to belong to Fi+1 accord- ing to the following performance function: > 0 x 0 is out of the(i + 1) th CL = 0 x 0 is at the limit of the CL < 0 x 0 belongs to the(i + 1) th CL (20) x (x 0) = d(x 0)−Di+1 gi+1 (vii) Set i = i + 1 and return to step 2. (viii) Stop the algorithm. The total number of generated samples is (21) NT = N + (n − 1)(1 − p0)N where n is the overall number of conditional levels required to reach the impact region. Since the conditional probability is equal to p0 for each level, the impact probability expressed by equation (19) becomes: p(F) = p(Fn) = p(FnFn−1)pn−1 (22) 0 where Nn is the number of samples belonging to the last conditional level whose geocentric distance is lower than the Earth radius. 0 Nn/N = pn−1 The main degrees of freedom of the method are the se- lected fixed conditional probability p0, the number of sam- ples per conditional level and the proposal auxiliary distri- bution for the MCMC phase, and they govern the accuracy MNRAS 000, 1 -- 19 (2017) ADP -- IS method for impact probability computation 17 (a) (b) Figure 15. Subset Simulation process: (a), initialization by standard MC, (b), CL 1 identification, (c), samples generation by means of MCMC, (d), new iterations and impact region identification. (c) (d) and efficiency of the method (Zuev et al. 2012). We used for our analysis 1000 samples per conditional level and a value of conditional probability equal to 0.1. A normal distribu- tion with spread equal to the original pdf was selected as proposal pdf for the MCMC algorithm. A comparison between the SS technique and the ADP -- IS method is shown in Table 9. The required number of samples, the overall computational time, the estimated im- pact probability and related Poisson statistics uncertainty are shown. Results for the ADP -- IS method are the ones referring to the 6σ case. Subset Simulation and ADP -- IS have a similar compu- tational burden, though the required number of samples is very different. This result is expected, as the propagation windows for the two cases are different. Figure 16 shows the distribution on the a − l plane of the generated condi- tional samples obtained with SS, along with the thresholds per conditional level and related colors. Impacting samples at the last conditional level are represented in black. Condi- tional samples progressively move to the left, until impact- ing samples at conditional level 4 are identified. If compared to Figs. 8-9, this region is practically coincident with the PHS's identified during the ADP propagation. That is, SS and ADP -- IS allow us to identify the same region in two completely independent ways. The advantage of the ADP -- IS method is that, by iden- tifying the PHS's, the propagation of the samples is drasti- cally reduced in time, which yields a similar computational MNRAS 000, 1 -- 19 (2017) burden though the number of generated samples is signif- icantly larger. Potentially, the ADP -- IS method could take advantage of parallelization both during ADP propagation and the sampling phase, while the advantages for SS would be lower, as parallelization could be introduced only for spe- cific phases of the algorithm. This approach would heighten the difference in efficiency between the two methods. How- ever, the great savings granted by the SS could be included in the ADP -- IS method during the sampling phase, by re- placing the standard MC performed in the ISD with a SS limited to the unpruned subsets. This aspect may represent a future development of the method. Overall, the combina- tion of ADP propagation and importance sampling allows us to achieve a computational burden that is competitive with both standard and advanced Monte Carlo techniques. Finally, it is worth comparing the performance of the presented approach with the reference technique in the field of impact probability computation, the LOV method. The LOV method takes advantage of the fact that the orbital uncertainty grows with time by stretching into a long slen- der ellipsoid in Cartesian space (Farnocchia et al. 2015). The tendency of uncertainty to stretch during propagation sug- gests the possibility of a one-dimensional parametrization of the uncertainty region, i.e. the sampling and the generation of the so-called Virtual Asteroids (VAs) is performed along the line of weakness of the orbit determination, and if all orbits are sufficiently close to the LOV, then significant sav- Matteo Losacco, PhD annual evaluation 1𝑥𝑥1𝑥𝑥2Standard MCImpact region 𝐹𝐹Matteo Losacco, PhD annual evaluation 1𝑥𝑥1𝑥𝑥2𝑃𝑃(𝐹𝐹1)=𝑝𝑝0𝐹𝐹1Impact region 𝐹𝐹Standard MCMatteo Losacco, PhD annual evaluation 1𝑥𝑥1𝑥𝑥2𝑃𝑃(𝐹𝐹1)=𝑝𝑝0Impact region 𝐹𝐹𝐹𝐹1Standard MCMatteo Losacco, PhD annual evaluation 1𝑥𝑥1𝑥𝑥2𝑃𝑃(𝐹𝐹1)=𝑝𝑝0Impact region 𝐹𝐹𝐹𝐹1𝐹𝐹2𝑃𝑃(𝐹𝐹2)=𝑝𝑝0𝑃𝑃(𝐹𝐹3)=𝑁𝑁𝑛𝑛/𝑁𝑁Standard MC 18 M. Losacco et al. Table 9. Comparison of the performance of ADP -- IS method (order 5, tolerance 10−10, Nmax equal to 10, 6σ uncertainty set) and SS method (N equal to 1000, p0 equal to 0.1) nsamples 353056 4600 tall p σ( p) 1 h 28 min 1 h 30 min 2.09 · 10−5 2.46 · 10−5 7.00 · 10−6 5.04 · 10−6 ADP -- IS SS Figure 16. Subset Simulation conditional samples projected onto the a − l plane (N equal to 1000, p0 equal to 0.1). In grey, boundaries of the 3σ domain. ings in computational time with respect to a standard Monte Carlo approach are obtained without sacrificing reliability. The analysis presented in Milani et al. (2005) offers a first term of comparison: the generic completion level of 10−7 can be obtained with the propagation of only ∼ 104 VAs. If compared to a standard MC approach, it would lead to com- pute times 3 -- 4 orders of magnitude below those required for similar completeness with MC simulations (Farnocchia et al. 2015). The analysis presented in the previous section showed that the ADP -- IS method grants a reduction in computation burden of around two orders of magnitude with respect to standard MC. Therefore, the LOV shows better performance than the ADS -- IS method in the current implementation. Nevertheless, there are some cases in which the LOV method does not guarantee the same level of accuracy of a standard MC approach. A first case occurs when the ob- served arc of the investigated object is very short, i.e. 1 or 2 days (Milani et al. 2005). In this case, the confidence re- gion is wide in two directions and the unidimensional sam- pling may not be suitable. What happens is that different LOVs, computed with different coordinates, provide inde- pendent sampling and may provide different results. That is, if some impacting samples lie well of the LOV and are separated from it by some strong nonlinearity, then the VAs selected along the LOV may fail to indicate some potential threatening encounters (Milani et al. 2002). In such cases, a standard MC approach would result more reliable, with unavoidable drawbacks in terms of computational time. As presented in this paper, the ADP -- IS method, though main- taining a six-dimensional sampling, allows us to drastically reduce the computational effort by limiting the sampling to just specific regions. For these reasons, the method may be considered as a valuable trade-off between the efficiency of the LOV method and the reliability of standard MC in all those cases in which the former may result inaccurate. The possibility of improving the efficiency of the method by means of parallelization in both ADP propagation and sam- pling phases represents another step in this direction, as well as an optimised coding of the dynamics. 9 CONCLUSIONS This paper introduced the combination of automatic domain pruning and importance sampling for uncertainty propaga- tion and impact probability computation for Earth resonant returns of Near Earth Objects. The automatic domain prun- ing represents an evolution of the DA based automatic do- main splitting technique, it allows us to estimate possible resonances after a planetary close encounter and limit the propagation of an uncertainty set to those subsets that may be involved in the resonant return of interest. During the propagation, the uncertainty domain is divided into sub- sets (Potentially Hazardous Subdomains) whose propaga- tion stops just before the epoch of the resonant return. The identification of PHS's represents the starting point for the sampling phase. An importance sampling probability density function is defined over these subdomains and samples are drawn directly from this auxiliary pdf. We tested the ADP -- IS method on the case of asteroid (99942) Apophis, provid- MNRAS 000, 1 -- 19 (2017) -8-6-4-202468-4-3-2-101234CL 0CL 1 (D1=5.8 107km)CL 2 (D2=1.8 106km)CL 3 (D3=1.3 106km)CL 4 (D4=2.2 104km)Impacting samples at CL 4x 10-6x 10-8 ADP -- IS method for impact probability computation 19 ing an estimate for the impact probability in 2036. We car- ried out a sensitivity analysis on the main parameters of the method, providing general guidelines for their selection. The comparison with a standard Monte Carlo approach showed how the ADP -- IS method can reduce the computation effort by more than two orders of magnitude, still granting the same accuracy level for the impact probability estimate. In addition, the current algorithm can be implemented to make use of parallelization techniques in both the ADP and the IS phase, thus significantly reduce the required computational time. All these considerations suggest that the method may be used as a valuable alternative to standard MC in all those cases in which the LOV method does not guarantee the re- quired level of accuracy. Future developments include a more rigorous formulation of the reference orbital period for sub- sets pruning allowing us to extend the pruning algorithm to the more critical case of intervening close encounters with other celestial bodies between the two encounters, and the testing to a wider set of cases. ACKNOWLEDGEMENTS M. Losacco gratefully acknowledges professors E. Zio, N. Pedroni and F. Cadini from Politecnico di Milano for their introduction to the importance sampling and subset simulation techniques. In addition, the authors are grateful to the reviewer Davide Farnocchia for the constructive comments that significantly improved the manuscript. This is a pre-copyedited, author-produced PDF of an article accepted for publication in Monthly Notices of the Royal Astronomical Society following peer review. The ver- sion of record Matteo Losacco, Pierluigi Di Lizia, Roberto Armellin, Alexander Wittig; A differential algebra-based importance sampling method for impact probability com- putation on Earth resonant returns of near-Earth objects, Monthly Notices of the Royal Astronomical Society, Volume 479, Issue 4, 1 October 2018, Pages 5474-5490 is available online at: https://doi.org/10.1093/mnras/sty1832. REFERENCES Armellin R., Di Lizia P., Berz M., Makino K., 2010a, Celestial Mechanics and Dynamical Astronomy, 107, 377 Armellin R., Di Lizia P., Bernelli Zazzera F., Berz M., 2010b, Celestial Mechanics and Dynamical Astronomy, 107, 451 Au S.-K., Beck J. L., 2001, Probabilistic Engineering Mechanics, 16, 263 Berz M., 1986, techreport AT-6:ATN-86-16, The new method of TPSD algebra for the description of beam dynamics to high orders. Los Alamos National Laboratory Berz M., 1987, Nuclear Instruments & Methods in Physics Re- search, Section A: Accelerators, Spectrometers, Detectors, and Associated Equipment, A258, 431 Berz M., 1999, Modern Map Methods in Particle Beam Physics, 1 edn. Advances in Imaging and Electron Physics, Academic Press Chesley S. R., 2005, in Proceedings of the International Astro- nomical Union Symposium. pp 215 -- 228 Chodas P. W., Yeomans D. K., 1999, in AAS/Division of Dynam- ical Astronomy Meeting. p. 1227 Di Lizia P., 2008, phdthesis, Politecnico di Milano Di Lizia P., Armellin R., Lavagna M., 2008, Celestial Mechanics and Dynamical Astronomy, 102, 355 Farnocchia D., Chesley S. R., Milani A., Gronchi G. F., Chodas P. W., 2015, in Michel P., DeMeo F. E., Bottke W. F., eds, , Asteroids IV, 1 edn, University of Arizona, pp 815 -- 834 Gronchi G. F., Milani A., 2001, Icarus, 152, 58 Kolchin E. R., 1973, Differential Algebra and Algebraic Groups. Academic Press Milani A., 1999, Icarus, 137, 269 Milani A., Chesley S. R., Valsecchi G. B., 1999, Astronomy and Astrophysics, 346, L65 Milani A., Chesley S. R., Valsecchi G. B., 2000a, Planetary and Space Science, 48, 945 Milani A., Chesley S. R., Boattini A., Valsecchi G. B., 2000b, Icarus, 145, 12 Milani A., Chesley S. R., Chodas P. W., Valsecchi G. B., 2002, in Bottke Jr. W. F., Cellino A., Paolicchi P., Binzel R. P., eds, , Asteroids III, 1 edn, The University of Arizona Press, pp 55 -- 69 Milani A., Chesley S. R., Sansaturio M. E., Tommei G., Valsecchi G. B., 2005, Icarus, 173, 362 Morselli A., Armellin R., Di Lizia P., Bernelli Zazzera F., 2012, in 63rd International Astronautical Congress 2012 (IAC 2012). Morselli A., Armellin R., Di Lizia P., Bernelli Zazzera F., 2014, Advances in Space Research, 55, 311 Rasotto M., Morselli A., Wittig A., Massari M., Di Lizia P., Armellin R., Valles C. Y., Ortega G., 2016, in 6th Interna- tional Conference on Astrodynamics Tools and Techniques (ICATT). Risch R. H., 1969, Transactions of the American Mathematical Society, 139, 167 Risch R. H., 1970, Bulletin of the American Mathematical Society, 76, 605 Ritt J. F., 1932, Differential Equations From the Algebraic Stand- point. American Mathematical Society Ritt J. F., 1948, Integration in finite terms: Liouville's theory of elementary methods. Columbia University Press Seidelmann P. K., 1992, Explanatory Supplement to the Astro- nomical Almanac. University Science Books Valli M., Armellin R., Di Lizia P., Lavagna M., 2013, Journal of Guidance, Control and Dynamics, 36, 48 Valsecchi G. B., Milani A., Gronchi G. F., Chesley S. R., 2003, Astronomy and Astrophysics, 408, 1179 Wittig A., Di Lizia P., Armellin R., Bernelli Zazzera F., Makino K., Berz M., 2014b, in 24th AAS/AIAA Space Flight Mechan- ics Meeting. Wittig A., Di Lizia P., Armellin R., Bernelli Zazzera F., Makino K., Berz M., 2014a, in 2nd IAA Conference on Dynamics and Control of Space Systems. Wittig A., Di Lizia P., Armellin R., Makino K., Bernelli Zazzera F., Berz M., 2015, Celestial Mechanics and Dynamical As- tronomy, 122, 239 Yeomans D. K., Chodas P. W., 1994, in Gehrels T., Matthews M. S., Schumann A. M., eds, Hazards Due to Comets and Asteroids. p. 241 Zio E., 2013, The Monte Carlo Simulation Method for System Reliability and Risk Analysis, 1 edn. Springer Series in Reli- ability Engineering, Springer Zio E., Pedroni N., 2009, in Proceedings of the European Safety and RELiability (ESREL) 2009 Conference. pp 687 -- 694 Chodas P. W., 1993, Bulletin of the American Astronomical So- Zuev K. M., Beck J. L., Au S.-K., Katafygiotis L. S., 2012, Com- ciety, 25, 1236 puters and Structures, 92-92, 283 MNRAS 000, 1 -- 19 (2017) 20 M. Losacco et al. This paper has been typeset from a TEX/LATEX file prepared by the author. MNRAS 000, 1 -- 19 (2017)
1610.01304
1
1610
2016-10-05T08:17:47
Jupiter's Para-H$_2$ Distribution from SOFIA/FORCAST and Voyager/IRIS 17-37 $\mu$m Spectroscopy
[ "astro-ph.EP" ]
Spatially resolved maps of Jupiter's far-infrared 17-37 $\mu$m hydrogen-helium collision-induced spectrum were acquired by the FORCAST instrument on the Stratospheric Observatory for Infrared Astronomy (SOFIA) in May 2014. Spectral scans in two grisms covered the broad S(0) and S(1) absorption lines, in addition to contextual imaging in eight broad-band filters (5-37 $\mu$m) with spatial resolutions of 2-4". The spectra were inverted to map the zonal-mean temperature and para-H$_2$ distribution ($f_p$, the fraction of the para spin isomer with respect to the ortho spin isomer) in Jupiter's upper troposphere (the 100-700 mbar range). We compared these to a reanalysis of Voyager-1 and -2 IRIS spectra covering the same spectral range. Para-H$_2$ increases from equator to pole, with low-$f_p$ air at the equator representing sub-equilibrium conditions (i.e., less para-H$_2$ than expected from thermal equilibration), and high-$f_p$ air and possible super-equilibrium at higher latitudes. In particular, we confirm the continued presence of a region of high-$f_p$ air at high northern latitudes discovered by Voyager/IRIS, and an asymmetry with generally higher $f_p$ in the north than in the south. We note that existing collision-induced absorption databases lack opacity from (H$_2$)$_2$ dimers, leading to under-prediction of the absorption near the S(0) and S(1) peaks. There appears to be no spatial correlation between para-H$_2$ and tropospheric ammonia, phosphine and cloud opacity derived from Voyager/IRIS at mid-infrared wavelengths (7-15 $\mu$m). We note, however, that para-H$_2$ tracks the similar latitudinal distribution of aerosols within Jupiter's upper tropospheric and stratospheric hazes observed in reflected sunlight, suggesting that catalysis of hydrogen equilibration within the hazes (and not the main clouds) may govern the equator-to-pole gradient. [Abridged]
astro-ph.EP
astro-ph
Jupiter's Para-H2 Distribution from SOFIA/FORCAST and Voyager/IRIS 17-37 µm Spectroscopy Leigh N. Fletchera, I. de Paterb, W.T. Reachc, M. Wongb, G.S. Ortond, P.G.J. Irwine, R.D. Gehrzf cStratospheric Observatory for Infrared Astronomy, Universities Space Research Association, Mail Stop 232-11, NASA Ames Research Center, Moffett Field, CA aDepartment of Physics & Astronomy, University of Leicester, University Road, Leicester, LE1 7RH, UK bUniversity of California, Berkeley, Astronomy Dept., 601 Campbell Hall, Berkeley, CA 94720-3411, USA dJet Propulsion Laboratory, California Institute of Technology, 4800 Oak Grove Drive, Pasadena, CA, 91109, USA eAtmospheric, Oceanic & Planetary Physics, Department of Physics, University of Oxford, Clarendon Laboratory, Parks Road, Oxford, OX1 3PU, UK fMinnesota Institute for Astrophysics, School of Physics and Astronomy, 116 Church Street, S.E., University of Minnesota, Minneapolis, MN 55455, USA. 94035, USA. 6 1 0 2 t c O 5 . ] P E h p - o r t s a [ 1 v 4 0 3 1 0 . 0 1 6 1 : v i X r a Abstract Spatially resolved maps of Jupiter's far-infrared 17-37 µm hydrogen-helium collision-induced spectrum were acquired by the FORCAST instrument on the Stratospheric Observatory for Infrared Astronomy (SOFIA) in May 2014. Spectral scans in two grisms covered the broad S(0) and S(1) absorption lines, in addition to contextual imaging in eight broad-band filters (5-37 µm) with spatial resolutions of 2-4". The spectra were inverted to map the zonal-mean temperature and para-H2 distribution ( fp, the fraction of the para spin isomer with respect to the ortho spin isomer) in Jupiter's upper troposphere (the 100-700 mbar range). We compared these to a reanalysis of Voyager-1 and -2 IRIS spectra covering the same spectral range. Tropospheric temperature contrasts match those identified by Voyager in 1979, within the limits of temporal variability consistent with previous investigations. Para-H2 increases from equator to pole, with low- fp air at the equator representing sub-equilibrium conditions (i.e., less para-H2 than expected from thermal equilibration), and high- fp air and possible super-equilibrium at higher latitudes. In particular, we confirm the continued presence of a region of high- fp air at high northern latitudes discovered by Voyager/IRIS, and an asymmetry with generally higher fp in the north than in the south. Far-IR aerosol opacity is not required to fit the data, but cannot be completely ruled out. We note that existing collision-induced absorption databases lack opacity from (H2)2 dimers, leading to under-prediction of the absorption near the S(0) and S(1) peaks. There appears to be no spatial correlation between para-H2 and tropospheric ammonia, phosphine and cloud opacity derived from Voyager/IRIS at mid-infrared wavelengths (7-15 µm). We note, however, that para-H2 tracks the similar latitudinal distribution of aerosols within Jupiter's upper tropospheric and stratospheric hazes observed in reflected sunlight, suggesting that catalysis of hydrogen equilibration within the hazes (and not the main clouds) may govern the equator-to-pole gradient, with conditions closer to equilibrium at higher latitudes. This gradient is superimposed onto smaller-scale variations associated with regional advection of para-H2 at the equator and poles. Keywords: Jupiter, Atmospheres, composition, Atmospheres, dynamics 1. Introduction Far-infrared (IR) spectra of the giant planets are shaped by the collision-induced absorption of hydrogen and helium, providing a sensitive measure of the atmospheric temperature structure and the abundances of the most common gases in gas giant atmospheres (e.g., Hanel et al., 1979; Conrath and Gau- tier, 1980; Conrath and Pirraglia, 1983; Conrath et al., 1998; Conrath and Gautier, 2000). However, atmospheric studies in this spectral range are hampered by several factors. Water vapour in Earth's atmosphere restricts our ability to measure the spectrum from the ground to narrow windows (known as the Q-band) between 17 and 24 µm. Furthermore, diffraction- limited spatial resolutions worsen with increasing wavelength, meaning that observatories with large-diameter primary mirrors Email address: [email protected] (Leigh N. Fletcher) Preprint submitted to Icarus are required to resolve spatial contrasts. Finally, it is difficult to disentangle the competing effects of a planet's temperature, aerosol distribution, helium abundance and the specific compo- sition of hydrogen (para- or ortho-hydrogen) on the shape of far-IR spectrum. To date, our only knowledge of the spatial variability of Jupiter's 17-55 µm spectrum comes from the In- frared Interferometer, Spectrometer and Radiometer (IRIS) ex- periments on the twin Voyager spacecraft (Burke, 1974; Hanel et al., 1977). Cassini did not have the spatial resolution to pro- vide maps of Jupiter in this spectral range (Flasar et al., 2004), and ISO could only provide disc-integrated views of the planet (Kessler et al., 1996). In this study, we use the Faint Object in- fraRed CAmera for the SOFIA Telescope (FORCAST, Adams et al., 2010) on the Stratospheric Observatory for Infrared As- tronomy (SOFIA, Young et al., 2012) to provide the first Earth- based spatially-resolved maps of Jupiter's full 17-37 µm spec- trum for comparison with Voyager (previous ground-based ef- forts have been restricted to wavelengths shortward of ∼ 24µm). November 8, 2018 The maps allow us to assess Jupiter's tropospheric tempera- tures and the distribution of para-hydrogen as a tracer for atmo- spheric circulation. Crucially, SOFIA flies at altitudes between 11-14 km, above 99% of the Earth's water vapour, opening up the 17-37 µm region for infrared astronomy. Voyager 1 and 2 flew by Jupiter in March and July 1979, re- spectively, providing spatially-resolved infrared spectra in the 180-2500 cm−1 region (4-55 µm). In the decades that followed, these data have been analysed by many authors to understand Jupiter's latitudinal temperature structure (Hanel et al., 1979; Conrath and Gautier, 1980; Conrath et al., 1981b; Conrath and Pirraglia, 1983; Conrath and Gierasch, 1984; Gierasch et al., 1986; Carlson et al., 1992; Conrath et al., 1998) and the ther- mal contrasts associated with discrete features like the Great Red Spot (Conrath et al., 1981a; Flasar et al., 1981; Griffith et al., 1992; Sada et al., 1996; Simon-Miller et al., 2002; Read et al., 2006). Compositional studies focused on Jupiter's bulk helium abundance (Gautier et al., 1981), ammonia distribu- tion (Gierasch et al., 1986), water ice signatures (Simon-Miller et al., 2000) and hydrocarbon distributions (Nixon et al., 2010), highlighting the richness of the Voyager dataset. This work focusses on the subrange of Jupiter's far-infrared spectrum (270-600 cm−1, 17-37 µm) that is accessible from SOFIA. This range is dominated by the broad collision-induced H2 S(0) and S(1) features near 354 cm−1 (28.2 µm) and 587 cm−1 (17.0 µm), respectively. By measuring both features si- multaneously, we can estimate the relative abundances of the two spin isomers of hydrogen - S(1) is formed from transi- tions within ortho-H2 (the odd spin state of H2 with parallel spins), whereas S(0) is formed from transitions within para-H2 (the even spin state of H2 with anti-parallel spins). Populat- ing or depopulating the S(0) states therefore affects the shape and gradient of the far-IR continuum in a way that is distin- guishable from temperature changes, which tend to scale the spectral continuum brightness more uniformly over the entire wave band. In 'normal' hydrogen, at the high temperatures of the deep atmosphere, the two isomers should be in a 3:1 ratio (a para-H2 fraction fp of 0.25, Massie and Hunten, 1982). At the colder temperatures near the tropopause, the equilibrium fp increases to near 0.35, with lower temperatures favouring the increased population of the J = 0 rotational state, the lower en- ergy state of the S(0) transition. However, vertical mixing of air parcels can cause significant disequilibrium, with regions of powerful uplift having lower para-H2 fractions than the equilib- rium ('sub-equilibrium' conditions) and regions of sinking be- neath the tropopause that move high- fp air downwards ('super- equilibrium' conditions). The rate of equilibration between the two gases is very slow (multiple decades, although the influ- ence of catalysis on aerosol surfaces remains unclear, Massie and Hunten, 1982), such that fast mixing can significantly affect the distribution of para-H2. Changes in the relative populations of J = 0 and J = 1 rotational states therefore cause variations in the shape of the 270-600 cm−1 spectrum, which can be used to trace vertical motions with Jupiter's troposphere. The distribution of Jupiter's para-H2 measured by Voyager has been presented by multiple studies (Conrath and Pirraglia, 1983; Conrath and Gierasch, 1984; Gierasch et al., 1986; Carl- son et al., 1992; Conrath et al., 1998). Carlson et al. (1992) used the similarity between the fp distribution and Jupiter's aerosol distribution to conclude that the spatial variations in fp were related to catalytic equilibration on aerosol particles. Most recently, Conrath et al. (1998) used Voyager 1 spectra to map zonal-mean fp between 200-500 mbar, and showed that al- though the fine belt/zone structure was not well resolved, large- scale contrasts were evident. Specifically, fp had a minimum at the equator and increased towards each pole, with the sug- gestion of an fp maximum over the north pole creating super- equilibrium conditions poleward of 45◦N at the tropopause. It is the existence of this hemispheric asymmetry in Jupiter's para- H2 distribution, and the potential correlation with tropospheric aerosols, that we sought to test with the new SOFIA/FORCAST dataset. Section 2 presents the FORCAST observations (both images and spectra) obtained in 2014. The SOFIA spectra are mod- elled in Section 3, along with a re-analysis of Voyager/IRIS spectra to permit quantitative comparisons. The resulting spa- tial distribution of fp is discussed in Section 4 and compared to the latitudinal variability of tropospheric aerosols, ammonia and phosphine. All latitudes in this paper are planetographic, all longitudes use System III west. 2. Observations 2.1. SOFIA/FORCAST Observations The Faint Object infraRed CAmera for the SOFIA Tele- scope (FORCAST, Adams et al., 2010; Herter et al., 2012) is a thermal-infrared camera designed for the SOFIA airborne ob- servatory (Young et al., 2012; Gehrz et al., 2009). It is com- prised of two cameras (a short-wave camera covering 5-25 µm and a long-wave camera covering 25-40 µm) and grism spec- troscopy. The two arrays of FORCAST (a Si:As blocked impu- rity band (BIB) array for λ < 25 µm and a Si:Sb BIB detector for λ > 25 µm) operate at cryogenic temperatures (4K) and have a plate scale of 0.768 arcseconds per pixel. The 256× 256 array translates to a wide 191 arcsecond field of view that is more than sufficient to capture Jupiter's ∼ 40" disc. When cou- pled with SOFIA's 2.7-m primary mirror (2.5-m effective aper- ture), this permits diffraction-limited observing for λ > 15 µm, with an additional 1.3" blurring due to the jitter of the telescope during flight. The resulting angular resolution ranges from 2- 4", depending on wavelength. The grism spectroscopy uses blazed diffraction gratings, and we specifically used the G227 (17.5-27.3 µm) and G329 (28.7-36.7 µm) grisms. 2.1.1. FORCAST Imaging Images were acquired in eight broad-band filters (Table 1) in order to provide calibration and context for the spectroscopy observations in Section 2.1.2. Imaging in each filter was ob- tained in two epochs on May 2nd 2014: a first group between 03:42-04:03UT and a second group between 06:44-07:11UT (a planned third imaging set could not be acquired). Examples of images acquired in the first group, with the Great Red Spot near the centre of the disc (45◦W longitude), are shown in Fig. 1. For 2 each filter, we compare the raw data (top row) with a contrast- enhanced version to show the spatial structure in the images (bottom row). An empirical function was used to reproduce the emission angle dependence in each image (λ > 8 µm) to cor- rect for the limb darkening, without which very little structure can be seen in the raw images at λ > 30 µm. The true limb darkening of the planet is not resolved, so the apparent func- tion is convolved with the point-spread function of the SOFIA primary mirror. Furthermore, the long-wavelength images have contribution functions that sample the warm lower stratosphere at the highest emission angles, making simple Minneart-type corrections (Minnaert, 1941) impossible. The absolute calibration of the images was determined by en- suring that photometric observations of standard stars, observed both on the flight containing the Jupiter observations and oth- ers in the flight series, matched the predicted flux within the FORCAST filters. The predicted fluxes were based on mod- els for the spectral shape of the stellar emission, tied to ab- solute photometry from ground and space-based observations (Herter et al., 2013). The absolute calibrators used for our im- ages included α Boo, σ Lib, and β UMi. For the 5.4 -- 31.5 µm images, between 6 and 8 standard star observations were com- bined to determine the calibration factors. For the 37.1 µm im- age, 23 standard star observations were combined, to enhance the signal-to-noise at this wavelength. The accuracy of the cal- ibration for each filter is shown in Table 1, with higher accu- racy resulting from wider bandwidths (i.e., more photons in the wider filters), the usage of the filter (more usage means more calibrations are available), the type of calibrator (i.e., the bright- ness of the standard star) and the integration time. For each fil- ter, we summed the Jupiter flux to measure the spectral irradi- ance in Janskys, converting this to a disc-integrated brightness temperature using an effective Jupiter radius of 16.745" (i.e., accounting for the oblateness of the disc). We repeated this for all longitudes observed (4-8 images per filter) to estimate the standard deviation, and show the median spectral irradiance and brightness temperatures in Table 1. These well-calibrated irradiance values will be used to cross-calibrate the FORCAST spectra in Section 2.1.2. Images were obtained in two separate epochs to achieve near- global coverage, allowing us to create cylindrical maps in Fig. 2. These maps are limited to the M-band (5 µm) and N-band (7-13 µm) channels because longer wavelengths showed very little spatial contrast, with the exception of the cold equatorial zone (EZ) and warm neighbouring north and south equatorial belts (the NEB and SEB). The cold Great Red Spot is visible in all images except 7.7 µm, where emission from CH4 gas limits the sensitivity to higher, stratospheric altitudes near 10 mbar. Longitudinal contrasts can be seen in Fig. 2, primarily along the warm equatorial belts. These cloud-free bands ap- pear bright at 5.4 µm due to the negligible attenuation of radi- ance from the deeper troposphere, and also bright in the N-band near 11.1 µm due to warmer 500-mbar temperatures in the belts compared to the zones. There is a strong correlation between longitudinal wave patterns observed on the NEB and SEB in the M and N-band images, confirming that the dynamics mod- ulating the tropospheric cloud opacity also modulate the tem- perature structure and ammonia humidity throughout the upper troposphere (e.g., Fletcher et al., 2016a). At mid-latitudes we observe patches of brighter emission (particularly near 30◦N, 45◦W) that are at the limit of SOFIA's spatial resolution. These warm patches are potentially associated with small cyclonic ovals and brown barges that were visible in Jupiter's northern hemisphere in 2014. Another notable aspect of the SOFIA maps in Fig. 2 is that we detect the warm polar hotspot near 180◦W that is associ- ated with upper stratospheric heating beneath the main auroral oval (e.g., Caldwell et al., 1980; Kostiuk et al., 1993; Liven- good et al., 1993). This bright emission shows up at 7.7 µm due to stratospheric methane, and near 11.1 µm due to stratospheric ethane. The 7.7-µm map shows a banded structure in the strato- sphere consisting of a cool region equatorward of ±15◦, as well as bands of bright emission between ±15 − 30◦ in both hemispheres. These stratospheric bands have been observed by Cassini (Flasar et al., 2004; Simon-Miller et al., 2006) and ground-based observers (Orton et al., 1994) and vary in their temperature over time. Longitudinal contrasts in these warm stratospheric bands are associated with horizontal thermal wave activity at mid-latitudes, and such activity was prominent in 2014 (Fletcher et al., 2016a). Finally, at the limit of the spa- tial resolution of SOFIA we observe a narrow, warm equatorial band at 7.7 µm, consistent with the local maximum in the ther- mal field associated with the current phase of Jupiter's Quasi- Quadrennial Oscillation (Leovy et al., 1991; Orton et al., 1991). However, each of these M- and N-band spatial contrasts are better studied with larger telescopes on the ground. Attempts to retrieve the thermal structure from these cylindrical maps (following the imaging retrieval techniques of Fletcher et al., 2009b) met with limited success given that latitudinal contrasts in the uncorrected images in Fig. 1 are so subtle. Indeed, the main strength of the SOFIA/FORCAST dataset is not in the contextual images, but in the 17-37 µm spectroscopy offered by the grisms, as discussed in the following sections. 2.1.2. FORCAST Spectroscopy Grism spectra were obtained with the 2.4 × 191 arcsecond slit (i.e., ∼ 5 times longer than Jupiter itself) and a spatial pixel scale of 0.768". Only the long-wave channel grisms were used in this study: G227 (17-27 µm) and G329 (28-37 µm) grisms to span 17-37 µm (Keller et al., 2010). Data were obtained on May 3rd 2014, corresponding to a planetocentric solar longi- tude (Ls) of 158◦, approaching the northern autumn equinox in February 2015. At the time of observation, the slit (which has a fixed orientation in the instrument) was oriented approximately North-South on Jupiter. For a single scan, twelve spectra were taken with each grism, starting at a position −20.5 pixels, i.e., −15.74" from the centre of Jupiter, stepping across the planet from west to east in increments of 4 pixels, or 3.07"; the last scan was taken at a position +18.05". The planet's diameter was 35.05" at the time. Details describing the spectra are pro- vided in Table 2. The spectra were taken using the standard IR observing two-position chop-with-nod, with a frequency of several Hz (Table 2). The chop amplitude was ±30" and nod slews were 60". 3 Table 1: Eight imaging filters used in this study, along with the disc-averaged brightness temperature and spectral irradiance measured in each filter. Uncertainties on the TB and total flux are the precision estimated from the standard deviation measured over multiple filters. The large uncertainty at 5.4 µm is related to the large longitudinal variability observed at this wavelength. The calibration uncertainty in the final column is the accuracy of determination of the fluxes of standard stars, based on repeatability of observations in a filter, after correction for atmospheric absorption, over multiple flights. (K) Filter Wavenumber Wavelength Bandwidth Disc-Average TB name 198.0 ± 29.1 5.4 µm 148.4 ± 3.1 7.7 µm 129.1 ± 4.9 11.1 µm 122.9 ± 1.8 19.7 µm 122.8 ± 3.7 31.5 µm 130.4 ± 2.9 33.5 µm 130.5 ± 4.9 34.8 µm 136.2 ± 4.7 37.1 µm (cm−1) 1867.06 1297.86 901.795 507.305 317.894 298.507 287.299 269.222 (µm) 5.356 7.705 11.089 19.712 31.457 33.500 34.807 37.144 (µm) 0.159 0.465 0.954 5.506 5.655 5.658 3.759 3.284 Total Flux (kJy) 6.9 ± 1.0 6.2 ± 0.1 26.2 ± 1.0 284.4 ± 4.2 654.0 ± 20.0 843.9 ± 18.5 857.8 ± 32.2 991.4 ± 34.2 Calibration Unc. (%) 5.1 4.2 9.1 2.3 7.9 12.7 5.0 9.9 Figure 1: Contextual imaging obtained in FORCAST filters on May 2nd, 2014. The Great Red Spot is in the centre of the disc. Each filter is shown in the raw form and with an empirical limb-darkening correction applied to enhance the contrast, particularly at wavelengths beyond 30 µm. This correction leads to some distortion in the oblate spheroidal shape of Jupiter at the edges. The central wavelengths and observation times are shown. Artefacts due to pixel defects and readout striping affect the longest-wavelength images. 4 SOFIA/FORCAST -- Jupiter 2014-05-025.4 µm03:53 UTRaw:Contrast-enhanced:Raw:Contrast-enhanced:7.7 µm04:03 UT11.1 µm03:56 UT19.7 µm03:51 UT31.5 µm03:55 UT33.5 µm03:47 UT34.8 µm03:58 UT37.1 µm04:00 UT Figure 2: Cylindrical maps of multiple observations at 5.4, 7.7 and 11.1 µm. 5.4 µm samples radiance attenuated by Jupiter's 1-4 bar cloud opacity. 7.7 µm samples stratospheric methane emission near 10 mbar. 11.1 µm senses a combination of 500-mbar temperatures and NH3 gas, although emission lines from stratospheric ethane dominate at higher emission angles. The top left panel is an attempt to combine these three wavelengths into a crude 3-colour image to reveal the differences between the three filters. 5 Composite: R(5.4 µm) G(7.7 µm) B(11.1 µm)36031527022518013590450Longitude (degrees, SysIII)-90-60-300306090Latitude (degrees, planetographic) 5.4 µm36031527022518013590450Longitude (degrees, SysIII)-90-60-300306090Latitude (degrees, planetographic) 7.7 µm36031527022518013590450Longitude (degrees, SysIII)-90-60-300306090Latitude (degrees, planetographic)11.1 µm36031527022518013590450Longitude (degrees, SysIII)-90-60-300306090Latitude (degrees, planetographic) The spectra, after pipeline processing through the step of spa- tial and spectral rectification (Rectified IMage, or RIM), were divided by the instrument response and atmospheric transmis- sion curves; the latter was also used to determine the precise wavelength scale. Individual spectra at each 3.07" step along the scan were placed into a 3D cube, and then the cube was interpolated in the scan direction to generate square pixels, 0.768 × 0.768". A polynomial fit to the background (on each row) was subtracted to remove residual flat-field variations and stray light. These final spectral data cubes were calibrated by summing the flux spatially over the disk of Jupiter and convolving spec- trally over the bandpass of five of the broad-band filters (cov- ering 17-37 µm), and scaling to match the flux of the broad- band images at the overlapping wavelengths. Figure 3 com- pares the calibrated, disc-integrated spectrum to the fluxes in the five filters, indicating the increasing uncertainty with wave- length. There was one spectral scan in the G227 grating, and three spectral scans in the G329 grating. The measured flux in the spectra and images deviate sub- stantially for λ > 33 µm (Fig. 3) due to known issues with the FORCAST G329 grism at the longest wavelengths. The cause of this discrepancy remains unresolved, and we omit this range from our subsequent analysis. Indeed, the fluxes mea- sured in the context images are consistent with spectral models that suggest the radiance should increase for λ > 28 µm towards the peak of Jupiter's black body emission beyond 50 µm. The cubes were geometrically registered to determine the latitudes, longitudes and emissions angles corresponding to each pixel. Zonal-mean spectra were generated by averaging pixels at each latitude within 20◦ of the central meridian, using a 5◦-wide lat- itudinal bin on a 2.5◦ latitudinal grid between 60◦N and 60◦S. These raw spectral data are shown in Fig. 4a, where we plot the radiance as a function of latitude and wavenumber and show how the emission angle for each spectrum varies from equa- tor to pole. These FORCAST spectra at specific latitudes are shown and modelled in Section 3. 2.2. Voyager/IRIS Observations In order to perform a quantitative comparison of the temper- atures and para-H2 distributions from the FORCAST observa- tions to previous data, we reanalysed spectra acquired by Voy- ager during the 1979 flybys. The closest approaches (March 5, 1979 for Voyager 1 and July 9, 1979 for Voyager 2) corre- spond to planetocentric solar longitudes of Ls = 170.3◦ and Ls = 180.4◦, respectively, near the northern autumn equinox. The Voyagers carried the Infrared Interferometer, Spectrometer and Radiometer (IRIS) instruments, with a Michelson interfer- ometer providing 180-2500 cm−1 spectra of the giant planets at a spectral resolution of 4.3 cm−1 (Hanel et al., 1979). IRIS spectra were extracted from the expanded volumes available on NASA's Planetary Data System1. Two temporal ranges were considered for each spacecraft, a dedicated north-south map- ping sequence taken during the approach to Jupiter and a full 1http://pds-rings.seti.org/voyager/iris/expanded volumes.html 6 Figure 3: Comparison of the disc-integrated FORCAST spectra from the G227 (17-27 µm) and G329 (28-37 µm) grisms (black dots with grey error bars) with disc-integrated brightness temperatures derived from the images in Table 1. For each filter, we show the bandwidth (horizontal error bar), the standard deviation of the measurement from multiple images (red vertical error bar) and the abso- lute calibration uncertainty (blue vertical error bar). The spectrum has been scaled to match the flux in the contextual images, and a 10% radiometric uncer- tainty (converted to brightness temperature) is shown, representing the 2-12% uncertainty range on the image calibration in Table 1. The downturn in the spectrum longward of 35 µm is a known issue with the G329 grism. average of all spectra acquired within 4 million km of clos- est approach (around ±72 hours from closest approach). The Voyager-1 inbound map was acquired from a distance of ap- proximately 3 million km between 19:00 UT on March 2nd and 10:00UT on March 3rd 1979 (see Fig. 1 of Gierasch et al., 1986). This is the sequence that was used by Conrath and Pir- raglia (1983); Conrath et al. (1998) in their studies of Jupiter's para-H2 and Nixon et al. (2010) in their study of Jupiter's hy- drocarbons. The Voyager-2 inbound map was acquired from a similar distance between 12:00UT on July 6th and 06:00UT on July 7th 1979. At closest approach, the 0.25◦ (4.4-mrad) di- ameter circular IRIS field of view provided spatial resolutions of 1230 km (1.0◦ latitude at the equator) and 2900 km (2.3◦ latitude at the equator) for Voyager 1 and Voyager 2, respec- tively. In the larger average, these high-resolution observations are blended with spectra with resolutions as low as 17600 km (14◦ latitude at the equator) from 4 million km away. In all four cases we had to manually remove corrupted spectra that hadn't been caught by the 'REJECT' flag in the database. All datasets were binned on a 5◦ latitude grid (with a 2.5◦ step), ensuring that the IRIS field of view was entirely on Jupiter's disc. Only emission angles within ±10◦ of the mean for each bin were retained. Uncertainties were derived from the Voyager noise-equivalent spectral radiance (NESR), as described by Sinclair et al. (2014). We note that a misalign- ment of the Voyager-2 interferometer (Hanel et al., 1982) led to slightly less sensitivity in the IRIS measurements in July 1979, and noisier spectra. 20253035Wavelength [µm]110115120125130135140Brightness Temperature [K]270.0300.0350.0400.0500.0600.0Wavenumber [cm-1] Table 2: Details on the grism spectroscopy. Grism wavelength range R = λ/(∆λ) UT time range Longitude range Latitude 1.67◦ G227 1.67◦ G329 17.50-27.30 µm 28.65-36.66 µm 265-270◦ 259-264◦ 140 220 05:36-05:45 05:26-05:35 Integration time Chop freq. 3.97 Hz 2.49 Hz 5.66 s/scan 6.19 s/scan Figure 4: Comparison of the SOFIA/FORCAST and Voyager/IRIS brightness temperature spectra as a function of latitude and wavenumber. The emission angles corresponding to each panel are shown in the right-hand column. The colour-scale for the brightness temperatures is shown in the right-centre panel. The white vertical band in the FORCAST spectrum in panel (a) is the gap between the two FORCAST grisms. 7 (a) SOFIA/FORCAST300400500600Wavenumber [cm-1]-60-40-200204060Planetographic Latitude (a) SOFIA/FORCAST0204060Emission Angle-60-40-200204060Planetographic Latitude(b) Voyager 1300400500600Wavenumber [cm-1]-60-40-200204060Planetographic Latitude (b) Voyager 10204060Emission Angle-60-40-200204060Planetographic Latitude(c) Voyager 2300400500600Wavenumber [cm-1]-60-40-200204060Planetographic Latitude (c) Voyager 20204060Emission Angle-60-40-200204060Planetographic Latitude105110115120125130 The 17-33 µm (300-600 cm−1) spectra from the two flybys are shown in Fig. 4b-c for comparison with the FORCAST data. We are limited to this wavelength range by the poor quality and increasing noise in the long-wavelength FORCAST grism in the 270-300 cm−1 region (not shown). The broader wavenum- ber range for the IRIS data (300-1350 cm−1) will be considered in Section 4 to compare the retrieved para-H2 distribution to zonal-mean distributions of ammonia, phosphine and clouds, following (Gierasch et al., 1986). Fig. 4 uses only the north- south mapping sequences, as these provided a smoother vari- ation of emission angle with latitude. The warmest brightness temperatures are seen at the lowest wavenumbers on the left in Fig. 4, but also near 470 cm−1 in between the broad S(0) and S(1) features. The three spectral panels show subtle asymme- tries between the northern and southern hemispheres which ap- pear to have changed with time - both FORCAST and Voyager- 1 show the lowest brightness temperatures at high northern lat- itudes, whereas Voyager 2 shows the lowest brightness temper- atures in the south, despite similar emission angles in the two Voyager sequences. These asymmetries will be important when deriving the distribution of para-H2 in Section 3. Finally, Fig. 5 shows a comparison between the FORCAST and IRIS brightness temperatures averaged in four spectral ranges coinciding with the lowest wavenumber portion of the FORCAST data, the S(0) and S(1) absorption centres and the peak emission in between. The observations from the two Voy- ager encounters are consistent with one another, with the ex- ception of cooler mid-northern latitudes during the Voyager 1 encounter. The FORCAST data show similar spatial structure at low latitudes and an equator-to-pole drop due to limb dark- ening that will be accounted for in our spectral modelling. The FORCAST data appear to be warmer than the IRIS data at all wavelengths, and we found that a 0.9× scaling factor applied to the FORCAST spectra would bring the two into agreement. This is within the 2-12% calibration accuracy range in Table 1. We discuss the implications for this uncertainty in the radiomet- ric calibration of FORCAST in Section 3. 3. Spectral Modelling SOFIA and Voyager spectra were analysed using the NEME- SIS forward model and spectral inversion code (Irwin et al., 2008), which has been previously used to study the distribu- tion of para-H2 on Saturn (Fletcher et al., 2016b), Uranus (Or- ton et al., 2015) and Neptune (Fletcher et al., 2014). NEME- SIS uses an optimal estimation retrieval architecture (Rodgers, 2000) to maximise the quality of the fit to the data whilst re- maining within the limits of physical plausibility using a priori information. The 17-33 µm spectrum is particularly simple, re- lying only on the collision-induced opacity of H2-H2 and H2- He. Jacobians in Fig. 6(functional derivatives of radiance with respect to the parameter of interest) demonstrate the vertical sensitivity of far-IR and mid-IR jovian spectra to tropospheric temperatures, para-H2, ammonia and phosphine. Spectra in this range are most sensitive to the temperature structure in the 100- 700 mbar region, depending on the potential contribution of Figure 5: Comparison of zonal-mean brightness temperatures between SOFIA/FORCAST (black), Voyager 1 (red) and Voyager 2 (blue), averaged over the four spectral ranges shown. The error bars show the standard devia- tion of the mean across the 20-cm−1 wide bands, rather than the uncertainty of the coadded IRIS spectra, which is smaller and relatively uniform in this wave- length range. A downward scaling of the SOFIA radiances by a factor of ∼ 0.9 makes them consistent with the Voyager radiances (see main text). 8 (a) 300-320 cm-16040200-20-40-60Planetographic Latitude110115120125130135140Brightness Temperature [K](b) 340-360 cm-16040200-20-40-60Planetographic Latitude110115120125130135140Brightness Temperature [K](c) 450-470 cm-16040200-20-40-60Planetographic Latitude110115120125130135140Brightness Temperature [K](d) 540-560 cm-16040200-20-40-60Planetographic Latitude110115120125130135140Brightness Temperature [K] aerosol opacity (Conrath et al., 1998). There exists some ad- ditional sensitivity to the lower stratosphere (50-100 mbar) in the peaks of the S(0) and S(1) lines. However, retrievals of temperature and para-H2 become highly correlated at the low pressures near Jupiter's tropopause, and are likely to be unre- liable in regions of aerosol opacity (p > 500 mbar), so these variables can only be uniquely retrieved in the 200-500 mbar range, approximately (Conrath et al., 1998). Our model atmosphere includes CH4, NH3, PH3, C2H2 and C2H6, as described in Fletcher et al. (2009a). NH3 and PH3 in- fluence both mid-IR spectra (Fig. 6c-d) and the far-IR spectra measured by Voyager (Fig. 6a), so are varied in Section 4 to compare their latitudinal distributions to that of para-H2. The prior T(p) is based on a low-latitude average of Cassini infrared observations (Fletcher et al., 2009a), which themselves used the T(p) from the Galileo Atmospheric Structure Instrument (ASI, Seiff et al., 1998) as the prior. The deep helium mole fraction was set to 0.136 based on in situ Galileo probe measurements (Niemann et al., 1998), and we caution the reader that the ab- solute temperatures and para-H2 fractions are sensitive to this assumption (although He is well-mixed, so this does not affect the measurement of relative spatial variability). The prior for the para-H2 fraction is based on equilibrium at the tempera- tures in our T(p) profile, reaching a maximum of fp = 0.34 at the 113-K tropopause. We retrieve continuous profiles of T(p) and fp(p) simultaneously, whilst holding the abundances of all other gases constant at their priors. Fig. 7 compares the FORCAST and IRIS data to our best- fitting spectral models at three latitudes. The sensitivity of the spectra to changes in para-H2 is shown by offsetting the fp dis- tribution by ±0.05 at all altitudes, demonstrating that the shape of the 300-500 cm−1 (centred on the S(0) line) is sufficiently al- tered, even at the highest latitudes and emission angles, to allow us to derive an fp distribution from these data. Furthermore, as the para-H2 change affects a broad swathe of the continuum, the high spectral sampling of multiple independent measurements allows us to derive a best-fit fp with uncertainties of ≈ 0.02. This figure also confirms the difference between the spectra at 55◦N and 55◦S, which has implications for the para-H2 asym- metry discussed below. The following subsections describe the challenges associated with analysing these data. Figure 6: Jacobians (functional derivatives of radiance with respect to the pa- rameter of interest) for tropospheric sensitivity in the far-IR (panels a,b) and mid-IR (panels c,d). The Jacobians have been normalised to their maximum in this wavelength range, and contours are plotted in units of 0.1, as shown by the key in panel c. In panel a, an increase in temperature causes an increase in the radiance (positive contributions). In panels b, c, and d, an increase in the abundances of para-H2, ammonia and phosphine cause a decrease in ra- diance (negative contributions). In panel b (para-H2) we also show the zero line (grey line), as some weak positive contribution exists longward of 520 cm−1. Vertical dotted lines in panels a and b show the wavelength range of our SOFIA/FORCAST data. The mid-IR influence of ammonia and phosphine will be discussed in Section 4. 9 3.1. Collision-induced absorption model Two different calculations of the collision-induced H2-H2 opacity are available for our forward model - Borysow et al. (1985) provided semi-empirical functions based on ab initio studies of H2-H2 collisions over a broad spectral range; Orton et al. (2007) then revised these calculations using updated quan- tum line shape calculations. Both used ab initio dipole data from Meyer et al. (1989b,a) with a stated accuracy of 3%, and neither incorporated the presence of (H2)2 dimers near to the S(0) and S(1) lines. However, Orton et al. (2007) found that some of the dipole components of Borysow et al. (1985) had been overestimated due to an accounting error in the original work. The compilation of Orton et al. (2007) now forms the 'Alternative' CIA database maintained by HITRAN (Richard et al., 2012). While the authors stated that differences were (a) Temperature200300400500600Wavenumber (cm-1)1.000.100.01Pressure (bar)(b) Para-H2200300400500600Wavenumber (cm-1)1.000.100.01Pressure (bar)0.000.00(c) Ammonia70080090010001100Wavenumber (cm-1)1.000.100.01Pressure (bar)(d) Phosphine900100011001200Wavenumber (cm-1)1.000.100.01Pressure (bar)0.00.20.40.60.81.0 Figure 7: Best fits to the SOFIA/FORCAST (top row) and Voyager-1/IRIS (bottom row) 300-570 cm−1 spectra at three locations: 55◦S, the equator and 55◦N. Spectra were coadded within ±2.5◦ of these central latitudes, and the small uncertainties associated with the IRIS spectra cannot be seen at this scale. Brightness temperature scales are the same for all the figures to show differences, although the primary effect is limb darkening from the equator to the high latitudes. The red lines show the spectral model produced with our best-fitting T(p) and fp(p). The blue/green lines show the sensitivity of the spectra to uniformly increasing/decreasing the para-H2 fraction by 0.05, the upper limit of the latitudinal variability of the para-H2 fraction detected on Jupiter. 10 350400450500550Wavenumber [cm-1]110115120125130135Brightness Temperature [K]350400450500550Wavenumber [cm-1]110115120125130135Brightness Temperature [K]350400450500550Wavenumber [cm-1]110115120125130135Brightness Temperature [K]350400450500550Wavenumber [cm-1]110115120125130135Brightness Temperature [K]350400450500550Wavenumber [cm-1]110115120125130135Brightness Temperature [K]350400450500550Wavenumber [cm-1]110115120125130135Brightness Temperature [K](a) 55oS(c) 55oN(b) EquatorSOFIA/FORCASTSOFIA/FORCASTSOFIA/FORCASTVoyager-1/IRISVoyager-1/IRISVoyager-1/IRIS found for wavenumbers beyond 600 cm−1 and temperatures be- low 120 K, we find that the two collision-induced-absorption (CIA) resources predict different opacities in the S(0) and S(1) lines at temperatures relevant to Jupiter's atmosphere. Fig. 8 compares the calculations of Borysow et al. (1985) and Orton et al. (2007) at 110 K (jovian tropopause temperatures) and two different para-H2 fractions. At equilibrium, the Borysow et al. (1985) model predicts around 7% higher absorption in the S(0) and S(1) peaks than the model presented by Orton et al. (2007). This has implications for our fits to the 17-33 µm spectra. Indeed, when we run our inversion of Voyager-1/IRIS spectra using the two different opacity databases we derive wholly dif- ferent zonal mean temperature cross-sections (see Fig. 9). Fits with the CIA opacity from Borysow et al. (1985) are qualita- tively similar to those presented in previous studies (e.g., Con- rath et al., 1998), with a broad tropopause region between 100- 200 mbar. Conversely, fits with the CIA opacity from Orton et al. (2007) require a much narrower tropopause region near 200 mbar and warmer temperatures in the lower stratosphere. Given that the latter CIA database has a smaller contrast be- tween the absorption maxima at 354 and 587 cm−1 and the minimum near 460 cm−1, our inversion requires warmer lower- stratospheric temperatures to reproduce the brightness temper- atures near the peaks of the S(0) and S(1) features. This is counter to the findings of all previous inversions of IRIS data, but cannot be discounted given that the newer calculations of Orton et al. (2007) are considered to be an improvement on those of Borysow et al. (1985). Furthermore, the newer calcu- lations showed better agreement with the experimental results of Birnbaum et al. (1996). Nevertheless, the newer calculations generally produce poorer fits to the data (worsening the χ2/N by a factor of two at all latitudes). We suggest that additional (H2)2 dimer absorption (a molecule bound by van der Waals forces, Frommhold et al., 1984; Schaefer and McKellar, 1990; Birnbaum et al., 1996) is required to resolve this discrepancy. In their study of Uranus' H2-He spectrum, Orton et al. (2014) incorporated new calcu- lations of the dimer opacity (provided by J. Schaefer) to their existing CIA database (Orton et al., 2007). A ∼ 5% enhance- ment in the absorption near the S(1) peak is shown as the red curve in Fig. 8, bringing the total absorption closer to the orig- inal erroneous values of Borysow et al. (1985). Furthermore, Frommhold et al. (1984) discussed dimer features near the S(0) line in Voyager/IRIS spectra of Jupiter, and found that bound- free transitions between dimers contributed approximately 5% to the opacity at 120 K. However, the dependence of the dimer contribution on the para-H2 fraction has not been modelled, and inclusion of these new calculations produced features in our CIA database that are not directly observed, implying that there is room for improvement of the theoretical simulations. We therefore use the CIA opacity from Borysow et al. (1985) in this paper for simplicity and comparison to previous works, and caution the reader about the 'missing absorption' in the HI- TRAN compilation (Orton et al., 2007; Richard et al., 2012) near the peaks of the S(0) and S(1) lines. Figure 8: H2-H2 absorption coefficient, normalised by the square of the gas density, at 110 K for 'normal' hydrogen (para-H2 fractions of 0.25, the high- temperature asymptote) and equilibrium H2 at 110 K, representative of the cold- est temperatures near Jupiter's tropopause). The black line is for the model of Borysow et al. (1985), the blue line is for Orton et al. (2007) (identical to that included in HITRAN 2012, Richard et al., 2012). The red line shows the in- creased opacity from adding (H2)2 dimer absorption to the CIA from Orton et al. (2007). 11 T= 110.0K, Normal para-H2200300400500600700Wavenumbers [cm-1]02×10-64×10-66×10-68×10-61×10-5Absorption [cm-1/amagat2]Orton + DimersOrton et al. (2007)Borysow et al. (1985)T= 110.0K, Equilibrium para-H2200300400500600700Wavenumbers [cm-1]02×10-64×10-66×10-68×10-6Absorption [cm-1/amagat2]Orton + DimersOrton et al. (2007)Borysow et al. (1985) Figure 9: Comparison of zonal mean temperature cross-sections derived from the Voyager-1 north-south mapping sequence assuming CIA opacities from (a) Borysow et al. (1985) and (b) Orton et al. (2007). Temperature precision ranges from 2 K at 700 mbar to 1.2 K at 400 mbar and 1.8 K at the tropopause, but this does not encompass the systematic differences between the two sources of CIA. 3.2. Aerosol influence and spectral windowing The reliable portion of the SOFIA/FORCAST data covers the 300-570 cm−1 (17.5-33.3 µm) spectral range. In their study of Voyager/IRIS spectra, Conrath et al. (1998) noted that spectra below 320 cm−1 and between 430-520 cm−1 needed to be ex- cluded due to the possible presence of aerosol opacity (Carlson et al., 1992). Carlson et al. (1992) showed that the inclusion of opacity from large aerosols (with radii comparable to the wavelength) would shift the Jacobian in the absorption mini- mum (near 460 cm−1 in Fig. 6a, 21.7 µm) upwards to cooler altitudes, thereby reducing the overall contrast in the 300-570 cm−1 spectrum. Subsequent retrieval studies have utilised com- pact aerosol layers with a base near 800 mbar (consistent with the expected condensation altitude of a 3× solar enriched NH3 gas, Atreya et al., 1999) to provide thermal-infrared opacity (Wong et al., 2004; Matcheva et al., 2005; Achterberg et al., 2006; Fletcher et al., 2009a). Using the Voyager-1 north-south mapping sequence, we re-ran the zonal mean temperature and para-H2 inversion using three different assumptions: (1) zero aerosol opacity; (2) a globally-homogeneous cloud layer with fixed optical depth of unity at 1 bar; and (3) allowing the aerosol opacity to vary as a free parameter during fits. The cloud was included as a compact cloud (aerosol-to-gas scale height ratio of 0.2) with a base at 800 mbar. Each experiment was repeated twice, once with the full 300-570 cm−1 range and once restrict- ing to the 320-430 cm−1 and 520-570 cm−1 ranges following Conrath et al. (1998). T(p), fp and para-H2 disequilibrium retrievals with and with- out aerosols are compared in Fig. 10. When fitting the full spectral range, we found closer fits to the data when no aerosol opacity was included (Fig. 10 top row), and negligible differ- ence between the fits when either homogeneous or inhomoge- neous cloud layers were used (Fig. 10 bottom row). In this instance, this argues against the addition of aerosol opacity of any kind to the 300-570 cm−1 fits. Furthermore, when aerosols were allowed to vary we obtained only a geometrical effect (a correlation with the emission angle), with lower optical depths required at the equator (∼ 0.3) and higher optical depths near the poles (> 2). This suggests that our assumed compact verti- cal distribution of aerosol opacity is incorrect. Furthermore, the goodness-of-fit was substantially improved at the equator when no aerosols were included, precisely in the region that would be expected to show the highest opacity. Unfortunately, neither the low-resolution FORCAST data nor the IRIS data provide the necessary centre-to-limb observations to further constrain this vertical distribution. The presence or absence of aerosols therefore adds uncertainty to our inversions - the addition of aerosols permits a warmer atmosphere at ∼ 700 mbar by 1-2 K and a slightly cooler atmosphere near 300 mbar by < 1 K, and negligible differences for p < 300 mbar. The addition of aerosols causes systematic changes to fp of around 0.01, and these are most apparent for p > 500 mbar where fp can no longer be constrained. Finally, we find that omission of the 300-320 cm−1 and 430- 520 cm−1 regions has negligible impact on our retrieved T(p) and f (p) distributions. Changes to the fp values are smaller than 0.01 and changes to the tropospheric temperatures are smaller than 1 K, both within the formal uncertainties on the T(p) and fp(p) structures shown in Fig. 10. Crucially, the lati- tudinal contrasts in temperature and fp are the same irrespective of how we include aerosols and spectral windowing in the in- version, and these contrasts match those of Conrath et al. (1998) extremely well. We will return to the importance of aerosols in Section 4. 12 6050403020100-10-20-30-40-50Planetographic Latitude0.500.300.200.100.05Pressure [bar](a) Borysow et al. (1985) [K]10811211211211211611611611612012012012012412412412412812812812813213213213613614014014450403020100-10-20-30-40-50-60Planetographic Latitude(b) Orton et al. (2007) [K]108108108112112112112116116116116120120120120124124124124128128128128132132132132136136136136140140144 Figure 10: Comparison of the temperature, para-H2 and para-H2 disequilibrium retrievals without (top row) and with (bottom row) the inclusion of aerosol opacity. Results are shown for fits of the full 300-570 cm−1 range of the Voyager-1/IRIS north-south mapping sequence. Temperature uncertainties range from 2 K at 700 mbar to 1.2 K at 400 mbar and 1.8 K at the tropopause; fp uncertainties range from 0.02 at 600 mbar and 100 mbar to 0.01 at 200 mbar. 3.3. Spectral selection from Voyager The tests conducted so far used only the north-south mapping sequence from Voyager 1. However, Section 2.2 describes the coaddition of spectra to create 'global averages' of each Voy- ager dataset within 4 million km of Jupiter and one north-south mapping sequence for each spacecraft. These four different datasets are compared to the zonal mean T(p), fp(p) and para- H2 disequilibrium from SOFIA/FORCAST in Fig. 11. Similar latitudinal structures are evident and will be discussed in Sec- tion 4. However, we note that there are differences depending on (i) whether we just take the north-south maps or include the full Voyager dataset; and (ii) whether we consider Voyager 1 or 2, only 4 months apart. In the first case, differences are due to the blending of a range of spatial resolutions in the global aver- ages and the higher signal-to-noise of the coadds (more spectra will reduce the noise and weight the inversion more strongly towards the data). In the second case, the Voyager-2 spectra were generally noisier due to the misalignment of the Voyager- 2 interferometer (Hanel et al., 1982) discussed in Section 2.2, which results in the retrieval being more strongly weighted to the prior. However, two features of the para-H2 distribution are robust no matter how we process the dataset - there is a mini- mum in fp (and hence sub-equilibrium conditions) at the equa- tor; and an asymmetry in fp between the northern and southern high latitudes which will be discussed in Section 4. Close inspection of the fits to the Voyager 1 and Voyager 2 spectral averages revealed that the residual always took on a common undulating shape, indicating that we were underfit- ting one side of the ∼ 470-cm−1 peak and overfitting the other. This undulation in the residual was worse for Voyager 2 than for Voyager 1, potentially as a result of the interferometer mis- alignment. To investigate the possibility of a wavelength offset in the IRIS spectra, we reran the zonal-mean retrievals shifting the spectrum in 0.25-cm−1 steps between 0 and 4.0 cm−1 (i.e., one FWHM for IRIS) using the CIA of Borysow et al. (1985). Better fits to the data could be obtained with shifts of 0.4-1.1 cm−1 for Voyager 1 and 1.6-2.5 cm−1 for Voyager 2. However, such large shifts are not viable, given that NH3 rotational fea- tures in the 200-250 cm−1 of the IRIS spectra match our models extremely well. Furthermore, we might expect an interferom- eter misalignment to stretch the wavenumber grid, rather than shifting it. Finally, Fig. 8 reveals a spectral offset between the absorption peaks between the Borysow et al. (1985) and the Or- ton et al. (2007) CIA data, indicating that the required spectral shift would be model dependent. The undulation in the residual could therefore be due to uncertainties in the collision-induced absorption data that have yet to be resolved. 3.4. Radiometric calibration of FORCAST Fig. 11a demonstrates that the temperatures derived from SOFIA/FORCAST spectra were systematically warmer, by ∼ 2 K at the tropopause and 8 − 10 K at 700 mbar, than those ob- tained by Voyager/IRIS 35 years earlier. Rather than invoking global changes to Jupiter's temperatures, we assume that this is caused by systematic offsets in the radiometric calibration. A 10% reduction of the SOFIA/FORCAST flux makes SOFIA and Voyager approximately consistent with one another (Fig. 12), so we investigate the impact of such a change on the re- trieved quantities, repeating the retrievals with scale factor of 13 0.500.300.200.100.05Pressure [bar](a) Zonal Mean Temperature [K]1121121121161161161161201201201201241241241241281281281281321321321361361401401446050403020100-10-20-30-40-50-60Planetographic Latitude0.500.300.200.100.05Pressure [bar]1081121121121121161161161161201201201201241241241241281281281281321321321361361401401441440.500.300.200.100.05Pressure [bar](b) Para-H2 Fraction0.290.310.310.330.330.330.330.350.370.396050403020100-10-20-30-40-50-60Planetographic Latitude0.500.300.200.100.05Pressure [bar]0.270.290.290.310.310.330.330.330.330.350.370.396050403020100-10-20-30-40-50-60Planetographic Latitude0.500.300.200.100.05Pressure [bar]30.0--0.02-0.02-0.02-0.01-0.01-0.010.000.000.010.010.020.030.040.050.500.300.200.100.05Pressure [bar](c) Para-H2 Disequilibrium30.0--0.02-0.02-0.02-0.01-0.01-0.01-0.010.000.000.000.000.010.010.010.020.030.040.05 Figure 11: Comparison of the retrieved atmospheric structure from four different Voyager spectral averages and SOFIA/FORCAST. The left-hand column shows the T(p) retrieval with 2-K contours, with uncertainties ranging from 2 K at 700 mbar to 1.2 K at 400 mbar and 1.8 K at the tropopause. The centre column shows the fp(p) retrieval with 0.01 spacing, with uncertainties ranging from 0.02 at 600 mbar and 100 mbar to 0.01 at 200 mbar. The right-hand column shows the degree of disequilibrium with 0.01 contours, from sub-equilibrium at low-latitudes to super-equilibrium at high latitudes. 14 6050403020100-10-20-30-40-50-60Planetographic Latitude0.500.300.200.100.05Pressure [bar]1121121121161161161161201201201201241241241241281281281281321321321321361361401401441441486050403020100-10-20-30-40-50-60Planetographic Latitude0.500.300.200.100.05Pressure [bar]1121121121161161161161201201201201241241241241281281281281321321321361361401401441441486050403020100-10-20-30-40-50-60Planetographic Latitude0.500.300.200.100.05Pressure [bar]1121121121161161161161201201201201241241241241281281281281321321321361361401401441446050403020100-10-20-30-40-50-60Planetographic Latitude0.500.300.200.100.05Pressure [bar]1081121121121121161161161161201201201201241241241241281281281281321321321361361401401441446050403020100-10-20-30-40-50-60Planetographic Latitude0.500.300.200.100.05Pressure [bar]0.270.290.290.310.310.330.330.330.330.350.376050403020100-10-20-30-40-50-60Planetographic Latitude0.500.300.200.100.05Pressure [bar]0.290.310.310.330.330.330.330.350.356050403020100-10-20-30-40-50-60Planetographic Latitude0.500.300.200.100.05Pressure [bar]0.290.310.310.330.330.330.330.350.356050403020100-10-20-30-40-50-60Planetographic Latitude0.500.300.200.100.05Pressure [bar]0.270.290.290.310.310.330.330.330.330.350.370.396050403020100-10-20-30-40-50-60Planetographic Latitude0.500.300.200.100.05Pressure [bar]30.0--0.02-0.02-0.02-0.01-0.01-0.010.000.000.010.010.020.030.040.056050403020100-10-20-30-40-50-60Planetographic Latitude0.500.300.200.100.05Pressure [bar]-0.0320.0--0.02-0.01-0.01-0.01-0.010.000.000.010.010.020.030.040.056050403020100-10-20-30-40-50-60Planetographic Latitude0.500.300.200.100.05Pressure [bar]0.0-2-0.01-0.01-0.010.000.000.000.000.010.010.010.020.020.026050403020100-10-20-30-40-50-60Planetographic Latitude0.500.300.200.100.05Pressure [bar]-0.02-0.01-0.01-0.010.000.000.000.000.010.010.010.010.02(b) Voyager 1 North-South Maps(c) Voyager 1 Global Average(d) Voyager 2 North-South Maps(e) Voyager 2 Global Average6050403020100-10-20-30-40-50-60Planetographic Latitude0.500.300.200.100.05Pressure [bar]0.250.270.270.290.290.310.310.310.330.330.330.330.350.370.396050403020100-10-20-30-40-50-60Planetographic Latitude0.500.300.200.100.05Pressure [bar]8011121121161161161161201201201201241241241241281281281321321361361401401441441481481521566050403020100-10-20-30-40-50-60Planetographic Latitude0.500.300.200.100.05Pressure [bar]-0.04-0.03-0.03-0.02-0.0210.0--0.01-0.0100.00.000.010.010.020.030.0450.0(a) SOFIA/FORCAST 2014 0.9, 1.0 and 1.1. These adjustments are within the 2-12% un- certainty range associated with the FORCAST absolute calibra- tion (Table 1). This results in equatorial 110-mbar temperatures ranging from 114-118 K and equatorial 330-mbar temperatures ranging from 120-123 K. These are within the ±2 K uncertainty range on the temperature retrieval. The absolute para-H2 fraction is larger (and closer to equi- librium) for the 0.9× scale factor (Fig. 12b) by 0.02-0.03 com- pared to the 1.0× scale factor, meaning that the downward scal- ing of the SOFIA/FORCAST flux is more consistent with equi- librium conditions (Fig. 12c). The formal retrieval uncertainties on the FORCAST para-H2 inversions are latitude-dependent, but at mid-latitudes they range from a minimum of 0.014 at 300 mbar to 0.022 at 100 and 750 mbar, excluding systematic un- certainties due to aerosols, helium, radiometric offsets, etc. At first glance, the small uncertainties might appear to be at odds with the spectral forward models in Fig. 7, where we showed the spectral effects of a ±0.05 offset in the para-H2 fraction. The resulting models appeared to bracket the 1σ measurement uncertainties. However, the para-H2 fraction affects the full 17- 37 µm range, and the ±0.05 offset curves were systematically above or below the best-fitting models, having a substantial ef- fect on the goodness-of-fit. This means that our formal para-H2 uncertainty is less than 0.05, and closer to the 0.01-0.02 range quoted here. Nevertheless, this implies that the measured dis- equilibrium in Fig. 12c is closer to zero within the uncertain- ties for the 0.9× scale factor, albeit with the same equator-to- pole increase in fp. Absolute offsets of ∼ 4 K remain present near 700 mbar, but these are not considered robust as they are primarily driven by noisy data from the long-wave FORCAST grism in Fig. 7. Furthermore, the absolute values of the temperature and para- H2 fraction are sensitive to the degree of smoothing applied to the vertical profiles during the inversion process. Relax this smoothing, and the temperature and fp excursions from the a priori increase in size until oscillations in the vertical profile become unrealistic and non-physical. The final result is there- fore a compromise between vertical smoothing and the quality of the fit to the data, using a priori uncertainties that are based on previous studies - latitudinal temperature excursions of < 5 K at 250 mbar (based on Cassini observations at higher spatial resolution, e.g., Simon-Miller et al., 2006) and latitudinal para- H2 excursions of < 0.05 at 250 mbar (Conrath et al., 1998). Unfortunately, however, the constraint on the retrieval imposed by the data depends on the signal-to-noise ratio, which is differ- ent for Voyager 1, Voyager 2 and the SOFIA/FORCAST data. The results in Fig. 11 are our best attempt at balancing these competing influences on the retrieval, but a more robust com- parison would require the same instrumentation for each epoch under comparison. 4. Discussion Interpretation of spatial variations in Jupiter's far-infrared spectrum remains a substantial challenge almost four decades after the Voyager/IRIS observations, despite the improvements in Earth-based observational capabilities at other wavelengths. The SOFIA/FORCAST observations represent an attempt to plug this capability gap, but - like Voyager/IRIS - can only re- solve the largest scales of Jupiter's banded structure and the Great Red Spot. Section 3 showed how absolute temperature and para-H2 estimates are adversely affected by uncertainties in radiometric calibration, the chosen source of spectral model for the collision-induced opacity, the degree of smoothing im- posed on the spectral inversion, and (to a lesser extent) the in- fluence of aerosol opacity and spectral windowing. In addition, the helium mole fraction remains poorly constrained and could serve to systematically scale the 17-37 µm brightness temper- atures up and down, which would have the effect of modify- ing the magnitude of the para-H2 disequilibrium (Gautier et al., 1981). The problem is made even more complex by using dif- ferent instruments with different sensitivities, even between the two Voyager flybys. Following Section 3.4, we estimate that temperatures and para-H2 fractions are no more accurate than ±2 K and ±0.02 in the 100-700 mbar range, respectively (using the Galileo-derived He abundance and CIA of Borysow et al., 1985). Nevertheless, Fig. 11 and Fig. 12 do show consistent fea- tures for the three epochs studied (March 1979, June 1979 and May 2014), irrespective of the inversion assumptions. To make this clearer, Fig. 13 extracts zonal-mean cross-sections of tem- perature, para-H2 and the degree of disequilibrium at 110-mbar (near the tropopause, and at the limit of the vertical sensitivity in Fig. 6a) and 330 mbar (near the peak of the Jacobians) from SOFIA and the two Voyager north-south sequences. These all show: an equatorial region (±0 − 15◦) that is warmer than tem- perate mid-latitudes (±20 − 40◦); a temperature drop poleward of ±50◦; sub-equilibrium para-H2 in the equatorial region and an equator-to-pole increase in fp to bring it closer to equilib- rium at the highest latitudes. We discuss each of these findings below. 4.1. Tropospheric Temperatures Jupiter's temperature field has been studied at higher spatial resolutions in the mid-infrared from Cassini and ground-based observatories (e.g., Flasar et al., 2004; Simon-Miller et al., 2006; Fletcher et al., 2011). These revealed equatorial temper- atures within 7◦ of the equator that were cooler than the neigh- bouring northern (7 − 17◦N) and southern (7 − 20◦S) equato- rial belts (NEB, SEB). In the majority of the SOFIA and Voy- ager temperature cross-sections in Fig. 11, these appear to be blended together to form a warm band throughout the ±15◦ re- gion. Fig. 6 of Conrath et al. (1998) and Fig. 2 of Simon-Miller et al. (2006) showed the same effect. The raw brightness tem- peratures in Figs. 4-5, as well as the contextual imaging in Fig. 1, suggests that there are maxima near ±10◦ latitude. However, these belts have such a limited contrast with the cooler equa- torial zone that they have very little influence on the spectral inversions. The cool equator is certainly present in the FOR- CAST and IRIS data, but has a limited visibility on the 2.5◦ latitude grid (with 5◦ bin widths) used to analyse the data. In conclusion, there is no inconsistency with the mid-IR retrievals, but we are left requiring higher spatial resolutions for future far- IR studies of Jupiter's tropics. 15 Figure 12: Effect of uniformly scaling the SOFIA/FORCAST flux downwards by 10% (top row) compared to the original calibration of the data (bottom row). Temperature, para-H2 and the disqequilibrium are plotted with exactly the same contour spacing as Fig. 11. Figure 13: Zonal-mean cross-sections of temperatures, para-H2 fraction and disequilibrium at 110 and 330 mbar. Black lines (with grey error bars) are from SOFIA/FORCAST with the 0.9× scaling, red lines are from the Voyager-1 north-south maps, and the blue lines are from the Voyager-2 north-south maps. 16 6050403020100-10-20-30-40-50-60Planetographic Latitude0.500.300.200.100.05Pressure [bar]8011121121161161161161201201201201241241241241281281281321321361361401401441441481481521560.500.300.200.100.05Pressure [bar](a) Zonal Mean Temperature [K]1041081121121121161161161161201201201201241241241241281281281281321321361361401401441441481520.500.300.200.100.05Pressure [bar](b) Para-H2 Fraction0.290.310.310.330.330.330.330.356050403020100-10-20-30-40-50-60Planetographic Latitude0.500.300.200.100.05Pressure [bar]0.250.270.270.290.290.310.310.310.330.330.330.330.350.370.390.500.300.200.100.05Pressure [bar](c) Para-H2 Disequilibrium10.0--0.01-0.010.000.000.000.000.010.010.026050403020100-10-20-30-40-50-60Planetographic Latitude0.500.300.200.100.05Pressure [bar]-0.04-0.03-0.0320.0--0.0210.0--0.01-0.0100.00.000.010.010.020.030.0450.0(a) Temperature at p=110 mbar100105110115120Temperature [K](d) Temperature at p=330 mbar6050403020100-10-20-30-40-50-60Planetographic Latitude116118120122124126Temperature [K](b) Para-H2 at p=110 mbar0.280.300.320.340.360.38Para-H2 Fraction(c) Disequilibrium at p=110 mbar-0.04-0.020.000.020.04Para-H2 Disequilibrium(e) Para-H2 at p=330 mbar6050403020100-10-20-30-40-50-60Planetographic Latitude0.280.300.320.340.36Para-H2 Fraction(f) Disequilibrium at p=330 mbar6050403020100-10-20-30-40-50-60Planetographic Latitude-0.04-0.020.000.020.04Para-H2 Disequilibrium The equatorial temperatures observed by FORCAST in 2014 (∼ 115 K at 200 mbar at the equator) are closest to those mea- sured by Voyager 1, but the differences are within our 1-2 K uncertainty range. Jupiter's equatorial temperatures are known to vary with time (by 2-4 K at the ∼ 250 mbar level, Orton et al., 1994; Simon-Miller et al., 2006), possibly as a result of sporadic equatorial brightening events that bring fresh updrafts of material to the tropopause. Gierasch et al. (1986) found equatorial differences between Voyager 1 and 2 that are con- firmed by Fig. 11c-d, with the latter being ∼ 2 K cooler at 150 mbar. This was consistent with the equatorial variability in the ground-based record of Orton et al. (1994), which showed the equatorial temperatures cooling by ∼ 3 K between 1978 and 1983. At temperate latitudes we do not resolve the belt/zone struc- ture in either the images (Fig. 1) or the spectra. The result is a broad cold band at 20 − 30◦ in both hemispheres, blending together the multiple belts and zones in Jupiter's temperate do- main. During the Voyager encounters, the southern cold band was colder and wider in latitudinal extent than the northern cold band, with a southerly warm belt near 40−50◦ marking the tran- sition into the polar environment. This asymmetry is reversed in the FORCAST data, with cooler temperatures at northern mid- latitudes than in the south, which can be seen in the raw data (Fig. 4a) as an asymmetry between the northern and southern hemispheres. The changes are at the 2-3 K level, and are con- sistent with the levels of temperature variability shown in the long-term records of Orton et al. (1994) and Simon-Miller et al. (2006). Finally, we note that all of our inversions (IRIS and FORCAST) suggest a transition to cold polar vortices poleward of ±50◦ (100 < p < 300 mbar), consistent with previous stud- ies. In summary, we find that although FORCAST can offer constraints on Jupiter's tropospheric temperatures, the spatial resolution of SOFIA (and the Voyager observations at far-IR wavelengths) blends together much of the finer belt/zone struc- tures required for dynamical studies. 4.2. Para-Hydrogen Despite the low spatial resolution, the spectral coverage of FORCAST and IRIS offer the capability to map Jupiter's para- H2 distribution using the broad S(0) and S(1) lines. This is a challenging measurement given the different sensitivities of the two instruments, but inversions in Figs. 11 and 13b show con- sistent trends with latitude - a sub-equilibrium minimum in fp within ∼ 15◦ of the equator that extends vertically throughout the range of sensitivity (approximately 200-500 mbar), and a transition to equilibrium (and possibly super-equilibrium) con- ditions at high latitudes. The para-H2 distribution appears to be asymmetric between the northern and southern hemispheres - FORCAST and Voyager 1 (both observing before northern autumnal equinox) suggest a more rapid increase of fp with lat- itude in the northern hemisphere than the southern hemisphere, resulting in super-equilibrium conditions poleward of 50◦N. Conversely, Voyager 2 observations (taken at northern autumnal equinox) suggest a stronger increase in fp towards the southern pole. The latitudinal gradients of the raw brightness tempera- tures in Fig. 4 support these findings, with a stronger northern gradient in Voyager 1 observations and a stronger southern gra- dient in Voyager 2 observations. The para-H2 distributions do show finer-scale variations with latitude, with some evidence fp minima in the mid-latitude regions associ- for additional ated with the cool temperate bands described earlier. However, given the retrieval uncertainties of 0.01-0.02 in the ∼ 250-mbar region, these contrasts are not considered to be significant. Qualitatively, our Voyager 1 and FORCAST results are con- sistent with those of Conrath et al. (1998), who also showed an asymmetry in fp and potential super-equilibrium conditions at Jupiter's high northern latitudes (their Fig. 9). The region of high- fp air is therefore a repeatable feature of Jupiter's high lat- itudes, but apparently not a permanent one, with observations by Voyager 2 showing no evidence for this feature in the north. However, we caution the reader that the fits to the Voyager 2 spectra showed a larger undulation in the residual than Voyager 1 (Section 3), suggestive of a discrepancy in the wavelength calibration that could influence the retrieved fp structure. The time evolution of this equator-to-pole gradient in para-H2 will only be truly revealed by observing Jupiter's poles over longer time spans with the same instrument. 4.2.1. Balancing dynamics and chemical equilibration The vertical distribution of para-H2 is approximately gov- erned by a balance between the strength of vertical mixing and the rate of chemical equilibration between the two spin isomers (Gierasch et al., 1986; Conrath et al., 1998). Unfortunately, neither the timescale for dynamical motions (τd) nor the hy- drogen equilibration time (τp) are well known, and both are expected to vary with latitude. Hydrogen conversion can oc- cur in the gas phase due to (i) hydrogen exchange between molecular and atomic hydrogen or (ii) paramagnetic conver- sion between two H2 molecules, but these are deemed less ef- ficient (i.e., longer τp) than (iii) paramagnetic conversion on the surfaces of aerosols (Massie and Hunten, 1982; Conrath and Gierasch, 1984; Fouchet et al., 2003). The efficiency of conversion depends on many factors - the surface area of the aerosols; UV irradiation of the aerosols to generate active sur- face sites; their optical depth; and the mixing of the air within the cloud layers. If the para-H2 distribution solely reflected the efficiency of this paramagnetic conversion and was not influ- enced by vertical mixing, then we should find that the cloudiest regions (i.e., the equator and the northern and southern tropi- cal zones) would have values closest to equilibrium, with more extreme disequilibrium over the cloud-free belts. Carlson et al. (1992), in their analysis of a small number of localised regions in the IRIS dataset, suggested that such a correlation did exist. However, we find that Jupiter's equator is the site of the largest sub-equilibrium conditions, counter to the purely chemical ex- planation. This led Conrath et al. (1998) to provide a quantitative as- sessment of the timescales governing the distribution of para- H2 taking dynamics into account, showing that that the dynam- ical timescale must be smaller (∼ 70 years) than the timescale required for the equilibration between the ortho- and para-H2 (∼ 110 years) in Jupiter's upper troposphere. Parcels of low- fp air could be advected upwards into colder atmospheric lay- 17 ers at a rate faster than the conversion to ortho-H2. However, (Fouchet et al., 2003) scaled their stratospheric para-H2 results to the 200-mbar level and suggested lower limits on τp of 15-22 years assuming paramagnetic conversion on aerosols, or 30-50 years assuming paramagnetic conversion in the gas phase (they did not favour hydrogen exchange as a dominant mechanism in Jupiter's troposphere, as this resulted in extremely long equili- bration timescales). These lower limits raise the possibility that τp could be smaller than τd on Jupiter (as found on the other giant planets by Conrath et al., 1998), suggesting that chemi- cal conversion could indeed dominate over dynamics. Further quantitative progress in separating the dynamical and chemical influences on fp(p) cannot be made without better constraints on the hydrogen conversion mechanism. 4.2.2. High-latitude para-H2 The polar tropopause is colder than elsewhere on the planet (Fig. 11), meaning that the equilibrium para-H2 is largest at the poles. However, we consistently retrieve fp in excess of this cold equilibrium in both the SOFIA and Voyager datasets. This matches the findings of Conrath et al. (1998), who suggested that these high-latitude fp maxima were due to air descend- ing below the tropopause, enriching the 200-500 mbar levels in high- fp air from the colder layers above. Such a circula- tion pattern, with air rising at low latitudes and descending over the poles, is consistent with that modelled by Conrath et al. (1990). However, given the uncertainties on the high-latitude super-equilibrium shown in Fig. 13, we must question the va- lidity of this conclusion. The inversions in Fig. 11 placed the largest fp in the 50- 100 mbar region above the tropopause, which makes little sense as the equilibrium fp should decline with altitude in the lower stratosphere as the temperature rises. These lower-stratospheric values are likely to be untrustworthy, given the poor vertical sensitivity of the far-IR inversions to the tropopause regions (Fig. 6), and the high emission angles that were used to sample the 50 − 60◦ latitude regions. These could lead to systematic errors in the para-H2 retrieval. However, the stratospheric in- vestigation of Fouchet et al. (2003) suggested the possibility that Jupiter's stratospheric fp matches the tropopause value (at least in the ±30◦ latitude range averaged in their ISO observa- tions), i.e., Jupiter's fp remains larger than equilibrium in the stratosphere without any conversion back to ortho-H2. The ver- tical T(p) is poorly known at high latitudes and p < 100 mbar, particularly within the cold polar vortex, so maybe a region of high- fp air could exist at lower pressures. The hypothesis that subsidence generates regions of high- fp air at high latitudes on Jupiter cannot be rejected by these data. However, Fig. 13 shows that the strongest conclusion that can be drawn is that fp is closest to equilibrium at Jupiter's high latitudes, so large-scale subsidence need not be invoked. Instead, if we assume the efficiency of hydrogen conversion to be latitudinally uniform, then the equator-to-pole gradient may simply reflect weaker tropospheric upwelling at high lati- tudes to replenish equilibrated air near the tropopause. Indeed, parameterisations of Jupiter's eddy diffusivity by Wang et al. (2015) demonstrate a strong decrease in the expected strength of vertical mixing from the equator to pole due to the planet's rotation. We investigate correlations with other tracers of tro- pospheric mixing in the next section. 4.3. Comparing para-H2 to tropospheric gases and aerosols To better explore the correlation between para-H2 and Jupiter's other tropospheric variables, we replicate the analy- sis of Gierasch et al. (1986) by extending the Voyager-1/IRIS spectral retrieval to cover the full 300-1350 cm−1 (7.4-33.3 µm) range (all the aforementioned results considered 300-600 cm−1 only). This allows us to study Jupiter's NH3, PH3 and aerosol opacity in the mid-IR (see Jacobians in Fig. 6c-d) at the same epoch and spatial resolution as the temperatures and para-H2 discussed above. Jupiter's T(p) and para-H2 distribution were retrieved simultaneously with a parameterised NH3 distribution (a variable deep abundance for p > 800 mbar, declining at higher altitudes according to a variable fractional scale height), and a scale factor for the PH3 distribution of Fletcher et al. (2009a). In addition, we scaled the optical depth of an 800- mbar compact cloud layer consisting of NH3 ice particles (re- fractive indices based on Martonchik et al., 1984) with a log- normal distribution of radius 10 ± 5 µm. Aerosol cross sec- tions and single scattering albedos were calculated assuming Mie scattering. The sources of spectral line data are identical to those used for Cassini/CIRS analyses in Fletcher et al. (2009a), and were converted to k-distributions for IRIS modelling using a Hamming line shape of full-width-at-half-maximum of 4.3 cm−1. The distribution of para-H2 at 330 mbar is compared to ammonia, phosphine and aerosols from Voyager-1/IRIS in Fig. 14. Firstly, we note that the zonal-mean NH3 distributions in Fig. 14b-c largely reproduce those found by Gierasch et al. (1986) using the same dataset, and have mole fractions comparable to those reported from Cassini (Achterberg et al., 2006). The para-H2 distribution also resembles that shown in Fig. 5 of Gierasch et al. (1986). Phosphine in Fig. 14e is found to be enhanced near the equator compared to the neighbouring belts, as confirmed later by Cassini (Irwin et al., 2004; Fletcher et al., 2009a). Both PH3 and NH3 show low-latitude enhancements in the regions of sub-equilibrium para-H2, suggesting that they are being enriched by the same upwelling motions that are respon- sible for the low- fp air at the equator. The aerosol optical depth in Fig. 14d was derived using the 600-1350 cm−1 (7.4-16.7 µm) region alone to ensure that it was independent of any degeneracies between aerosols and para-H2 in the 300-600 cm−1 region. We recover the latitudinal varia- tion of optical depth identified by Gierasch et al. (1986) (their Figure 4), including the high opacity in the 20− 30◦N band. In- triguingly, this does not closely resemble the cloud opacity dis- tributions derived later by Cassini (e.g., Matcheva et al., 2005; Fletcher et al., 2016a), which showed distinct maxima at the equator and the NTrZ and STrZ (northern and southern tropi- cal zones, respectively) and strong minima associated with the NEB and SEB (northern and southern equatorial belts). The absence of the equatorial aerosol enhancement could be related to spatial resolution - the cold equatorial troposphere is also not well resolved by the IRIS data in Fig. 11. 18 However, Fig. 14 makes it clear that para-H2 is not well cor- related with Jupiter's tropospheric aerosols, ammonia and phos- phine distributions. None of these species reflect a decline in the strength of vertical mixing with latitude (e.g., Wang et al., 2015), which again raises the possibility that chemical equili- bration, rather than dynamic mixing, governs the equator-to- pole para-H2 gradient. Our experiments do not yield the cor- relation between para-H2 and cloud opacity favoured by Carl- son et al. (1992), but given the limited sensitivity of thermal-IR spectra to aerosols, we cannot completely reject the aerosol- catalysis hypothesis. Despite the lack of correlation with the tropospheric aerosols in the main cloud deck (Fig. 14d), interconversion of para- and ortho-H2 on the surfaces of aerosols remains an intriguing possibility that has been favoured by previous studies (Carlson et al., 1992; Fouchet et al., 2003). The character of Jupiter's stratospheric aerosols is known to change at the highest lati- tudes. Inversions of Cassini/ISS visible-light observations by Zhang et al. (2013) show that the number density and optical depth of stratospheric aerosols increases by orders of magni- tude from the equator to the poles (with a peak in the 20-50 mbar region), and that this haze extends down to the tropopause in the ±60◦ latitude region. The increased availability of these upper tropospheric and stratospheric hazes at Jupiter's high lat- itudes could increase the efficiency of aerosol catalysis there, explaining why para-H2 is closer to equilibrium at high lati- tudes compared to low latitudes. Furthermore, there is a subtle asymmetry in the stratospheric haze optical thickness (Fig. 12 of Zhang et al., 2013), with more aerosol loading in the north- ern hemisphere than the south, potentially as a result of the au- roral generation mechanism. If the high-latitude fp enrichment is indeed linked to the hazes, and not to the deeper tropospheric clouds, then this could explain the observed asymmetry. Mid- and far-infrared spectral inversions are insensitive to these haze particles because of their small radii and low optical depths, so they could not be derived from the IRIS and FORCAST data. At Jupiter's high latitudes, we propose that fp is close to equilibrium due to the efficiency of equilibration on the upper tropospheric/stratospheric aerosols compared to low-latitudes, potentially with super-equilibrium conditions created by subsi- dence over the poles (provided a region of high- fp air resides above the altitude range of far-IR sensitivity). Conversely, the low-latitude sub-equilibrium fp is due to strong vertical mix- ing in the tropics, as well as inefficient hydrogen equilibration due to a clearer upper troposphere. In this regard, the large- scale equator-to-pole gradient is related to the availability of aerosols to catalyse the equilibration, whereas the local-scale features (the strong perturbations at the equator and high lat- itudes) are the result of vertical advection and mixing. This upwelling might not be uniform over the equatorial zone, but confined to temporally-variable plume activity, similar to the low- fp air that was produced by Saturn's northern springtime storm (Achterberg et al., 2014). Higher resolution observations of fp at all latitudes will be required to better correlate para-H2 with dynamic phenomena. Finally, Saturn also exhibits sub-equilibrium conditions at the equator, but with high-latitude fp that changes with time Figure 14: Comparison of para-H2 with tropospheric properties derived from Voyager-1/IRIS spectra. The solid line in panel (a) shows the equilibrium para- H2 fraction based on the retrieved temperatures, showing that the majority of Jupiter's atmosphere is sub-equilibrium. Panels b-e show the ammonia, phos- phine and aerosol distributions derived from the full 300-1350 cm−1 IRIS spec- trum (all previous figures have used only the 300-570 cm−1 region). 19 (b) Ammonia Mole Fraction at 440 mbar6050403020100-10-20-30-40-50-60Planetographic Latitude02468Mole Fraction [ppm](c) Ammonia Mole Fraction at 860 mbar6050403020100-10-20-30-40-50-60Planetographic Latitude050100150200250300350Mole Fraction [ppm](d) Aerosol Optical Depth at 1 bar6050403020100-10-20-30-40-50-60Planetographic Latitude0.00.20.40.60.81.01.21.4Optical Depth(e) Phosphine Mole Fraction at 650 mbar6050403020100-10-20-30-40-50-60Planetographic Latitude0.00.51.01.5Mole Fraction [ppm](a) Para-H2 Fraction at 330 mbar6050403020100-10-20-30-40-50-60Planetographic Latitude0.290.300.310.320.330.340.350.36Para-H2 Fraction (Fletcher et al., 2016b). During Cassini's observations of Sat- urn, a region of high- fp air has been steadily developing in Saturn's northern hemisphere (poleward of ∼ 50◦N) as Saturn passed northern spring equinox and approached summer sol- stice. Fletcher et al. (2016b) suggested that this was due to tro- pospheric subsidence in the springtime hemisphere and that it could be related to the growth of springtime aerosols. Jupiter's own fp is unlikely to vary seasonally due to the low obliquity, which would suggest that the observed asymmetry in the FOR- CAST and Voyager-1 data could be relatively stable. However, we have no explanation for why the high- fp airmass is absent from the Voyager-2 observations of the northern hemisphere. 5. Conclusions Observations of Jupiter's far-infrared spectrum from the FORCAST instrument on the SOFIA aircraft in May 2014 have provided spectroscopic maps of Jupiter with a 2-4" spatial resolution. The spectra cover the hydrogen-helium collision- induced absorption in the 17-37 µm region that is largely in- accessible to ground-based observatories. These spectra have been inverted to determine Jupiter's temperature and para-H2 distribution as a function of latitude and pressure in the 70-700 mbar region. Despite the low spatial resolution, this is the first time that such a measurement has been possible since the Voy- ager flybys of Jupiter in 1979, and permit a quantitative com- parison between the two epochs. Jupiter's disc-integrated brightness varies from 122.8± 3.7 K at 31 µm to 136.2±4.7 K at 37.1 µm. Tropospheric temperatures observed by FORCAST agree with those found by Voyager, and show the following characteristics: (i) Jupiter's cold equatorial zone has limited influence on the far-IR spectra and shows low contrast with the neighbouring warm equatorial belts; (ii) the temperate banded structure is blended together to form broad, cool zones between 20 − 40◦ which have varied in latitudinal size between the Voyager and SOFIA datasets; (iii) longitudi- nal undulations of tropospheric brightness on the NEB are well- correlated between the M (5.4 µm) and N-band (8-11 µm) ob- servations, suggestive of wave activity affecting cloud opacity, tropospheric temperatures, and ammonia humidity throughout a broad region of the atmosphere between the tropopause and the cloud decks; (iv) both the Voyager and SOFIA inversions suggest the presence of cold vortices poleward of ±50◦ latitude and 100 < p < 300 mbar; and (v) stratospheric heating asso- ciated with the northern auroral oval is detectable by SOFIA as increased brightness near 180◦ in the 7.7- and 11.1-µm filters sensing stratospheric methane and ethane, respectively. Many of these features are better resolved at mid-infrared wavelengths with larger observatories, although the key strength of SOFIA is the ability to derive the para-H2 distri- bution from the relative strength of the S(0) and S(1) absorp- tion features. The SOFIA results are comparable with previous Voyager findings that show an equator-to-pole increase in the para-H2 fraction, with low- fp and sub-equilibrium conditions at the equator and high- fp and weak super-equilibrium condi- tions at ±60◦ latitude. We confirm the presence of a region of high- fp air at high northern latitudes that appears asymmet- ric when compared to the south, which was previously seen by Voyager 1 (Gierasch et al., 1986; Conrath et al., 1998). This region is therefore a repeatable feature, although it does not ap- pear to be permanent (it was not evident in the noisier Voyager 2 data), and the magnitude of the super-equilibrium depends on our retrieval assumptions. We note that the choice of collision- induced H2-He opacity influences our quantitative results, and that the use of recent opacity tables (Orton et al., 2007; Richard et al., 2012) worsens the fits to FORCAST and IRIS spectra when compared to the original (erroneous) opacity results from Borysow et al. (1985). We suggest that this is due to missing (H2)2 dimer absorption in these tables, which increases the ab- sorption coefficient near the S(0) and S(1) lines to values closer to those of Borysow et al. (1985). Inversions of Voyager-1 mid-infrared spectra revealed that the latitudinal distributions of tropospheric ammonia, phos- phine and cloud opacity at ∼ 800 mbar are not correlated with the observed distribution of para-H2 in the SOFIA and Voyager observations. Nevertheless, we note a similarity between the para-H2 distribution and that of small-radii aerosols comprising the upper tropospheric and stratospheric hazes (Zhang et al., 2013). These hazes are also asymmetric, with a greater aerosol mass loading in the north polar regions where we find the cold- est tropopause temperatures and the highest para-H2 fraction. We therefore propose that the para-H2 equator-to-pole gradient is governed primarily by the efficiency of paramagnetic con- version catalysed by aerosols in Jupiter's hazes (rather than in the cloud decks sensed in the mid- and far-infrared); with sec- ondary perturbations from localised circulations. In this sce- nario, the sub-equilibrium conditions at low latitudes are due to the clear-atmosphere conditions and inefficient equilibration, coupled with strong upwelling enriching low- fp air at the equa- tor. Conversely, high-latitude air is closer to equilibrium due to efficient hydrogen conversion and weaker upwelling, and may be locally enhanced in the north by subsidence beneath the cold polar vortex. Observations of Jupiter at higher spatial resolutions could start to distinguish the competing dynamical and chemical in- fluences on the para-H2 distribution, particularly if fp is found to be substantially sub-equilibrium in localised plumes at the equator and other latitudes, such as those found on Saturn (Achterberg et al., 2014). Furthermore, the strength of the para- H2 asymmetry and the north polar anomaly could be tracked with a time series of measurements at various seasons during Jupiter's 11.8-year orbit. Unfortunately Jupiter's H2 continuum is beyond the saturation limits for the James Webb Space Tele- scope (which could have reached the S(0) line near 350 cm−1 (28.5 µm), Norwood et al., 2016), although this facility should certainly be used to conduct similar studies of the para-H2 dis- tributions on Uranus and Neptune. Acknowledgments This work was based on observations made with the NASA/DLR Stratospheric Observatory for Infrared Astron- omy (SOFIA), and was financially supported through SOF0012 20 to the University of California, Berkeley. SOFIA is jointly operated by the Universities Space Research Association, Inc. (USRA), under NASA contract NAS2-97001, and the Deutsches SOFIA Institut (DSI) under DLR contract 50 OK 0901 to the University of Stuttgart. We are grateful for all those involved in the telescope engineering, operations support and the flight crews. We thank Luke Keller, Matthew Bellardini and Joseph Quinn (Ithaca College) and Joseph Adams for their as- sistance with the initial planning and calibration of the SOFIA data. Fletcher was supported by a Royal Society Research Fel- lowship at the University of Leicester. The UK authors ac- knowledge the support of the Science and Technology Facili- ties Council (STFC). A portion of this work was performed by Orton at the Jet Propulsion Laboratory, California Institute of Technology, under a contract with NASA. Gehrz received par- tial support from the United States Air Force. This research used the ALICE High Performance Computing Facility at the University of Leicester. References Achterberg, R.K., Conrath, B.J., Gierasch, P.J., 2006. Cassini CIRS retrievals of Ammonia in Jupiter's Upper Troposphere. Icarus 182, 169 -- 180. doi:10. 1016/j.icarus.2005.12.020. Achterberg, R.K., Gierasch, P.J., Conrath, B.J., Fletcher, L.N., Hesman, B.E., Bjoraker, G.L., Flasar, F.M., 2014. Changes to Saturn's Zonal-mean Tropo- spheric Thermal Structure after the 2010-2011 Northern Hemisphere Storm. ApJ 786, 92. doi:10.1088/0004-637X/786/2/92. Adams, J.D., Herter, T.L., Gull, G.E., Schoenwald, J., Henderson, C.P., Keller, L.D., De Buizer, J.M., Stacey, G.J., Nikola, T., 2010. FORCAST: a first light facility instrument for SOFIA, in: Ground-based and Airborne Instru- mentation for Astronomy III, p. 77351U. doi:10.1117/12.857049. Atreya, S.K., Wong, M.H., Owen, T.C., Mahaffy, P.R., Niemann, H.B., de Pa- ter, I., Drossart, P., Encrenaz, T., 1999. A comparison of the atmospheres of Jupiter and Saturn: deep atmospheric composition, cloud structure, vertical mixing, and origin. Plan. & Space Sci. 47, 1243 -- 1262. Birnbaum, G., Borysow, A., Orton, G.S., 1996. Collision-Induced Absorption of H2-H2 and H2-He in the Rotational and Fundamental Bands for Planetary Applications. Icarus 123, 4 -- 22. doi:10.1006/icar.1996.0138. Borysow, J., Trafton, L., Frommhold, L., Birnbaum, G., 1985. Modeling of pressure-induced far-infrared absorption-spectra: molecular-hydrogen pairs. Astrophys. J. 296, 644 -- 654. Burke, T.E., 1974. The Mariner Jupiter/Saturn IRIS experiment, in: Palluconi, F.D., Pettengill, G.H. (Eds.), The Saturn's Rings Workshop. Caldwell, J., Gillett, F.C., Tokunaga, A.T., 1980. Possible infrared aurorae on Jupiter. Icarus 44, 667 -- 675. doi:10.1016/0019-1035(80)90135-9. Carlson, B.E., Lacis, A.A., Rossow, W.B., 1992. Ortho-para-hydrogen equili- bration on Jupiter. Astrophys. J. 393, 357 -- 372. doi:10.1086/171510. Conrath, B.J., Flasar, F.M., Pirraglia, J.A., Gierasch, P.J., Hunt, G.E., 1981a. Thermal structure and dynamics of the Jovian atmosphere. II - Visible cloud features. Journal of Geophysical Research 86, 8769 -- 8775. doi:10.1029/ JA086iA10p08769. Conrath, B.J., Gautier, D., 1980. Thermal structure of Jupiter's atmosphere ob- tained by inversion of Voyager 1 infrared measurements, in: Remote Sensing of Atmospheres and Oceans, pp. 611 -- 628. Conrath, B.J., Gautier, D., 2000. Saturn Helium Abundance: A Reanalysis of Voyager Measurements. Icarus 144, 124 -- 134. doi:10.1006/icar.1999. 6265. Conrath, B.J., Gierasch, P.J., 1984. Global variation of the para hydrogen frac- tion in Jupiter's atmosphere and implications for dynamics on the outer plan- ets. Icarus 57, 184 -- 204. doi:10.1016/0019-1035(84)90065-4. Conrath, B.J., Gierasch, P.J., Leroy, S.S., 1990. Temperature and circulation in the stratosphere of the outer planets. Icarus 83, 255 -- 281. doi:10.1016/ 0019-1035(90)90068-K. Conrath, B.J., Gierasch, P.J., Nath, N., 1981b. Stability of zonal flows on Jupiter. Icarus 48, 256 -- 282. doi:10.1016/0019-1035(81)90108-1. Conrath, B.J., Gierasch, P.J., Ustinov, E.A., 1998. Thermal Structure and Para Hydrogen Fraction on the Outer Planets from Voyager IRIS Measurements. Icarus 135, 501 -- 517. doi:10.1006/icar.1998.6000. Conrath, B.J., Pirraglia, J.A., 1983. Thermal structure of Saturn from Voyager infrared measurements - Implications for atmospheric dynamics. Icarus 53, 286 -- 292. doi:10.1016/0019-1035(83)90148-3. Flasar, F.M., Conrath, B.J., Pirraglia, J., Clark, P.C., French, R.G., Gierasch, P.J., 1981. Thermal structure and dynamics of the Jovian atmosphere. I - The Great Red Spot. J. Geophys. Res. 86, 8759 -- 8767. doi:10.1029/ JA086iA10p08759. Flasar, F.M., Kunde, V.G., Achterberg, R.K., Conrath, B.J., Simon-Miller, A.A., Nixon, C.A., Gierasch, P.J., Romani, P.N., B´ezard, B., Irwin, P., Bjo- raker, G.L., Brasunas, J.C., Jennings, D.E., Pearl, J.C., Smith, M.D., Orton, G.S., Spilker, L.J., Carlson, R., Calcutt, S.B., Read, P.L., Taylor, F.W., Par- rish, P., Barucci, A., Courtin, R., Coustenis, A., Gautier, D., Lellouch, E., Marten, A., Prang´e, R., Biraud, Y., Fouchet, T., Ferrari, C., Owen, T.C., Ab- bas, M.M., Samuelson, R.E., Raulin, F., Ade, P., C´esarsky, C.J., Grossman, K.U., Coradini, A., 2004. An intense stratospheric jet on Jupiter. Nature 427, 132 -- 135. Fletcher, L.N., de Pater, I., Orton, G.S., Hammel, H.B., Sitko, M.L., Ir- win, P.G.J., 2014. Neptune at summer solstice: Zonal mean tempera- tures from ground-based observations, 2003-2007. Icarus 231, 146 -- 167. doi:10.1016/j.icarus.2013.11.035, arXiv:1311.7570. Fletcher, L.N., Greathouse, T.K., Orton, G.S., Sinclair, J.A., Giles, R.S., Irwin, P.G.J., Encrenaz, T., 2016a. Mid-infrared mapping of Jupiter's temperatures, aerosol opacity and chemical distributions with IRTF/TEXES. Icarus 278, 128 -- 161. doi:10.1016/j.icarus.2016.06.008, arXiv:1606.05498. Fletcher, L.N., Irwin, P.G.J., Achterberg, R.K., Orton, G.S., Flasar, F.M., 2016b. Seasonal variability of Saturn's tropospheric temperatures, winds and para-H2 from Cassini far-IR spectroscopy. Icarus 264, 137 -- 159. doi:10.1016/j.icarus.2015.09.009, arXiv:1509.02281. Fletcher, L.N., Orton, G.S., Rogers, J.H., Simon-Miller, A.A., de Pater, I., Wong, M.H., Mousis, O., Irwin, P.G.J., Jacquesson, M., Yanamandra-Fisher, P.A., 2011. Jovian temperature and cloud variability during the 2009-2010 fade of the South Equatorial Belt. Icarus 213, 564 -- 580. doi:10.1016/j. icarus.2011.03.007. Fletcher, L.N., Orton, G.S., Teanby, N.A., Irwin, P.G.J., 2009a. Phosphine on Jupiter and Saturn from Cassini/CIRS. Icarus 202, 543 -- 564. doi:10.1016/ j.icarus.2009.03.023. Fletcher, L.N., Orton, G.S., Yanamandra-Fisher, P., Fisher, B.M., Parrish, P.D., Irwin, P.G.J., 2009b. Retrievals of atmospheric variables on the gas giants from ground-based mid-infrared imaging. Icarus 200, 154 -- 175. doi:10. 1016/j.icarus.2008.11.019. Fouchet, T., Lellouch, E., Feuchtgruber, H., 2003. The hydrogen ortho-to- Icarus 161, 127 -- 143. para ratio in the stratospheres of the giant planets. doi:10.1016/S0019-1035(02)00014-3. Frommhold, L., Samuelson, R., Birnbaum, G., 1984. Hydrogen dimer struc- tures in the far-infrared spectra of Jupiter and Saturn. ApJ Letters 283, L79 -- L82. doi:10.1086/184338. Gautier, D., Conrath, B., Flasar, M., Hanel, R., Kunde, V., Chedin, A., Scott, N., 1981. The helium abundance of Jupiter from Voyager. Journal of Geo- physical Research 86, 8713 -- 8720. doi:10.1029/JA086iA10p08713. Gehrz, R.D., Becklin, E.E., de Pater, I., Lester, D.F., Roellig, T.L., Woodward, C.E., 2009. A new window on the cosmos: The Stratospheric Observatory for Infrared Astronomy (SOFIA). Advances in Space Research 44, 413 -- 432. doi:10.1016/j.asr.2009.04.011. Gierasch, P.J., Magalhaes, J.A., Conrath, B.J., 1986. Zonal mean properties of Jupiter's upper troposphere from Voyager infrared observations. Icarus 67, 456 -- 483. doi:10.1016/0019-1035(86)90125-9. Griffith, C., B´ezard, B., Owen, T., Gautier, D., 1992. The tropospheric abun- dances of nh3 and ph3 in jupiter's great red spot, from voyager iris observa- tions. Icarus 98, 82 -- 93. Hanel, R., Conrath, B., Flasar, F.M., Kunde, V., Maguire, W., Pearl, J.C., Pir- raglia, J., Samuelson, R., Cruikshank, D.P., Gautier, D., Gierasch, P.J., Horn, L., Ponnamperuma, C., 1982. Infrared observations of the Saturnian system from Voyager 2. Science 215, 544 -- 548. Hanel, R., Conrath, B., Flasar, M., Kunde, V., Lowman, P., Maguire, W., Pearl, J., Pirraglia, J., Samuelson, R., Gautier, D., Gierasch, P., Kumar, S., Pon- namperuma, C., 1979. Infrared observations of the Jovian system from Voy- ager 1. Science 204, 972 -- 976. doi:10.1126/science.204.4396.972-a. Hanel, R., Conrath, B., Kunde, V., Lowman, P., Maguire, W., Pearl, J., Pirraglia, 21 icarus.2015.07.004. J., Gautier, D., Gierasch, P., Kumar, S., 1977. The Voyager infrared spec- troscopy and radiometry investigation. Space Science Reviews 21, 129 -- 157. doi:10.1007/BF00200848. Herter, T.L., Adams, J.D., De Buizer, J.M., Gull, G.E., Schoenwald, J., Hender- son, C.P., Keller, L.D., Nikola, T., Stacey, G., Vacca, W.D., 2012. First Sci- ence Observations with SOFIA/FORCAST: The FORCAST Mid-infrared Camera. ApJ Letters 749, L18. doi:10.1088/2041-8205/749/2/L18, arXiv:1202.5021. Herter, T.L., Vacca, W.D., Adams, J.D., Keller, L.D., Schoenwald, J., Hirsch, L., Wang, J., De Buizer, J.M., Helton, L.A., Llorens, M.C., 2013. Data Reduction and Early Science Calibration for FORCAST, A Mid-Infrared Camera for SOFIA. PASP 125, 1393 -- 1404. doi:10.1086/674144. Irwin, P., Teanby, N., de Kok, R., Fletcher, L., Howett, C., Tsang, C., Wilson, C., Calcutt, S., Nixon, C., Parrish, P., 2008. The NEMESIS planetary at- mosphere radiative transfer and retrieval tool. Journal of Quantitative Spec- troscopy and Radiative Transfer 109, 1136 -- 1150. Irwin, P.G.J., Parrish, P., Fouchet, T., Calcutt, S.B., Taylor, F.W., Simon-Miller, A.A., Nixon, C.A., 2004. Retrievals of jovian tropospheric phosphine from Cassini/CIRS. Icarus 172, 37 -- 49. doi:10.1016/j.icarus.2003.09.027. Keller, L., Deen, C.P., Jaffe, D.T., Ennico, K.A., Greene, T.P., Adams, J.D., Herter, T., Sloan, G.C., 2010. Progress report on FORCAST grism spec- troscopy as a future general observer instrument mode on SOFIA, in: Ground-based and Airborne Instrumentation for Astronomy III, p. 77356N. doi:10.1117/12.857127. Kessler, M.F., Steinz, J.A., Anderegg, M.E., Clavel, J., Drechsel, G., Estaria, P., Faelker, J., Riedinger, J.R., Robson, A., Taylor, B.G., Xim´enez de Ferr´an, S., 1996. The Infrared Space Observatory (ISO) mission. Astron. Astrophys 315, L27 -- L31. Kostiuk, T., Romani, P., Espenak, F., Livengood, T.A., 1993. Temperature and abundances in the Jovian auroral stratosphere. 2: Ethylene as a probe of the microbar region. Journal of Geophysical Research 98, 18. doi:10.1029/ 93JE01332. Leovy, C.B., Friedson, A.J., Orton, G.S., 1991. The quasiquadrennial oscil- lation of Jupiter's equatorial stratosphere. Nature 354, 380 -- 382. doi:10. 1038/354380a0. Livengood, T.A., Kostiuk, T., Espenak, F., 1993. Temperature and abundances in the Jovian auroral stratosphere. 1: Ethane as a probe of the millibar region. Journal of Geophysical Research 98, 18. doi:10.1029/93JE01043. Martonchik, J.V., Orton, G.S., Appleby, J.F., 1984. Optical properties of NH3 ice from the far infrared to the near ultraviolet. Applied Optics 23, 541 -- 547. Massie, S.T., Hunten, D.M., 1982. Conversion of para and ortho hydrogen in the Jovian planets. Icarus 49, 213 -- 226. doi:10.1016/0019-1035(82) 90073-2. Matcheva, K., Conrath, B., Gierasch, P., Flasar, F., 2005. The cloud structure of the jovian atmosphere as seen by the Cassini/CIRS experiment. Icarus 179, 432 -- 448. Meyer, W., Borysow, A., Frommhold, L., 1989a. Absorption spectra of H2-H2 pairs in the fundamental band. Physical Review A 40, 6931 -- 6949. doi:10. 1103/PhysRevA.40.6931. Meyer, W., Frommhold, L., Birnbaum, G., 1989b. Rototranslational absorption spectra of H2-H2 pairs in the far infrared. Physical Review A 39, 2434 -- 2448. doi:10.1103/PhysRevA.39.2434. Minnaert, M., 1941. The reciprocity principle in lunar photometry. Astrophys. J. 93, 403 -- 410. doi:10.1086/144279. Niemann, H.B., Atreya, S.K., Carignan, G.R., Donahue, T.M., Haberman, J.A., Harpold, D.N., Hartle, R.E., Hunten, D.M., Kasprzak, W.T., Mahaffy, P.R., Owen, T.C., Way, S.H., 1998. The composition of the Jovian atmosphere as determined by the Galileo probe mass spectrometer. J. Geophys. Res. 103, 22831 -- 22846. doi:10.1029/98JE01050. Nixon, C.A., Achterberg, R.K., Romani, P.N., Allen, M., Zhang, X., Teanby, N.A., Irwin, P.G.J., Flasar, F.M., 2010. Abundances of Jupiter's trace hy- drocarbons from Voyager and Cassini. Plan. & Space Sci. 58, 1667 -- 1680. doi:10.1016/j.pss.2010.05.008, arXiv:1005.3959. Norwood, J., Moses, J., Fletcher, L.N., Orton, G., Irwin, P.G.J., Atreya, S., Rages, K., Cavali´e, T., S´anchez-Lavega, A., Hueso, R., Chanover, N., 2016. Giant Planet Observations with the James Webb Space Tele- scope. PASP 128, 018005. doi:10.1088/1538-3873/128/959/018005, arXiv:1510.06205. Orton, G.S., Fletcher, L.N., Encrenaz, T., Leyrat, C., Roe, H.G., Fujiyoshi, T., Pantin, E., 2015. Thermal imaging of Uranus: Upper-tropospheric tem- peratures one season after Voyager. Icarus 260, 94 -- 102. doi:10.1016/j. 22 Orton, G.S., Fletcher, L.N., Moses, J.I., Mainzer, A.K., Hines, D., Hammel, H.B., Martin-Torres, F.J., Burgdorf, M., Merlet, C., Line, M.R., 2014. Mid- infrared spectroscopy of Uranus from the Spitzer Infrared Spectrometer: 1. Determination of the mean temperature structure of the upper troposphere and stratosphere. Icarus 243, 494 -- 513. doi:10.1016/j.icarus.2014. 07.010, arXiv:1407.2120. Orton, G.S., Friedson, A.J., Caldwell, J., Hammel, H.B., Baines, K.H., Bergstralh, J.T., Martin, T.Z., Malcom, M.E., West, R.A., Golisch, W.F., Griep, D.M., Kaminski, C.D., Tokunaga, A.T., Baron, R., Shure, M., 1991. Thermal maps of Jupiter - Spatial organization and time dependence of stratospheric temperatures, 1980 to 1990. Science 252, 537 -- 542. Orton, G.S., Friedson, A.J., Yanamandra-Fisher, P.A., Caldwell, J., Hammel, H.B., Baines, K.H., Bergstralh, J.T., Martin, T.Z., West, R.A., Veeder, Jr., G.J., Lynch, D.K., Russell, R., Malcom, M.E., Golisch, W.F., Griep, D.M., Kaminski, C.D., Tokunaga, A.T., Herbst, T., Shure, M., 1994. Spatial Or- ganization and Time Dependence of Jupiter's Tropospheric Temperatures, 1980-1993. Science 265, 625 -- 631. doi:10.1126/science.265.5172. 625. Orton, G.S., Gustafsson, M., Burgdorf, M., Meadows, V., 2007. Revised Ab Initio Models for H2-H2 Collision Induced Absorption at Low Tempera- tures. Icarus 189, 544 -- 549. Read, P., Gierasch, P., Conrath, B., 2006. Mapping potential-vorticity dynamics on Jupiter. II: The Great Red Spot from Voyager 1 and 2 data. Quarterly Journal of the Royal Meteorological Society 132, 1605 -- 1625. Richard, C., Gordon, I.E., Rothman, L.S., Abel, M., Frommhold, L., Gustafs- son, M., Hartmann, J.M., Hermans, C., Lafferty, W.J., Orton, G.S., Smith, K.M., Tran, H., 2012. New section of the HITRAN database: Collision- induced absorption (CIA). Journal of Quantitative Spectroscopy and Radia- tive Transfer 113, 1276 -- 1285. doi:10.1016/j.jqsrt.2011.11.004. Rodgers, C.D., 2000. Inverse Methods for Atmospheric Remote Sounding: Theory and Practice. World Scientific. Sada, P.V., Beebe, R.F., Conrath, B.J., 1996. Comparison of the Structure and Dynamics of Jupiter's Great Red SPOT between the Voyager 1 and 2 En- counters. Icarus 119, 311 -- 335. doi:10.1006/icar.1996.0022. Schaefer, J., McKellar, A.R.W., 1990. Faint features of the rotational S0(0) and S0(1) transitions of H2. A comparison of calculations and measurements at 77 K. Zeitschrift fur Physik D Atoms Molecules Clusters 15, 51 -- 65. doi:10.1007/BF01436911. Seiff, A., Kirk, D.B., Knight, T.C.D., Young, R.E., Mihalov, J.D., Young, L.A., Milos, F.S., Schubert, G., Blanchard, R.C., Atkinson, D., 1998. Thermal structure of Jupiter's atmosphere near the edge of a 5-µm hot spot in the north equatorial belt. J. Geophys. Res. 103, 22857 -- 22890. doi:10.1029/ 98JE01766. Simon-Miller, A., Gierasch, P., Beebe, R., Conrath, B., Flasar, F., Achterberg, R., 2002. New Observational Results Concerning Jupiter's Great Red Spot. Icarus 158, 249 -- 266. Simon-Miller, A.A., Conrath, B., Gierasch, P.J., Beebe, R.F., 2000. A detection of water ice on Jupiter with Voyager IRIS. Icarus 145, 454 -- 461. doi:10. 1006/icar.2000.6359. Simon-Miller, A.A., Conrath, B.J., Gierasch, P.J., Orton, G.S., Achterberg, R.K., Flasar, F.M., Fisher, B.M., 2006. Jupiter's atmospheric temperatures: From Voyager IRIS to Cassini CIRS. Icarus 180, 98 -- 112. doi:10.1016/j. icarus.2005.07.019. Sinclair, J.A., Irwin, P.G.J., Fletcher, L.N., Greathouse, T., Guerlet, S., Hur- ley, J., Merlet, C., 2014. From Voyager-IRIS to Cassini-CIRS: Interannual variability in Saturn's stratosphere? Icarus 233, 281 -- 292. doi:10.1016/j. icarus.2014.02.009. Wang, D., Gierasch, P.J., Lunine, J.I., Mousis, O., 2015. New insights on Icarus 250, Jupiter's deep water abundance from disequilibrium species. 154 -- 164. doi:10.1016/j.icarus.2014.11.026, arXiv:1412.0690. Wong, M.H., Bjoraker, G.L., Smith, M.D., Flasar, F.M., Nixon, C.A., 2004. Identification of the 10-µm ammonia ice feature on Jupiter. Plan. & Space Sci. 52, 385 -- 395. doi:10.1016/j.pss.2003.06.005. Young, E.T., Becklin, E.E., Marcum, P.M., Roellig, T.L., De Buizer, J.M., Herter, T.L., Gusten, R., Dunham, E.W., Temi, P., Andersson, B.G., Back- man, D., Burgdorf, M., Caroff, L.J., Casey, S.C., Davidson, J.A., Erick- son, E.F., Gehrz, R.D., Harper, D.A., Harvey, P.M., Helton, L.A., Horner, S.D., Howard, C.D., Klein, R., Krabbe, A., McLean, I.S., Meyer, A.W., Miles, J.W., Morris, M.R., Reach, W.T., Rho, J., Richter, M.J., Roeser, H.P., Sandell, G., Sankrit, R., Savage, M.L., Smith, E.C., Shuping, R.Y., Vacca, W.D., Vaillancourt, J.E., Wolf, J., Zinnecker, H., 2012. Early Science with SOFIA, the Stratospheric Observatory For Infrared Astronomy. ApJ Letters 749, L17. doi:10.1088/2041-8205/749/2/L17, arXiv:1205.0791. Zhang, X., West, R.A., Banfield, D., Yung, Y.L., 2013. Stratospheric aerosols on Jupiter from Cassini observations. Icarus 226, 159 -- 171. doi:10.1016/ j.icarus.2013.05.020. 23
0904.1661
1
0904
2009-04-10T08:42:38
Supernova Propagation And Cloud Enrichment: A new model for the origin of $^{60}$Fe in the early solar system
[ "astro-ph.EP" ]
The radioactive isotope $^{60}$Fe ($T_{1/2} = 1.5 $ Myr) was present in the early solar system. It is unlikely that it was injected directly into the nascent solar system by a single, nearby supernova. It is proposed instead that it was inherited during the molecular cloud stage from several supernovae belonging to previous episodes of star formation. The expected abundance of $^{60}$Fe in star forming regions is estimated taking into account the stochasticity of the star-forming process, and it is showed that many molecular clouds are expected to contain $^{60}$Fe (and possibly $^{26}$Al [$T_{1/2} = 0.74 $ Myr]) at a level compatible with that of the nascent solar system. Therefore, no special explanation is needed to account for our solar system's formation.
astro-ph.EP
astro-ph
In press, Astrophysical Journal Letters Supernova Propagation And Cloud Enrichment: A new model for the origin of 60Fe in the early solar system Matthieu Gounelle1, Anders Meibom1, Patrick Hennebelle2 & Shu-ichiro Inutsuka3 [email protected] ABSTRACT The radioactive isotope 60Fe (T1/2 = 1.5 Myr) was present in the early solar system. It is unlikely that it was injected directly into the nascent solar system by a single, nearby supernova. It is proposed instead that it was inherited during the molecular cloud stage from several supernovae belonging to previous episodes of star formation. The expected abundance of 60Fe in star forming regions is estimated taking into account the stochasticity of the star-forming process, and it is showed that many molecular clouds are expected to contain 60Fe (and possibly 26Al [T1/2 = 0.74 Myr]) at a level compatible with that of the nascent solar system. Therefore, no special explanation is needed to account for our solar system's formation. Subject headings: solar system, formation, meteors, ISM: clouds, ISM:evolution, supernovae: general 1. Introduction Short-lived radionuclei (SLRs) are radioactive isotopes with half-lives shorter than 100 Myr, which were present in the early solar system (ESS, Russell et al. 2001). Because of 1Laboratoire d'´Etude la Mati`ere Extraterrestre, Mus´eum National d'Histoire Naturelle, 57 rue Cuvier, 75 005 Paris, France. 2Laboratoire de Radioastronomie Millim´etrique, ´Ecole Normale Sup´erieure et Observatoire de Paris, 24 rue Lhomond, 75005 Paris, France. 3Department of Physics, Kyoto University, Kyoto, 606-8502, Japan. -- 2 -- their relatively high abundances with respect to that of the interstellar medium (ISM), some SLRs must have been produced within, or close in space and time to the ESS rather than during continuous Galactic nucleosynthesis (e.g. Meyer & Clayton 2000). Iron-60 (T1/2 = 1.5 Myr) holds a special position because it is only produced efficiently by stellar nucleosynthesis unlike other SLRs, which can also be made in the protoplanetary disk via irradiation of dust/gas by accelerated energetic particles such as protons (Lee et al. 1998). As such, 60Fe provides important clues about the immediate stellar environment of the nascent solar system (Montmerle et al. 2006). AGB stars are not considered a likely source of 60Fe in the solar system because of their low probability of encounter with a star-forming region (Kastner & Myers 1994). Elaborating on the pioneering work of Cameron & Truran (1977), two different quanti- tative scenarios with nearby, single supernova have been proposed whereby 60Fe is injected either into the solar protoplanetary disk (e.g. Ouellette et al. 2007) or into the molecular cloud (MC) core progenitor of our solar system (e.g. Cameron et al. 1995). Some models envision both possibilities (Takigawa et al. 2008). In the supernova-disk scenario, the small size of the disk requires that it lies within 0.4 pc from the injecting supernova belonging to the same stellar cluster (Ouellette et al. 2007). However, when massive stars become supernovae after a few Myr of evolution, remaining disks around low-mass stars are several pc away from the massive star (e.g. Sicilia-Aguilar et al. 2005), and receive only minute amounts of SLRs (Williams & Gaidos 2007; Gounelle & Meibom 2008). In the supernova- core scenario, the supernova shockwave triggers the core gravitational collapse in addition to delivering SLRs only for very restricted conditions in term of distance and shockwave velocity (Boss et al. 2008). At present, there is therefore no satisfying model which can explain the presence of 60Fe in the ESS. Here, we quantitatively evaluate a scenario proposing that 60Fe was inherited in the progenitor molecular cloud from supernovae belonging to previous episodes of star formation. This scenario differs from scenarios favoring direct injection into a disk/core from a contemporaneous, single, nearby supernova mainly because the 60Fe adduction occurs at the larger MC scale. As previous stellar models trying to account for the presence of 60Fe in the early solar system (Ouellette et al. 2007; Meyer & Clayton 2000; Mostefaoui et al. 2005, Boss et al. 2008), our model cannot solve the problem of 53Mn (T1/2 = 3.7 Myr) overproduction relative to 60Fe and their relative abundances in the early solar system (e.g. Wasserburg et al. 2006). -- 3 -- 2. The Supernova Propagation And Cloud Enrichment Model 2.1. Model sketch Recently, a new paradigm concerning the formation mechanisms and lifetimes of molecu- lar clouds emerged (see Hennebelle et al. 2007 and references therein). In this new paradigm, referred to as the turbulent convergent flow model, MCs result from the collision of coherent flows and large-scale shocks in the interstellar medium driven by winds from massive stars and supernova explosions. Such collisions compress the interstellar atomic gas and after 10 to 20 Myr of evolution, the gas is dense enough to be shielded from the UV radiation and to become molecular (e.g. Glover & MacLow 2007). Star formation follows immediately after the formation of the dense molecular gas. The turbulent convergent flow model provides a natural explanation for the wind-swept appearance of molecular clouds (V´azquez-Semadeni et al. 2005), but is also consistent with the short lifetime of MCs (Hartmann et al. 2001) and the formation of stars in molecular clouds within a crossing time (Elmegreen 2000). In addition, the turbulent convergent flow model elegantly accounts for the division of OB associations in subgroups of different ages (Lada & Lada 2003). A famous example is the Scorpio-Centaurus region made of the Lower Centaurus Crux (LCC, ∼ 16 Myr), the Upper Centaurus Lupus (UCL, ∼ 17 Myr) and the Upper Scorpius (Upper Sco, ∼ 5 Myr) subregions (Preibisch & Zinnecker 2007). If the turbulent convergent flow model is correct, relatively high concentrations of 60Fe and other radioactivities with half-lives & 1 Myr are expected in molecular clouds. This is because supernova ejecta, whose compression effects build molecular clouds, also carry large amount of radioactive elements such as 60Fe. Although it can take as long as 20 Myr to build a molecular cloud depending on the starting density of the atomic gas, live 60Fe is continuously replenished in the second generation molecular cloud by supernovae originating from the first episode of star formation, which explode every few Myr. We therefore suggest that 60Fe in the ESS was inherited from multiple supernovae belonging to previous episodes of star formation and name our model SPACE for Supernova Propagation And Cloud Enrichment. 2.2. Quantitative estimate We consider a first generation of stars formed in molecular cloud 1 (MC1) and a second generation of stars formed in molecular cloud 2 (MC2). After dissipation of the gas, the first (second) generation of stars become the OB1 (OB2) association. In our model, a number of supernovae from MC1 deliver 60Fe into the second generation molecular cloud MC2 (Fig. 1). -- 4 -- The mass of 60Fe in MC2 as a function of time t (time zero being the onset of star formation in MC1) reads: MMC2(60Fe)[t] = f η i=NSN Xi=1 YSNi(60Fe)e−(t−ti)/τ , (1) where f is a geometrical dilution factor, η is the mixing efficiency, NSN is the number of supernovae which have exploded in OB1 before time t, τ is the mean life of 60Fe, YSNi(60Fe) is the 60Fe yield of the ith supernova in MC1 and ti is the time of the ith supernova explosion in MC1. The stellar masses (M) in MC1 are calculated following the stellar Initial Mass Function (1993), M = 0.01 + (0.19ξ1.55 + (IMF), using the generating function of Kroupa et al. 0.05ξ0.6)/(1 − ξ0.58), where ξ is a random number to be chosen between 0 and 1 (Brasser et al. 2006). We consider only the distributions whose most massive star is less massive than 150 M⊙, a likely upper limit for stellar masses (Weidner & Kroupa 2006). Importantly, we find that fSN, the fraction of stars more massive than 8 M⊙ which will go supernova, is 2.3 × 10−3. The yields of 60Fe have been determined for a diversity of supernovae, corresponding to progenitor massive stars with masses ranging from 11 to 120 M⊙ (Woosley & Weaver 1995; Rauscher et al. 2002; Limongi & Chiffi 2006; www.nucleosynthesis.org). Though in relatively good agreement, the yields somehow vary because of differences in the stellar and nuclear physics used by the different groups. For a given mass, when different yields are available, we use the average of the different yields. For stellar masses for which the yields are unknown, we take the yield of the star closest in mass. The explosion time ti of each massive star depends on the mass of the progenitor and is given by the evolutionary tracks of Schaller et al. (1992). The model input parameters are summarized in Table 1. The value of η depends on the efficiency of mixing of ejecta material into cold compressed gas that eventually becomes molecular material. The interface of the ejecta and the shocked ambient medium is expected to be turbulent due to various instabilities. Especially, when the elapsed time of supernova expansion becomes comparable to the cooling timescale of post-shock gas, thermal instability makes the interface highly turbulent. This process is clearly shown in hydrodynamical simulations of the propagation of a shock wave or of a shocked layer into warm neutral medium, which results in the creation of cold turbulent clumps embedded in warm neutral medium via thermal instability (Koyama & Inutsuka 2002; Audit & Hennebelle 2005). The spatial scale of the smallest turbulent eddy is probably comparable to the characteristic size (λC) of the smallest cold clump that is on the order of the critical length scale (< 0.01 pc) of thermal instability. The characteristic mass (MC) -- 5 -- of the smallest cold clumps is much smaller than the solar mass and can be given by the following equation: MC ≡ ρCλ3 C ∼ 10−5M⊙(cid:18) ρC 10−21 g cm−3(cid:19)(cid:18) λC 0.01 pc(cid:19)3 , (2) where ρC is the gas density of cold clumps. Thus, the mixing of the metal-rich ejecta and the ambient medium should be very efficient on this small mass scale. Therefore, we expect the efficiency of mixing to be very high, and use η = 1 hereafter, in line with the value of η adopted by Looney et al. (2006) in the case of a disk and a starless core. Given that in our model, the supernovae from OB1 by definition face MC2 (see Fig. 1), we expect f to be close to 0.5. We conservatively assume that only one tenth of the supernova ejecta contributes to the sweeping-up of atomic gas and the adduction of 60Fe into the new molecular cloud, and therefore adopt f = 0.1. The evolution of 60Fe in MC2 is calculated for different sizes of MC1, i.e. for different values of its number of stars, N1. For each N1, the calculation is realized about 100 times to account for the stochastic nature of star formation. A typical example, with N1 = 5000 stars, is given in Fig. 2 where each thin line represents one realization of the simulation, while the thick red line is the average of 102 realizations. The average number of supernovae is 11.8, varying from from 4 to 22. Increasing the number of realizations does not change the shape of the red average curve depicted in Fig. 2, nor any of the calculated properties, indicating that 100 realizations of the IMF suffice to give a fair account of the 60Fe abundance expected in MC2. The positive slope at small times (or large masses) is due to the rarity of occurrence of very massive stars. The positive slope at t = 16 Myr is due to the high 60Fe yield of the 13 M⊙ star compared to stars with neighboring masses (see Table 1). Note that both Woosley & Weaver (1995) and Limongi & Chiffi (1996) find similar high 60Fe yields for the 13 M⊙ model. Because MCs need between 10 and 20 Myr to be built by turbulent convergent flows (e.g. V´azquez-Semadeni et al. 2007; Hennebelle et al. 2008), we calculate the average of the 60Fe abundance between 10 and 20 Myr, MMC2 (60Fe). In our typical case (N1 = 5000 stars), MMC2 (60Fe) = (3.2 ± 2.0) × 10−6 M⊙, where the uncertainty corresponds to one standard deviation (Fig. 3). We find that the value of MMC2 (60Fe) scales linearly with N1, the number of stars in MC1. From Fig. 2 and 3, it is clear that second generation MCs are expected to contain a significant amount of 60Fe due to contamination by supernovae of a first generation of stars. It remains to compare that amount of 60Fe contained in MC2 to its abundance in the ESS. Note that because the observed collapse timescales of cores (a few 105 yr, Onishi et al. 2002) are far shorter than the 60Fe half-life, there is no need for an extra decay term between the -- 6 -- molecular cloud stage (assumed to start 10 to 20 Myr after the onset of star formation in MC1) and the disk stage, implying that the abundance in protoplanetary disks is identical to that of MC2 from which they form. 2.3. Comparison with the solar system The initial abundance of 60Fe in the solar system is not precisely known (Mostefaoui et al. 2005; Gounelle & Meibom 2008). The most recent and precise studies failed to detect an isochron and placed upper limits of 6 × 10−7 and 1 × 10−7 respectively for the initial 60Fe/56Fe ratio (Dauphas et al. 2008; Regelous et al. 2008). Adopting a conservative initial ratio 60Fe/56Fe = 3 × 10−7, it is estimated that the molecular cloud progenitor of our solar system had an 60Fe concentration of [60Fe]SS = 4 × 10−10 M⊙ per unit of solar mass, assuming a 56Fe/1H ratio of 3.2 × 10−5, and a metallicity of 0.7 (Lodders 2003). Our model can account for the 60Fe abundance in the ESS provided it formed in a molecular cloud with a mass MMC2 = MMC2(60Fe)/[60Fe]SS = (0.80 ± 0.50) × 104 M⊙. Our typical case, N1 = 5000 stars, corresponds to the estimated number of stars in the UCL-LCC association which formed ∼ 12 Myr before the Upper Sco association (de Geus 1992), within the 10-20 Myr interval defined above. If we take 2350 M⊙ as the stellar content of Upper Sco (de Geus 1992) and a molecular gas mass of (0.8 ± 0.5) × 104 M⊙, we obtain a star formation efficiency of 29+50 −11 %. This star formation efficiency is in line with the observed star formation efficiencies (5-30 %) of nearby star-forming regions (Lada & Lada 2003), implying that if Upper Sco molecular gas was swept-up by the explosions of supernovae from the UCL-LCC association (Preibisch & Zinnecker 2007), it is expected to contain 60Fe at a concentration similar to that of the ESS. This indicates that our model is self-consistent and offers a plausible astrophysical setting for the presence of 60Fe within the nascent solar system. 3. Discussion It is obvious that given the stochastic nature of star formation and the variable formation timescales of molecular clouds, a range of 60Fe abundance is expected in molecular clouds (Fig. 2 and 3), and therefore in protoplanetary disks. Some of the input parameters such as f or η could be a factor of a few smaller than the adopted values, lowering accordingly MMC2 (60Fe). The 60Fe content of the ESS might however be a factor of 3 smaller than the one we adopted (Regelous et al. 2008). In addition, fSN could be significantly higher -- 7 -- (fSN = 3 × 10−3, Adams & Laughlin 2001) than the value we adopted (fSN = 2.1 × 10−3), resulting in a higher value of 60Fe in MC2 due to a larger number of SNe in MC1. Finally, any increase of the half-life of 60 Fe (Gunter Korschinek, personal communication) would increase (exponentially) MMC2 (60Fe). The point of the calculations above is to show that for typical numbers (60Fe yields, molecular clouds masses and formation timescales, star formation efficiency...), the estimated ESS abundance of 60Fe can be reproduced in the context of a reasonable astrophysical model. Our model differs from previous models on several important points. First, 60Fe is delivered to a molecular cloud by a diversity of supernovae rather than by a single supernova. Second, the mass of the receiving phase (∼ 104 M⊙) is orders of magnitude larger than for the disk and the core model (0.013 and 1 M⊙ respectively, e.g. Looney et al. 2006). Third, 60Fe is not injected into a dense phase (nH2 ∼ 105 cm−3 and ∼ 1014 cm−3 for the core and disk respectively) isolated from the rest of the ISM, but is delivered into a relatively diffuse ISM phase interacting with other ISM components leading to high mixing efficiency (see §2.2). Fourth, the 60Fe-producing supernovae belong to previous generations of massive stars rather than to the same generation of stars. Fifth, the supernovae shock waves do not trigger the collapse of a pre-existing molecular cloud core, but rather contribute to build on a timescale of 10-20 Myr a new, second generation MC. Finally, this new model takes quantitatively into account the stochasticity of the star-forming process (though this approach has been used in a different context, Cervino et al. 2000). The paradigm for MC formation described in §2.1 is however not universally accepted. An alternative or complementary view is that gravitational instabilities represent the main driver for MC formation rather than large scale convergent flows (e.g. Hennebelle et al. 2008). However, independently of the main driver for MC formation, it remains observa- tionally true that most stars form in Giant Molecular Clouds (GMCs). In such a context, MC1 and MC2 would represent two different regions of the same GMC which have evolved at different paces (e.g. Fellhauer & Kroupa 2005). The aforementioned observation that OB associations are made of subgroups of different ages can indeed be interpreted as an evidence for sequential star formation in GMCs (Elmegreen & Lada 1977), preserving the essence of our model. The strength of our model is that the 60Fe content in the ESS is easily reproduced within a plausible, if not common, astrophysical setting, unlike the disk or core models (§1), and that all solar systems (within MC2 or the younger region of a GMC) will receive 60Fe instead of a few disks or cores. Therefore, even if a significant fraction of stars (50 %) form in MC1, the overall probability for 60Fe MC inheritance is far higher than that of injection into a disk or a core. In that respect, though there is not one typical solar system, there is no need to -- 8 -- call for an unlikely astrophysical setting for our solar system formation. An important and so far unresolved problem associated with all models based on su- pernova ejecta, is the over production of 53Mn relatively to 60Fe and their inferred ratio in the early solar system (Wasserburg et al. 2006). In the current supernova 1D models, 53Mn and 60Fe are produced deep in the supernova interior, together with 56Ni (T1/2 = 6 days). Nickel-56 is known to make it to the surface (Arnett et al. 1989), ruling out fallback as a solution to the 53Mn overproduction problem (e.g. Meyer & Clayton 2000). Our model does not offer a solution to this general problem, but it is not inconceivable that a solution might come from new developments in supernovae nucleosynthetic models, many aspects of which are not fully understood (Woosley & Heger 2007; Magkotsios et al.2008). Though 26Al can be produced by energetic particles irradiation (Lee et al. 1998), it might be difficult for irradiation to account for the entire ESS inventory (Duprat & Tatischeff 2007; Fitoussi et al. 2008). The ESS 26Al/60Fe mass ratio was ∼ 8, adopting 60Fe/56Fe = 3 × 10−7, 26Al/27Al = 5 × 10−5 and an 27Al/56Fe ratio of 0.11 (Lodders 2003). Given the strong heterogeneity in the 26Al distribution (Diehl et al. 2006), this compares relatively well with the observed ISM 26Al/60Fe mass ratio of ∼ 3 (Wang et al. 2007). This suggests that together with 60Fe, a substantial amount of 26Al could also have been inherited from the progenitor molecular cloud. M. Serrano is thanked for producing Fig. 1. The reviewers significantly helped to improve the paper. This study was funded by the PNP and the France-´Etats-Unis fund from CNRS, the ANR and the european grant ORIGINS [MRTN-CT-2006-035519]. Adams, F. C., & Laughlin, G. 2001, Icarus, 150, 151 REFERENCES Arnett, W. D., Bahcall, J. N., Kirshner, R. P., & Woosley, S. E. 1989, ARA&A, 27, 629 Audit, E., & Hennebelle, P. 2005, A&A, 433, 1 Boss, A. P., Ipatov, S. I., Keiser, S. A., Myhill, E. A., & Vanhala, H. A. T. 2008, ApJ, 686, L119 Brasser, R., Duncan, M.J & Levison, H.F. 2006, Icarus, 184, 59 Cameron, A. G. W. & Truran, J. W. 1977, Icarus, 30, 447 -- 9 -- Cameron, A. G. W., Hoflich, P., Myers, P. C. & Clayton, D. D. 1995, ApJ, 447, L53 Cervino, M., Knodlseder, J., Schaerer, D., von Ballmoos, P., & Meynet, G. 2000, A&A, 363, 970 Dauphas, N., et al. 2008, ApJ, 686, 560 de Geus, E. J. 1992, A&A, 262, 258 Diehl, R., et al. 2006, Nature, 439, 45 Duprat, J., & Tatischeff, V. 2007, ApJ, 671, L69 Elmegreen, B.G. 2000, ApJ, 277 Elmegreen, B. G., & Lada, C. J. 1977, ApJ, 214, 725 Fellhauer, M., & Kroupa, P. 2005, ApJ, 630, 879 Fitoussi, C., et al. 2008, Phys. Rev. C, 78, 044613 Glover, S. C. O., & Mac Low, M.-M. 2007, ApJ, 659, 1317 Gounelle, M., & Meibom, A. 2008, ApJ, 680, 781 Hartmann, L., Ballesteros-Paredes, J. & Bergin, E. A. 2001, ApJ, 562, 852 Hennebelle, P., Banerjee, R., V´azquez-Semadeni, E., Klessen, R. S., & Audit, E. 2008, A&A, 486, L43 Hennebelle, P., Mac Low, M. -., & Vazquez-Semadeni, E. 2007, ArXiv e-prints, 711, arXiv:0711.2417 Kastner, J. H. & Myers, P. C. 1994, ApJ, 421, 605 Koyama, H. & Inutsuka, S. 2002, ApJ, 564, L97 Kroupa, P., Tout, C. A., & Gilmore, G. 1993, MNRAS, 262, 545 Lada, C. J. & Lada, E. A. 2003, ARA&A, 41, 57 Lee, T., Shu, F. H., Shang, H., Glassgold, A. E. & Rehm, K.E. 1998, ApJ, 506, 898 Limongi, M. & Chieffi, A. 2006, ApJ, 647, 483 Lodders, K. 2003, ApJ, 591, 1220 -- 10 -- Looney, L. W., Tobin, J. J., & Fields, B. D. 2006, ApJ, 652, 1755 Magkotsios, G., et al. 2008, arXiv:0811.4651 Meyer, B. S. & Clayton, D. D. 2000, SSR, 92, 133 Montmerle, T., Augereau, J. C., Chaussidon, M., Gounelle, M., Marty, B. & Morbidelli, A. 2006, EMP, 98, 39 Mostefaoui, S., Lugmair, G. W. & Hoppe, P. 2005, ApJ, 625, 271 Onishi, T., Mizuno, A., Kawamura, A., Tachihara, K., & Fukui, Y. 2002, ApJ, 575, 950 Ouellette, N., Desch, S. J., & Hester, J. J. 2007, ApJ, 662, 1268 Preibisch, T., & Zinnecker, H. 2007, IAU Symposium, 237, 270 Rauscher, T., Heger, A., Hoffman, R. D. & Woosley, S. E. 2002, ApJ, 576, 323 Regelous, M., Elliott, T., & Coath, C. D. 2008, EPSL, 272, 330 Russell, S. S., Gounelle, M. & Hutchison, R. 2001, Phil. Trans. Roy. Soc. A, 359, 1991 Schaller, G., Schaerer, D., Meynet, G., & Maeder, A. 1992, A&AS, 96, 269 Sicilia-Aguilar, A., Hartmann, L. W., Hern´andez, J., Briceno, C., & Calvet, N. 2005, AJ, 130, 188 Takigawa, A., Miki, J., Tachibana, S., Huss, G. R., Tominaga, N., Umeda, H., & Nomoto, K. 2008, ApJ, 688, 1382 V´azquez-Semadeni, E. et al. 2007, ApJ, 657, 870 V´azquez-Semadeni, E., Kim, J., Shadmehri, M., Ballesteros-Paredes, J. 2005, ApJ, 618, 344 Wang, W. et al. 2007, A&A, 469, 1005 Wasserburg, G. J., Busso, M., Gallino, R. & Nollett, K. M. 2006, Nuclear Physics A, 777, 5 Weidner, C., & Kroupa, P. 2006, MNRAS, 365, 1333 Williams, J.P. & Gaidos, E. 2007, ApJ, 663, L33 Woosley, S. E. & Weaver, T. A. 1995, ApJS, 101, 181 Woosley, S. E., & Heger, A. 2007, Phys. Rep., 442, 269 -- 11 -- This preprint was prepared with the AAS LATEX macros v5.2. -- 12 -- T = 0 Myr T = 3 Myr MC1 T = 5 Myr 60Fe SN SN1 60Fe OB1 T = 10-15 Myr 60Fe SN4 60Fe MC2 Proto high-mass star Proto low-mass star Disk-free low-mass star Windy high-mass star Fig. 1. -- Sketch of the SPACE model. At t = 0 Myr, star formation starts in MC1. At t = 3 Myr, high-mass stars emit powerful winds which dissipate the molecular gas and start to accumulate interstellar gas further away. At t = 5 Myr, the first supernova in OB1 explodes and mixes 60Fe in the swept-up gas. At t = 10-15 Myr, the swept-up gas becomes dense enough to become molecular and star formation starts in MC2. As a result of sequential molecular cloud formation, MC2 contains a relatively high abundance of 60Fe produced by the previous supernovae (4 in that particular case), which contributed to its formation. -- 13 -- N1 =5000 stars 5 10 Time (Myr) 15 20 10−2 10−3 10−4 10−5 10−6 10−7 10−8 0 ) s e s s a m l r a o s ( 2 C M n i e c n a d n u b a e F 0 6 Fig. 2. -- Time evolution of the 60Fe abundance in the MC2 molecular cloud. Each thin line represents one realization of the simulation, while the thick red line is the average of 102 realizations. -- 14 -- N1 =5000 stars Averaged over the time interval 10−20 Myr 20 40 60 Realization # 80 100 10−5 10−6 ) s e s s a m l r a o s ( 2 C M n i e c n a d n u b a e F 0 6 10−7 0 Fig. 3. -- Average of the 60Fe abundance in the MC2 molecular cloud over the time interval 10-20 Myr for different realizations of the model (100). -- 15 -- Table 1. Model input data M (M⊙) tSN(Myr) YSN(60Fe) (M⊙) 11 12 13 14 15 16 17 18 19 20 21 22 25 30 35 40 60 80 120 20.8 17.8 15.5 13.8 12.5 11.4 10.6 9.9 9.2 8.8 8.3 7.9 7.0 6.0 5.3 4.9 3.9 3.4 3.9 5.25E-6 3.62E-6 9.03E-5 5.72E-6 3.31E-5 4.39E-6 7.96E-6 2.54E-5 7.83E-5 2.09E-5 2.45E-5 5.19E-5 6.96E-5 3.75E-5 7.37E-5 5.93E-5 2.27E-4 7.55E-4 9.93E-4 Note. -- M is the stellar mass, tSN is the stellar lifetime and YSN(60Fe) the 60Fe yield. The stellar lifetime is calculated us- ing the formula log(tSN) = 1.4/(logM)1.5 (Schaller et al. 1992; Williams & Gai- dos 2007). Iron-60 yields are the aver- age of the yields modelized by Woosley & Weaver (1995), Rauscher et al. (2002) and Limongi & Chiffi (2006).
1210.1217
1
1210
2012-10-03T20:00:02
The first planet detected in the WTS: an inflated hot-Jupiter in a 3.35 day orbit around a late F-star
[ "astro-ph.EP", "astro-ph.SR" ]
We report the discovery of WTS-1b, the first extrasolar planet found by the WFCAM Transit Survey, which began observations at the 3.8-m United Kingdom Infrared Telescope. Light curves comprising almost 1200 epochs with a photometric precision of better than 1 per cent to J=16 were constructed for 60000 stars and searched for periodic transit signals. For one of the most promising transiting candidates, high-resolution spectra taken at the Hobby-Eberly Telescope allowed us to estimate the spectroscopic parameters of the host star, a late-F main sequence dwarf (V=16.13) with possibly slightly subsolar metallicity, and to measure its radial velocity variations. The combined analysis of the light curves and spectroscopic data resulted in an orbital period of the substellar companion of 3.35 days, a planetary mass of 4.01+-0.35 Mj and a planetary radius of 1.49+-0.17 Rj. WTS-1b has one of the largest radius anomalies among the known hot Jupiters in the mass range 3-5 Mj.
astro-ph.EP
astro-ph
Mon. Not. R. Astron. Soc. 000, 1 -- 15 (2012) Printed 14 August 2018 (MN LATEX style file v2.2) The first planet detected in the WTS: an inflated hot-Jupiter in a 3.35 day orbit around a late F-star⋆ M. Cappetta1†, R.P. Saglia1,2, J.L. Birkby3,6, J. Koppenhoefer1,2, D.J. Pinfield4, S.T. Hodgkin3, P. Cruz5, G. Kov´acs3, B. Sipocz4, D. Barrado5,16, B. Nefs6, Y.V. Pavlenko7, L. Fossati8, C. del Burgo9,10,11, E.L. Mart´ın12, I. Snellen6, J. Barnes4, A. M. Bayo18, D. A. Campbell4, S. Catalan4, M.C. G´alvez-Ortiz12, N. Goulding4, C. Haswell8, O. Ivanyuk7, H. Jones4, M. Kuznetsov7, N. Lodieu13, F. Marocco4, D. Mislis3, F. Murgas13,14, R. Napiwotzki4, E. Palle13,14, D. Pollacco15, L. Sarro Baro17, E. Solano5,19, P. Steele1, H. Stoev5, R. Tata13,14, J. Zendejas1,2 1Max-Planck-Institut fur extraterrestrische Physik, Giessenbachstrasse, D-85741 Garching, Germany 2Universitatssternwarte Scheinerstrasse 1, D-81679 Munchen, Germany 3Institute of Astronomy, University of Cambridge, Madingley Road, Cambridge, CB3 0HA, UK 4Center for Astrophysics Research, University of Hertfordshire, College Lane, Hatfield, Hertfordshire AL10 9AB, UK 5Departamento de Astrof´ısica, Centro de Astrobiolog´ıa (CSIC/INTA), PO Box 78, E-28691 Villanueva de la Canada, Spain 6Leiden Observatory, Lieiden University, Postbus 9513, 2300 RA, Leiden, The Netherlands 7Main Astronomical Observatory of Ukrainian Academy of Sciences, Golosiiv Woods, Kyiv-127, 03680, Ukraine 8Department of Physical Sciences, The Open University, Walton Hall, Milton Keynes, MK7 6AA, UK 9UNINOVA-CA3, Campus da Caparica, Quinta da Torre, Monte de Caparica 2825-149, Caparica, Portugal 10School of Cosmic Physics, Dublin Institute for Advanced Studies, Dublin 2, Ireland 11Instituto Nacional de Astrof´ısica, ´Optica y Electr´onica (INAOE), Aptdo. Postal 51 y 216, 72000 Puebla, Pue., Mexico 12Centro de Astrobiolog´ıa (CSIC-INTA). Crta, Ajalvir km 4. E-28850, Torrej´on de Ardoz, Madrid, Spain 13Instituto de Astrof´ısica de Canarias, Calle V´ıa L´actea s/n, E-38200 La Laguna, Tenerife, Spain 14Departamento de Astrof´ısica, Universidad de La Laguna (ULL), E-38205 La Laguna, Tenerife, Spain 15Astrophysics Research Centre, School of Mathematics & Physics, Queens University, University Road, Belfast BT7 1NN 16Calar Alto Observatory, Centro Astronmico Hispano Alemn, C/ Jess Durbn Remn, E-04004 Almera, Spain 17Departamento de Inteligencia Artificial, UNED, Juan del Rosal, 16, 28040 Madrid, Spain 18European Southern Observatory, Alonso de C´ordova 3107, Vitacura, Santiago, Chile 19Spanish Virtual Observatory Accepted XXXX Received XXXX ; in original form XXXX ABSTRACT We report the discovery of WTS-1b, the first extrasolar planet found by the WFCAM Transit Survey, which began observations at the 3.8-m United Kingdom Infrared Tele- scope. Light curves comprising almost 1200 epochs with a photometric precision of better than 1 per cent to J ∼ 16 were constructed for ∼ 60 000 stars and searched for periodic transit signals. For one of the most promising transiting candidates, high- resolution spectra taken at the Hobby-Eberly Telescope allowed us to estimate the spectroscopic parameters of the host star, a late-F main sequence dwarf (V=16.13) with possibly slightly subsolar metallicity, and to measure its radial velocity variations. The combined analysis of the light curves and spectroscopic data resulted in an orbital period of the substellar companion of 3.35 days, a planetary mass of 4.01±0.35 MJ , and a planetary radius of 1.49+0.16 −0.18 RJ . WTS-1b has one of the largest radius anomalies among the known hot Jupiters in the mass range 3-5 MJ . Key words: Extrasolar planet, Hot-Jupiter, Radius anomaly ⋆ Based on observations collected at the 3.8-m United Kingdom Infrared Telescope (Hawaii, USA), the Hobby-Eberly Telescope c(cid:13) 2012 RAS 2 M. Cappetta et al. 1 INTRODUCTION The existence of highly-irradiated, gas-giants planets orbit- ing within < 0.1 AU of their host stars, and the unexpected large radii of many of them, is an unresolved problem in the theory of planet formation and evolution (Baraffe et al. 2010). Their prominence amongst the 777 confirmed ex- oplanets1 is unsurprising; their large radii, large masses, and short orbital periods make them readily accessible to ground-based transit and radial velocity surveys, on account of the comparatively large flux variations and reflex motions that they cause. Exoplanet searches often focus on the detec- tion of small Earth-like exoplanets, but understanding the formation mechanism and evolution of the giant planets, particularly in the overlap mass regime with brown dwarfs, is a key question in astrophysics. This paper reports on WTS-1b, the first planet discov- ery from the WFCAM Transit Survey (WTS; Kov´acs et al. 2012; Birkby et al. 2011). The WTS is the only large-scale ground-based transit survey that operates at near infrared (NIR) wavelengths. The advantage of photometric monitor- ing at NIR wavelengths is an increased sensitivity to photons from M-dwarfs (M⋆ < 0.6 M⊙). These small stars undergo similar flux variations and reflex motions in the presence of an Earth-like companion as solar-type stars do with hot Jupiter companions. While a primary goal of the WTS is the detection of terrestrial planets around cool stars, WTS can also provide observational constraints on the mechanism for giant planet formation, by accurately measuring the hot Jupiter fraction for M-dwarfs. The light curve quality of the WTS is sufficient to reveal transiting super-Earths around mid-M dwarfs (Kov´acs et al. 2012), but scaling-up means that any hot Jupiter companions to the ∼80 000 FGK stars in the survey are also detectable. Theoretical models of isolated giant planets predict an almost constant radius for pure H+He objects in the mass range 0.5 − 10 MJ as a result of the equilibrium between the electron degeneracy in the core and the pressure sup- port in the external gas layers (Zapolsky & Salpeter 1969; Guillot 2005; Seager et al. 2007; Baraffe et al. 2010). Hence the larger radii of many hot Jupiters (hereafter HJ) must arise from other factors. Due to the proximity of these plan- ets to their host star, the irradiation of the surface of the planet is thought to play a major role in the so-called ra- dius anomaly, by altering the thermal equilibrium and de- laying the Kelvin-Helmholtz contraction of the planet from birth (e.g. Showman & Guillot 2002). This is supported by a correlation between the mean planetary density and the incident stellar flux (Laughlin et al. 2010; Demory & Seager 2011; Enoch et al. 2012). However, it has been shown that this cannot be the only explanation for the radius anomaly (Burrows et al. 2007), and other sources must contribute to (Texas, USA), the 2.5-m Isaac Newton Telescope (La Palma, Spain), the William Herschel Telescope (La Palma, Spain), the German-Spanish Astronomical Center (Calar Alto, Spain), the Kitt Peak National Observatory (Arizona, USA) and the Hert- fordshire's Bayfordbury Observatory. † E-mail: [email protected] 1 http://www.exoplanet.eu at the time of the publication of this work the large amount of energy required to keep gas giants radii above ∼1.2 RJ (Baraffe et al. 2010). There are a number of physical mechanisms thought to be responsible for radius inflation, including (but not lim- ited to): tidal heating due to the circularisation of close-in orbits (Bodenheimer et al. 2001, 2003; Jackson et al. 2008), reduced heat loss due to enhanced opacities in the outer layers of the planetary atmosphere (Burrows et al. 2007), double-diffusive convection leading to slower heat trans- portation (Chabrier et al. 2007; Leconte & Chabrier 2012), increased heating via Ohmic dissipation in which ionised atoms interact with the planetary magnetic field as they move along strong atmospheric winds (Batygin & Stevenson 2010), and a slower cooling rate due to the mechanical greenhouse effect in which turbulent mixing drives a down- wards flux of heat simulating a more intense incident irra- diation (Youdin & Mitchell 2010). A radically different ex- planation has been proposed by Mart´ın et al. (2011) who point out a correlation between radius anomaly and tidal decay timescale and suggest that inflated HJs are actually young because they have recently formed as a result of bi- nary mergers. Studying the radius anomaly in higher mass HJs (> 3 RJ ) is useful as they are perhaps more resilient to atmospheric loss due to their larger Roche lobe radius. This paper is organised as follows: in Section 1.1 we briefly describe the strategy of the WFCAM Transit Sur- vey, while the photometric and spectroscopic observations are presented in Section 2.1 along with their related data reduction. Section 3 describes how we characterized the host star (Section 3.1) and determined the properties of the planet (Section 3.2), using a combination of low- and high- resolution spectra and photometric follow-up observations. A discussion on the nature and the peculiarity of this new planet, our conclusions and further considerations are given in Section 4. Throughout this paper we refer to the planet as WTS-1b, and to the parent star and the whole star-planet system as WTS-1. 1.1 The WFCAM Transit Survey The WTS is an on-going photometric monitoring cam- paign using the Wide Field Camera (WFCAM) on the United Kingdom Infrared Telescope (UKIRT) at Mauna Kea, Hawaii, and has been in operation since August 2007. UKIRT is a 3.8-m telescope, designed solely for NIR obser- vations and operated in queue-scheduled mode. The survey was awarded 200 nights of observing time, of which ∼50% has been observed to-date, and runs primarily as a back-up program when observing conditions are not optimal for the main UKIDSS programs (seeing > 1 arcsec). The survey targets four 1.6 deg2 fields, distributed seasonally in right ascension at 03, 07, 17 and 19 hours to allow year-round visibility. WTS-1b is located in the 19 hour field (hereafter 19hr). The fields are close to the Galactic plane but have b > 5 degrees to avoid over-crowding and high reddening. The exact locations were chosen to maximise the number of M-dwarfs, while keeping giant contamination at a minimum and maintaining an E(B − V ) < 0.1. Due to the back-up nature of the program, observations of a given field are ran- domly distributed throughout a given night, but on average occur in a one hour block at its beginning or end. Seasonal visibility also leaves long gaps when no observations are c(cid:13) 2012 RAS, MNRAS 000, 1 -- 15 Table 1. WFCAM J-band light curve. The full epoch list, which contains 1 182 entries, will be available electronically. Table 2. The host star WTS-1. Inflated HJ in close orbit around a late F-star 3 HJD -2 400 000 54317.8138166 54317.8258565 ... 55896.7105702 Normalized Error flux 1.0034 0.9971 ... 0.0057 0.0056 ... 0.9979 0.0064 possible (Kov´acs et al. 2012). Since the WTS was primar- ily designed to find planets transiting M-dwarf stars, the observations are obtained in the J-band (∼1.25 µm). This wavelength is near to the peak of the spectral energy dis- tribution (SED) of a typical M-dwarf. Hotter stars, with an emission peaking at shorter wavelength, are consequently fainter in this band. Interestingly, photometric monitoring at NIR wavelengths may have a further advantage as it is potentially less susceptible to the effect of star-spot induced variability (Goulding et al. 2012). The discovery reported in this paper demonstrates that despite the survey optimisa- tion for M-dwarfs, its uneven epoch distribution and the increased difficulty in obtaining high-precision light curves from ground-based infrared detectors, it is still able to detect HJs around FGK stars. 2 OBSERVATIONS 2.1 Photometric data 2.1.1 UKIRT/WFCAM J-band photometry The instrument used for this campaign is WFCAM (Casali et al. 2007), which has a mosaic of four Rockwell Hawaii-II PACE infrared imaging 2048 × 2048 pixels detec- tors, covering 13.65′ × 13.65′ (0.4′′/pixel) each. The detec- tors are placed in the four corners of a square (this pattern is called a paw-print) with a separation of 12.83′ between the chips, corresponding to 94% of a chip width. Each of the four fields consists of eight pointings of the WFCAM paw-print, each one comprising a 9-point jitter pattern of 10 second exposures each, and tiled to give uniform cover- age across the field. It takes 15 minutes to observe an entire WTS field (9 × 10s × 8 + overheads). In this way, the NIR light-curves have an average cadence of 4 data points per hour. A full description of our 2-D image processing and light curve generation is given by Kov´acs et al. (2012) and closely follows that of Irwin et al. (2007). Briefly, a mod- ified version of the CASU INT wide-field survey pipeline (Irwin & Lewis 2001) 2 is used to remove the dark cur- rent and reset anomaly from the raw images, apply a flat- field correction using twilight flats, and to decurtain and sky subtract the final images. Astrometric and photometric calibration is performed using 2MASS stars in the field-of- view (Hodgkin et al. 2009). A master catalogue of source positions is generated from a stacked image of the 20 best 2 http://casu.ast.cam.ac.uk/surveys-projects/wfcam/technical/ c(cid:13) 2012 RAS, MNRAS 000, 1 -- 15 Parameter Value RAa Deca la ba µαcosδb b µδ 19h 35m 58.37s +36d 17m 25.17s +70.0140 deg +7.5486 deg -7.7±2.4 mas yr−1 -2.8±2.4 mas yr−1 a Epoch J2000; b Proper motion from SDSS. frames and co-located, variable aperture photometry is used to extract the light curves. We attempt to remove system- atic trends that may arise from flat-fielding inaccuracies or varying differential atmospheric extinction across the image by fitting a 2-D quadratic polynomial to the flux residuals in each light curve as a function of the source's spatial po- sition on the chip. As a final correction, for each source we remove residual seeing-correlated effects by fitting a second- order polynomial to the correlation between its flux and the stellar image FWHM per frame (see Irwin et al. 2007 for a discussion of the effectiveness of these techniques). The final J-band light curves have a photometric preci- sion of 1% down to J = 16 mag (∼7% for J = 18 mag while the uncertainty per data point is just 3 mmag for the bright- est stars (saturation occurs between J ∼ 12 − 13 mag). The unfolded J-band light curve for WTS-1 is shown in Figure 1 (all the measurements are reported in Table 1) and its final out-of-transit RMS is 0.0064 (equivalent to 6.92 mmag). 2.1.2 Transit detection algorithm We identified WTS-1b in the J-band light curves by us- ing the Box-Least-Squares transit search algorithm occfit, as described in Aigrain & Irwin (2004), which takes a maxi- mum likelihood approach to fitting generalized periodic step functions. Before inspecting transit candidates by eye, we employed several criteria to speed up the detection process. The first is a magnitude cut, in which we removed all sources fainter than J = 17 mag. We also required that the source have an image morphology consistent with stellar sources (Irwin et al. 2007). Despite our attempts to remove system- atic trends in the light curves, we invariably suffered from residual correlated red noise, so we modified the detection significance statistic, S, from occfit according to the pre- scription (equation 4) of Pont et al. (2006), to obtain Sred. Our transit candidates must then have Sred > 5 to sur- vive, although we note that this is more permissive with respect to the limit recommended by Pont and collabora- tors (Sred > 7). We further discarded any detections with a period in the range 0.99 <P< 1.005 days, in order to avoid the common ∼1 day alias of the ground-based photometric surveys. Next, as fully described in Birkby et al. (2012a), the WFCAM ZY JHK single epoch photometry was combined with five more optical photometric data points (ugriz bands) available for the 19hr field from the Sloan Digi- 4 M. Cappetta et al. Figure 1. The unfolded J-band light curve for WTS-1. tized Sky Survey archive (SDSS 7th release, Nash 1996) to create an initial SED. The SEDs were fitted with the NextGen models (Baraffe et al. 1998) in order to estimate effective temperatures and hence a stellar radius for each source. We could then impose a final threshold on the de- tected transit depths, by rejecting those that corresponded to a planetary radius greater than 2 RJ . It is worth noting that occfit tends to under-estimate transit depths because it does not consider limb-darkening effects and the trape- zoidal shape of the transits. Moreover, the NextGen models systematically under-estimate the temperature of solar-like stars (Baraffe et al. 1998), so the initial radius estimates are also under-predicted. Hence, the genuine HJ transit events are unlikely to be removed in this final threshold cut. As a result, our transits detection procedure was conservative. Many of the ∼3 500 phase-folded light curves which satis- fied our criteria were false-positives arising from nights of bad data or single bad frames. Others were binary systems, folded on half the true orbital period. A more detailed anal- ysis of the candidate selection procedure and of advanced selection steps in the survey can be found in Sipocz et al. (2012). The WTS-1 J-band light curve passed all our selection criteria and the object (see Table 2) progressed to the follow- ing phases of candidate's confirmation. The descriptions of the photometric and spectroscopic data in the next sections are organized in order to match the following chronologi- cal sequence of analysis. First, optical i′-band photometric follow-up of WTS-1 was conduced in order to prove the real presence of the transit and check the transit depth consis- tency between the two bands. Then, ISIS/WHT intermedi- ate resolution spectra enabled us to place strong constrains on the velocity amplitude of host star (at the km s−1level) and therefore to rule out the false-positive eclipsing binaries scenarios. Finally we moved to the high-resolution spectro- scopic follow-up. The confirmation of the planetary nature of WTS-1b came with the RV measurements obtained with the HET spectra. 2.1.3 Broad band photometry The WFCAM and SDSS photometric data for WTS-1 are reported in Table 3 with other single epoch broad band pho- tometric observations. Johnson B V R bands were observed for WTS-1 on the night of 6th April 2012 at the Univer- sity of Hertfordshire's Bayfordbury Observatory. We used a Meade LX200GPS 16-inch f/10 telescope fitted with an Table 3. Broad band photometric data of WTS-1 measured within the WFCAM (Vega), SDSS (AB), 2MASS (Vega) and WISE (Vega) surveys. Johnson magnitudes in the visible are pro- vided too (Vega). Effective wavelength λef f (mean wavelength weighted by the transmission function of the filter), equivalent width (EW) and magnitude are given for each single pass-band. The bands are sorted by increasing λef f . Band λef f [A] EW[A] Magnitude SDSS-u Johnson-B SDSS-g Johnson-V SDSS-r Johnson-R SDSS-i UKIDSS-Z SDSS-z UKIDSS-Y 2MASS-J UKIDSS-J UKIDSS-H 2MASS-H 2MASS-Ks UKIDSS-K WISE-W1 WISE-W2 3546 4378 4670 5466 6156 6696 7471 8817 8918 10305 12350 12483 16313 16620 21590 22010 34002 46520 558 970 1158 890 1111 2070 1045 879 1124 1007 1624 1474 2779 2509 2619 3267 6626 10422 18.007 (±0.014) 17.0 (±0.1) 16.785 (±0.004) 16.5 (±0.1) 16.434 (±0.004) 16.1 (±0.1) 16.249 (±0.004) 15.742 (±0.005) 16.189 (±0.008) 15.642 (±0.007) 15.375 (±0.052) 15.387 (±0.005) 15.103 (±0.006) 15.187 (±0.081) 15.271 (±0.199) 15.048 (±0.009) 15.041 (±0.044) 15.886 (±0.157) SBIG STL-6303E CCD camera, and integration times of 300 seconds per band. Images were bias, dark, and flat-field cor- rected, and extracted aperture photometry was calibrated using three bright reference stars within the image. Photo- metric uncertainties combine contributions from the SNR of the source (typically ∼20) with the scatter in the zero point from the calibration stars. The Two Micron All Sky Survey (2MASS, Skrutskie et al. 2006) and the Wide-field Infrared Survey Explorer (WISE, Wright et al. 2010) provide further NIR data points (J H Ks bands and W 1 W 2 bands respec- tively). 2.1.4 INT i′-band data In addition to the WFCAM J-band light curve, we observed one half transit of the WTS-1 system in the i′-band using the Wide Field Camera on the 2.5-m Isaac Newton Tele- scope (McMahon et al. 2001) on July 23, 2010. A total of c(cid:13) 2012 RAS, MNRAS 000, 1 -- 15 Table 4. INT i′-band light curve of WTS-1. The full epoch list, which contains 82 entries, will be available electronically. HJD -2 400 000 55401.3703254 55401.3809041 ... 55401.4894461 Normalized Error flux 1.0244 1.0205 ... 0.0109 0.0034 ... 0.9936 0.0022 82 images, sampling the ingress of the transit, were taken with an integration time of 60 seconds. The data were re- duced with the CASU INT/WFC data reduction pipeline as described in detail by Irwin & Lewis (2001) and Irwin et al. (2007). The pipeline performs a standard CCD reduction, including bias correction, trimming of the overscan and non- illuminated regions, a non-linearity correction, flat-fielding and defringing, followed by astrometric and photometric cal- ibration. A master catalogue for the i′-band filter was then generated by stacking 20 frames taken under the best con- ditions (seeing, sky brightness and transparency) and run- ning source detection software on the stacked image. The extracted source positions were used to perform variable aperture photometry on all of the images, resulting in a time-series of differential photometry. The final out-of-transit RMS in the WTS-1 i′-band light curve is 0.0026 (equivalent to 2.87 mmag) and is used to refine the transit model fitting procedure in Section 3.2.1. The i′-band light curve for WTS-1 is given in Table 4. 2.2 Spectroscopic data 2.2.1 ISIS/WHT We carried out intermediate-resolution spectroscopy of the star WTS-1 over two nights between July 29 − 30, 2010, as part of a wider follow-up campaign of the WTS planet candidates, using the William Herschel Telescope (WHT) at Roque de Los Muchachos, La Palma. We used the single-slit Intermediate dispersion Spectrograph and Imaging System (ISIS). The red arm with the R1200R grating centred on 8500 A was employed. We did not use the dichroic during the ISIS observations because it can induce systematics and up to 10% efficiency losses in the red arm, which we wanted to avoid given the relative faintness of our targets. The four spectra observed have a wavelength coverage of 8100 -- 8900 A. The wavelength range was chosen to be optimal for the majority of the targets for our spectroscopic observa- tion which were low-mass stars. The slit width was chosen to match the approximate seeing at the time of observation giv- ing an average spectral resolution R ∼ 9000. An additional low-resolution spectrum was taken on July 16th, 2010, us- ing the ISIS spectrograph with the R158R grating centred on 6500 A. This spectrum has a resolution (R ∼ 1000), a SNR of ∼ 40 and a wider wavelength coverage (5000 -- 9000 A). The 3 package spectra were processed using the iraf.ccdproc 3 iraf is distributed by National Astronomy Observatories, which is operated by the Association of Universities of Research in Astronomy, Inc., under contract to the National Science, USA c(cid:13) 2012 RAS, MNRAS 000, 1 -- 15 Inflated HJ in close orbit around a late F-star 5 for instrumental signature removal. We optimally extracted the spectra and performed wavelength calibration using the semi-automatic kpno.doslit package. The dispersion func- tion employed in the wavelength calibration was performed using CuNe arc lamp spectra taken after each set of expo- sures. 2.2.2 CAFOS/2.2-m Calar Alto Two spectra of WTS-1 were obtained with CAFOS at the 2.2-m telescope at the Calar Alto observatory (as a Direc- tors Discretionary Time - DDT) in June, 2011. CAFOS is a 2k×2k CCD SITe#1d camera at the RC focus, and it was equipped with the grism R-100 that gives a dispersion of ∼ 2.0 A/pix and a wavelength coverage from 5850 to 9500 A, approximately. Their resolving power is of around R ∼ 1900 at 7500 A, with a SN R ∼ 25. The data reduction was per- formed following a standard procedure for CCD processing and spectra extraction with iraf. The spectra were finally averaged in order to increase the SNR. 2.2.3 KPNO A low resolution spectrum of WTS-1 was observed in September 2011 with the Ritchey-Chretien Focus Spectro- graph at the 4-m telescope at Kitt Peak (Arizona, USA). The grism BL-181, which gives a dispersion of ∼ 2.8 A/pix, was used. Calibration, sky subtraction, wavelength and flux calibration were performed following a standard procedure for long slit observations using dedicated iraf tasks. The ThAr arc lamp and the standard star spectra, employed for the wavelength and the flux calibration respectively, were taken directly after the science exposure. The mea- sured spectrum covers the wavelength range 6000-9000 A with SNR∼40 at 7500 A and has a resolution of R ∼ 1000. This is the only flux calibrated spectrum we have available. 2.2.4 HET In the late 2010 and in the second half of the 2011 the star WTS-1 was observed during 11 nights with the High Res- olution Spectrograph (HRS) housed in the insulated cham- ber in the basement of the 9.2-m segmented mirror Hobby- Eberly Telescope (HET, see Ramsey et al. 1998). The HRS is a single channel adaptation of the ESO UVES spectrome- ter linked to the corrected prime focus of the HET through its fiber-fed instrument as described by Tull (1998). It uses an R-4 echelle mosaic, which we used with a resolution of R=60 000, and a cross-dispersion grating to separate spec- tral orders, while the detector is a mosaic of two thinned and anti-reflection coated 2K × 4K CCDs. One science fiber was used to get the spectrum of the target star and two sky fibers were used in order to subtract the sky contamination. A couple of ThAr calibration expo- sures were taken immediately before and after the science exposure. This strategy allowed us to keep under control any undesired systematic effect during the observation. Each sci- ence observation (except one, see Section 3.2.2) was split in two exposures, of about half an hour each, in order to limit the effect of the cosmic rays hits, which can affect the data 6 M. Cappetta et al. The data performed with (Willmarth & Barnes reduction was package reduction and analysis steps. Due to the faintness of the star, the Iodine gas cell was not used for our observations. the 1994). iraf.echelle After the standard calibration of the raw science frame, bias-subtraction and flat-fielding, the stellar spectra were extracted order by order and the related sky spectrum was subtracted. The extracted ThAr spectra were used to compute the dispersion functions, which are characterized by an RMS of the order of 0.003 A. Consistency between the two solutions (computed from the ThAr taken before and after the science exposure) was checked for all the nights in order to detect possible drifts or any other technical hitch that could take place during each run. Successively, the spectra were wavelength calibrated using a linear interpolation of the two dispersion functions. In the subsequent data analysis, custom Matlab pro- grams were used. After the continuum estimation and the following normalization, the spectra were filtered to remove the residual cosmic rays peaks left after the previous filtering performed on the raw science frame with the iraf task cos- micrays. Comparing the spectra of all the nights for each single order, the pixels with higher flux, due to a cosmic hit on the detector, were detected and masked. The telluric ab- sorption lines present in the redder orders were then removed in order to avoid contaminations using a high SNR observed spectrum of a white dwarf as template for the telluric lines. Finally, the spectra related to the two split science expo- sures were averaged to obtain the final set of spectra used in the following analysis. A total of 40 orders (18 from the red CCD and 22 from the blue one) cover the wavelength range 4400-6300 A. The spectra have a SNR ∼ 8. 3 ANALYSIS AND RESULTS 3.1 Stellar parameters 3.1.1 Spectral energy distribution A first characterization of the parent star can be performed comparing the shape of the Spectral Energy Distribution (SED), constructed from broad band photometric observa- tions, with a grid of synthetic theoretical spectra. The data relative to the photometric bands, collected in Table 3, were analysed with the application VOSA (Virtual Observatory SED Analyser, Bayo et al. 2008, 2012). VOSA offers a valu- able set of tools for the SED analysis, allowing the estima- tion of the stellar parameters. It can be accessed through its web-page interface and accepts as input file an ASCII table with the following data: source identifier, coordinates of the source, distance to the source in parsecs, visual extinction (AV ), filter label, observed flux or magnitude and the related uncertainty. For our purpose, we tried to estimate only the main intrinsic parameters of the star: effective temperature, surface gravity and metallicity. The extinction AV was as- sumed as a further free parameter. The synthetic photometry is calculated by convolving the response curve of the used filter set with the theoreti- cal synthetic spectra. Then a statistical test is performed, via χ2 ν minimization, to estimate which set of synthetic pho- tometry best reproduces the observed data. We decided to employ the Kurucz ATLAS9 templates set (Castelli et al. Figure 2. Broad band photometric data of WTS-1: SDSS (filled dots), Johnson (empty dots), WFCAM (squares), 2MASS (trian- gles) and WISE (stars). The best fitting template (black line) is the ATLAS9 Kurucz Teff =6500 K, log g=4.5, [Fe/H]=-0.5 model with AV =0.44. Vertical errorbars correspond to the flux uncer- tainties while those along the X-axis represent each band's EW (see Table 3). 1997) to fit our photometric data. These templates repro- duce the SEDs in the high temperature regime better than the NextGen models (Baraffe et al. 1998), more suitable at lower temperatures (< 4500 K). In order to speed up the fitting procedure, we restricted the range of Teff and log g to 3500-8000 K and 3.0-5.0, re- spectively. These constrains in the parameter space did not affect the final results as it was checked a posteriori that the same results were obtained considering the full avail- able range for both parameters. The resulting best fitting synthetic template, plotted in Figure 2 with the photo- metric data, corresponds to the Teff =6500 K, log g=4.5 and [Fe/H]=-0.5 Kurucz model and AV =0.44. Uncertainties on the parameters were estimated both using χ2 ν statistical analysis and a bayesian (flat prior) approach. The related errors result to be of the order of 250 K, 0.2, 0.5 and 0.07 for Teff , log g, [Fe/H] and AV , respectively. As it can be seen in Figure 2, the WISE W 2 data point is not consistent with the best fitting model. Firstly, it is worth noting that the observed value of W 2 = 15.886 (±0.157) is below the 5 sigma point source sensitivity expected in the W 2-band (> 15.5). The number of single source detections used for the W 2-band measurement is also considerably less than that of the W 1-band (4 and 19 respectively) increasing the uncertainty in the measurement. Finally, the poor an- gular resolution of WISE in the W 2-band (6.4′′) could add further imprecisions to the final measured flux, especially in a field as crowded as the WTS 19hrs field. For these reasons, the WISE W 2 data point was not considered in the fitting procedure. Once the magnitude values were corrected for the in- terstellar absorption according to the best fitting value (AV =0.44±0.07), colour -- temperature relations were used to further check the effective temperature and spectral type of the host star. From the SDSS g and r magnitudes, we c(cid:13) 2012 RAS, MNRAS 000, 1 -- 15 Inflated HJ in close orbit around a late F-star 7 Figure 3. U pper row: comparison of a degraded spectrum of the Sun (black) with a synthetic spectrum (red) computed with Teff =5750 K, log g=4.5 and [Fe/H]=0.0. Lower row: comparison of the WTS-1 spectrum (black) with different synthetic spectra (colours). The comparison between the observed WTS-1 spectrum and the synthetic models took into account the differences of the core of the lines shown in the upper row plots. From this analysis, the best fitting model is the one with Teff =6250 K, log g=4.5 and [Fe/H]=0.0. obtained a value of (B-V)0=0.43±0.04 (Jester et al. 2005) which imply Teff =6300±600 K assuming log g=4.4 and [Fe/H]=-0.5 (Sekiguchi & Fukugita 2000). Following the ap- pendix B of Collier Cameron et al. (2007), the 2MASS (J − H) index of 0.23±0.09 leads to a value of Teff =6200±400 K while Table 3 of Covey et al. (2007) suggests the host star to be a late-F considering different colour indices at once. These results are all compatible within the uncertainties. 3.1.2 Spectroscopic analysis The spectroscopic spectral type determination was done firstly by comparing the spectrum observed with CAFOS with a set of spectra of template stars. Stars of different spectral types, uniformly spanning the F5 to G2 range, were observed with the same instrumental setting. Since the ob- served spectrum has a relatively low SNR, we focused the analysis on the strongest features present which are the Hα (6562.8 A) and the Ca ii triplet (8498.02 A, 8542.09 A, 8662.14 A). The best match was obtained, via minimization of the RMS of the difference between the WTS-1 and tem- plate star spectra, with the spectrum of an F6V star with solar metallicity. Afterwards, we tried to estimate the stellar parame- ters comparing the observed spectrum with a simulated spectrum with known parameters. For that aim, we used a library of high resolution synthetic stellar spectra by Coelho et al. (2005), created by the PFANT code (Barbuy c(cid:13) 2012 RAS, MNRAS 000, 1 -- 15 1982; Cayrel et al. 1991; Barbuy et al. 2003) that computes a synthetic spectrum assuming local thermodynamic equilib- rium (LTE). The synthetic spectra were achieved using the model atmospheres presented by Castelli & Kurucz (2003). Since the core of these lines are strongly affected by cu- mulative effects of the chromosphere, non-LTE (local ther- modynamical equilibrium) and inhomogeneity of velocity fields, we firstly compared a spectrum of the Sun observed with HIRES spectrograph at the Keck telescope (Vogt et al. 1994). The spectrum of the Sun was degraded to lower reso- lution and resampled to match our CAFOS spectrum speci- fications to see how such effects appear at this resolution and how different the solar spectrum is from a synthetic spec- trum from the library by Coelho and collaborators. Look- ing at the upper plots in Figure 3, we concluded that the cores of the lines in the simulated spectrum are systemati- cally higher than those in the observed spectrum of the Sun. Nevertheless, the Ca ii triplet line at 8498.02 A seems to be less affected by the above-mentioned problems. Consider- ing these differences between central depth of the observed and simulated spectra in the best fitting procedure, we esti- mated that the synthetic model with Teff =6250 K, log g=4.5 and [Fe/H]=0.0 best reproduces the observed spectrum of WTS-1. The expected uncertainties are of the order of the step size of the used library (δTeff=250 K, δlog g=0.5 and δ[Fe/H]=0.5). The high resolution HET spectra were employed to at- tempt a more detailed spectroscopic analysis of the host 8 M. Cappetta et al. x u l F . m r o N 1 0.9 0.8 0.7 0.6 HET spectrum ISIS spectrum Teff=6000; logg=4.4 Teff=6250; logg=4.4 Teff=6500; logg=4.4 6540 6570 Wavelength [Ao ] 6600 Figure 4. Intermediate resolution ISIS spectrum (black line) and high resolution HET spectrum (grey line) of the Hα line of WTS- 1 in comparison with synthetic spectra calculated with different effective temperatures of 6000 K (blue dash-dotted line), 6250 K (dashed line; finally adopted Teff ), and 6500 K (blue dotted line). The synthetic spectra were convolved in order to match the res- olution of the ISIS spectrum. star. The spectra observed each single night were stacked together obtaining a final spectrum with a SNR of about 12, calculated over a 1 A region at 5000 A. To compute model atmospheres, LLmodels stellar model atmosphere code (Shulyak et al. 2004) was used. For all the calcula- tions, LTE and plane-parallel geometry were assumed. We used the VALD database (Piskunov et al. 1995; Kupka et al. 1999; Ryabchikova et al. 1999) as a source of atomic line parameters for opacity calculations with the LLmodels code. Finally, convection was implemented according to the Canuto & Mazzitelli (1991) model of convection. The parameter determination and abundance analysis were performed iteratively, self-consistently recalculating a new model atmosphere any time one of the parameters, in- cluding the abundances, changed. As a starting point, we adopted the parameters derived from the CAFOS spectrum. We performed the atmospheric parameter determination by imposing the iron excitation and ionization equilibria mak- ing use of equivalent widths measured for all available un- blended and weakly blended lines. We converted the equiva- lent width of each line into an abundance value with a mod- ified version (Tsymbal 1996) of the WIDTH9 code (Kurucz 1993). Unfortunately, the faintness of the observed star, cou- pled with the calibration process (including the sky subtrac- tion), led to a distortion of the stronger lines, weakening their cores. For this reason, our analysis took into account only the measurable weak lines, making therefore impossi- ble a determination of the microturbulence velocity (vmic), which we fixed at a value of 0.85 km s−1 (Valenti & Fisher 2005). With the fixed vmic, we imposed the Fe excitation and ionization equilibria, which led us to an effective tempera- ture of 6000±400 K and a surface gravity of 4.3±0.4. Im- posing the ionization equilibrium we took into account the expected non-LTE effects for Fe i (∼0.05 dex, Mashonkina 2011), while for Fe ii non-LTE effects are negligible. In this temperature regime, the ionization equilibrium is sensitive to both Teff and log g variations, therefore it is important to simultaneously and independently further con- strain temperature and/or gravity. For this reason we looked at the Hα line to further constrain Teff, as in this temper- ature regime Hα is sensitive primarily to temperature vari- ations (Fuhrmann et al. 1993). We did this by fitting syn- thetic spectra, calculated with SYNTH3 (Kochukhov 2007), to the Hα line profile observed in the HET high resolution and in the ISIS/WHT low resolution spectra. As the Hα line of the HET spectrum was also affected by the before men- tioned reduction problems, only the wings of the line were considered. Although we could calculate synthetic line pro- files of Hα on the basis of atmospheric models with any Teff , because of the low SNR of our spectra, we performed the line profile fitting on the basis of models with a temperature step of 100 K (Fuhrmann et al. 1993, small variations in gravity and metallicity are negligible). From the fit of the Hα line we obtained a best fitting Teff of 6100±400 K. Further details on method, codes and techniques can be found in Fossati et al. (2009), Ryabchikova et al. (2009), Fossati et al. (2011) and references therein. Figure 4 shows the Hα line profile ob- served with ISIS and HET in comparison with synthetic spectra calculated assuming a reduced set of stellar parame- ters. On the basis of the previous analysis, we finally adopted Teff=6250±250 K, log g=4.4±0.1. With this set of param- eters and the equivalent widths measured with the HET spectrum, a final metallicity of -0.5±0.5 dex dex was derived, where the uncertainty takes into account the internal scatter and the uncertainty on the atmospheric parameters. By fit- ting synthetic spectra, calculated with the final atmospheric parameters and abundances, to the observation, we derived a projected rotational velocity v sin(i)=7±2 km s−1. The HET spectrum allowed us to measured the atmo- spheric lithium abundance from the Li i line at ∼6707 A. Lithium abundance is important as it can constrain the age of the star (see Section 3.1.3). As this line presents a strong hyperfine structure and is slightly blended by a nearby Fe i line, we measured the Li i abundance by means of spectral synthesis, instead of equivalent widths. By adopt- ing the meteoritic/terrestrial isotopic ratio Li6/Li7=0.08 by Rosman & Taylor (1998), we derived log N(Li)=2.5±0.4 (corresponding to an equivalent width of 41.12±24.40 mA), where the uncertainty takes into account the uncertainty on the atmospheric parameters, Teff in particular (see Figure 5). The low resolution spectrum obtained at the KPNO observatory, being the only flux calibrated spectrum, was compared to the Kurucz ATLAS9 templates set (the same employed in the SED analysis in Section 3.1.1). The lim- ited wavelength range covered by the observed spectrum (3000 A), allowed to achieve usable results fitting only one parameter of the template spectra. We therefore decided to leave Teff as a free parameter, fixing the other quantities to log g=4.5, [Fe/H]=-0.5 and AV =0.44. We concluded that the temperature range 6000-6500 K brackets the Teff with 95% confidence level. 3.1.3 Properties of the host star WTS-1 The atmospheric parameters of the star WTS-1 were com- puted combining the results coming from the analysis de- scribed in the previous sections. We finally obtained an effec- tive temperature of 6250±200 K, a surface gravity of 4.4±0.1 c(cid:13) 2012 RAS, MNRAS 000, 1 -- 15 Inflated HJ in close orbit around a late F-star 9 Table 5. Properties of the WTS-1 host star. Parameter Value Teff log g 6250±200 K 4.4±0.1 [Fe/H] [-0.5, 0] dex Ms Rs mV MV v sin(i)a ρs Age Distance True space motion Uc True space motion Vc True space motion Wc 1.2±0.1 M⊙ 1.15+0.10 −0.12 R⊙ 16.13±0.04 3.55+0.27 −0.38 7±2 km s−1 0.79+0.31 −0.18 ρsun −0.4 kpc [0.6, 4.5] Gyr 3.2+0.9 13±28 km s−1 20±38 km s−1 -12±26 km s−1 a We assumed vmic=0.85 km s−1; b Epoch J2000; c Left-handed coordinates system (see text). determine to which component of our Galaxy it belongs. In order to compute the U, V and W components of the motion, the following set of quantities was required: dis- tance, systemic velocity (from the RV fit, see Section 3.2.2), proper motion (from SDSS 7th release, Munn et al. 2004, 2008) and coordinates of the star. The distance to the ob- served system was estimated according to the extinction, fitted in the SED analysis (AV =0.44 ± 0.07), and a model of dust distribution in the galaxy (Amores & L´epine 2005, axis-symmetric model). The UVW values and their errors are calculated using the method in Johnson & Soderblom (1987), with respect to the Sun (heliocentric) and in a left- handed coordinate system, so that they are positive away from the Galactic centre, Galactic rotation and the North Galactic Pole respectively. All the quantities here discussed are listed in Table 5. Considering the uncertainties on the derived quantities, maily affected by the error on the dis- tance, the host star is consistent with both the definitions of Galactic young-old disk and young disk populations (metal- licity between -0.5 and 0 dex, solar-metallicity respectively, Leggett 1992). But we could assess that WTS-1 is not a halo member. Combining ρs and Ms, a value of 1.15+0.10 −0.12 R⊙ was com- puted for the stellar radius. The same result was obtained considering the stellar mass and the surface gravity mea- sured from the spectroscopic analysis. Scaling by the dis- tance the apparent V magnitude calculated from the SDSS g and r magnitudes (Jester et al. 2005), we computed an absolute V magnitude of 3.55+0.27 −0.38. This value and all the other derived stellar parameters are consistent with each other and with the typical quantities expected for an F6-8 main-sequence star (Torres et al. 2009). They are collected in Table 5. Figure 5. High-resolution HET stacked spectrum (black full line) of the WTS-1 Li i line at ∼6707 A in comparison with three syn- thetic spectra calculated with the final adopted stellar parameters and lithium abundances (log N(Li)) of 2.5 (red dashed line), 2.9 (blue dotted line), and 2.1 (blue dash-dotted line). The vertical line indicates the position of the centre of the multiplet. and a metallicity range −0.5 ÷ 0.0 dex. As described in Sec- tion 3.1.2, the faintness of the star and the reduction process led to a distortion of the stronger lines in the HET high res- olution spectrum, reducing the number of reliable lines em- ployed in the measure of the metallicity. For these reasons, the relative uncertainty on the metallicity is larger with re- spect to those of the other parameters. Further observations would be needed to pin down the exact value of the star metallicity. In order to determine the parameters of the stellar com- panion, mass and radius of the host star must be known. The fit of the transit in the light curves provides important constrains on the mean stellar density (see Section 3.2.1). Joining this quantity to the effective temperature, we could place WTS-1 in the modified ρ−1/3 vs. Teff H-R diagram and compare its position with evolutionary tracks and isochrones models (Girardi et al. 2000) in order to estimate stellar mass and age. In this way, we estimated the stellar mass to be 1.2±0.1 M⊙ and the age of the system to range between 200 Myr and 4.5 Gyr. s Further constrains on the stellar age can be fixed consid- ering the measured Li i line abundance. Depletion of lithium in stars hotter than the Sun is thought to be due to a not yet clearly identified slow mixing process during the main- sequence evolution, because those stars do not experience pre-main sequence depletion (Mart´ın 1997). Comparison of the lithium abundance of WTS-1 (log N(Li)=2.5±0.4) with those of open clusters rise the lower limit on the age to 600 Myr, because younger clusters do not show lithium de- pletion in their late-F members (Sestito & Randich 2005). On the other hand, it is not feasible to derive an age con- strain from the WTS-1 rotation rate (v sin(i)=7±2 km s−1) because stars with spectral types earlier than F8 show no age-rotation relation (Wolff et al. 1986) and thus they are left out from the formulation of the spin down rate of low- mass stars (Stepi´en 1988). The true space motion knowledge of WTS-1 allows to c(cid:13) 2012 RAS, MNRAS 000, 1 -- 15 10 M. Cappetta et al. Table 6. Quadratic limb-darkening coefficients used for the tran- sit fitting, for a star with effective temperature Teff =6250 K, surface gravity log g=4.4 and metallicity [Fe/H]=-0.5 dex (Claret & Bloemen 2011). filter γ1 γ2 J i′ 0.14148 0.25674 0.24832 0.26298 3.2 Planetary parameters 3.2.1 Transit fit The light curves in J- and i′-band were fitted with ana- lytic models presented by Mandel & Agol (2002). We used quadratic limb-darkening coefficients for a star with effec- tive temperature Teff =6250 K, surface gravity log g=4.4 and metallicity [Fe/H]=-0.5 dex, calculated as linear interpola- tions in Teff , log g and [Fe/H] of the values tabulated in Claret & Bloemen (2011). We use their table derived from ATLAS atmospheric models using the flux conservation method (FCM) which gave a slightly better fit than the ones derived using the least-squares method (LSM). The values of the limb-darkening coefficients we used in our fitting are given in Table 6. Scaling factors were applied to the error values of the J- and i′-band light curves (0.94 and 0.9 re- spectively) in order to achieve a reduced χ2 of the constant out-of-transit part equal to 1. Using a simultaneous fit to both light curves, we fitted the period P , the time of the central transit t0, the radius ratio Rp/Rs, the mean stellar density, ρs = Ms/R3 s in solar units and the impact parameter βimpact in units of Rs. The light curves and the model fit are shown in Figures 6 and 7, while the resulting parameters are listed in Table 7. The errors were calculated using a multi-dimensional grid on which we search for extreme grid points with ∆χ2=1 when varying one parameter and simultaneously minimizing over the others. Figure 8 shows the correlations between the parameters of the WTS-1 system derived from the simulta- neous fit of the J- and i′-band light curves. Considering the solar metallicity scenario, with different limb-darkening co- efficients, the change in the final fitting parameters is smaller than 1% of their uncertainties. Note that the combined fit as- sumes a fixed radius ratio although in a hydrogen-rich atmo- sphere, molecular absorption and scattering processes could result in different radius ratios in each band (an attempt to detect such variations has recently been undertaken by de Mooij et al. 2012). In our case, the uncertainties are too large to see this effect in the light curves and the assump- tion of a fixed radius ratio is a good approximation. The estimated stellar density of the host star (0.79+0.31 −0.18 ρsun) is consistent with the expected value based on its spectral type (Seager & Mallen-Ornelas 2003). The noise in the data did not allow a secondary transit detection. Subsequently, we searched for further periodic signals in the light curve after the removal of the data points related to the transit. No significant signals were detected in the Lomb- Scargle periodogram up to a period of 400 days. Since WTS- 1 is a late F-star, there are not many spots on the surface Figure 6. WFCAM J-band light curve data of WTS-1. U pper panel: whole set of the folded WFCAM J-band data points. M iddle panel: folded photometric data centred in the transit and best model (red line) fitted in combination with the INT i′-band data. Lower panel: residuals of the best fit. Figure 7. INT i′-band light curve data of WTS-1. U pper panel: photometric data and best model (red line) fitted in combination with the WFCAM J-band data. Lower panel: residuals of the best fit. and they do not live long enough to produce a stable signal over a timescale of several years. 3.2.2 Radial velocity One of the most pernicious transit mimics in the WTS are eclipsing binaries. On one hand, a transit can be mimicked c(cid:13) 2012 RAS, MNRAS 000, 1 -- 15 Inflated HJ in close orbit around a late F-star 11 4 2 0 ] g e d [ n o i t a n i l c n ∆ i −2 −0.5 0.0 ∆Rplanet [Rjup] ] g e d [ n o i t a n i l c n ∆ i 4 2 0 −2 ] p u R j 0.0 [ t e n a p R ∆ l −0.5 −0.2 0.0 ∆Rstar [Rsun] 0.2 −0.2 0.0 ∆Rstar [Rsun] 0.2 ] s [ d o i r e p ∆ 2 0 −2 −4 −10 10 20 0 ∆t0 [min] Figure 8. Correlation plots of the quantities derived from the simultaneous fit of the transit in the J and i′-band light curves. The χ2 minima are indicated by crosses while the different tones of grey correspond to the 68%, 95% and 99% confidence level (darker to lighter respectively). The other couples of parameters do not show significant correlations. Table 7. Fitted parameters of the WTS-1 system as determined from the simultaneous fit of the J- and i′-band light curves. Scal- ing factors to the uncertainties of the J and i′ data points (0.94 and 0.9 respectively) were applied in order to achieve a reduced χ2=1 in the constant out-of-transit part of the light curves. Parameter Value Porb = 3.352059+1.2×10−5 −1.4×10−5 days t0 = 2 454 318.7472+0.0043 −0.0036 HJD Rp/Rs = 0.1328+0.0032 −0.0035 −0.18 ρsun ρs = 0.79+0.31 βimpact = 0.69+0.05 −0.09 by an eclipsing binary that is blended with foreground or background star. On the other hand, grazing eclipsing bi- naries with near-equal radius stars also have shallow, near- equal depth eclipses that can phase-fold into transit-like sig- nals at half the binary orbital period. In order to rule out the eclipsing binaries scenarios, the RV variation of the WTS-1 system were first measured using the ISIS/WHT intermedi- ate resolution spectra. Eclipsing binaries systems typically show RV amplitudes of tens of km s−1, while the measured RVs were all consistent with a flat trend within the RV un- certainties of ∼ 1 km s−1. Afterwards, we analysed the high-resolution HET spec- tra in order to accurately investigate the properties of the sub-stellar companion of WTS-1. The spectra related to each single night were cross-correlated using the iraf.rv.fxcorr task with the synthetic spectrum of a star with Teff =6250 K, log g=4.4 and [Fe/H]=-0.5. Changes in the effective temper- ature of the synthetic templates, even of the order of sev- eral hundreds of Kelvin, cause variations of the measured RV values smaller than the statistical uncertainties due to noise in the spectra (suggesting the absence of contami- nation of back/foreground stars of different spectral type). Even smaller variations occur changing surface gravity and metallicity of the template. The 40 single RV values, each one coming from a different order, were used to compute the RVs and related uncertainty at each epoch of the obser- vations. Resampling statistical tools were used in order to c(cid:13) 2012 RAS, MNRAS 000, 1 -- 15 Figure 9. T op panel: RVs values measured with the high- resolution HET spectra of WTS-1 as a function of the orbital phase. Black dots and blue triangles refers to observations per- formed in 2010 and 2011 respectively. The data point at φ = 0.33, empty dot, was excluded from the fitting procedure (see text for details). Best-fit circular orbit model (χ2 ν =1.45) and 1σ uncer- tainty of the semi-amplitude (K⋆=479±34 m s−1) are indicated with solid and dashed red lines respectively. The black dotted line refers to the fitted radial systemic velocity (γ=-1 714±35 m s−1). Zero phase corresponds to the mid-transit time. M iddle panel: Phase folded O-C residuals from the best-fit. The residual scatter is of the order of ∼ 150 m s−1, consistent with the RVs uncer- tainties. Bottom panel: Bisector spans. No significant deviations from zero and no correlation with the RVs were found. 12 M. Cappetta et al. Table 8. Radial velocities and bisector spans measurements for WTS-1 obtained by HET spectra. The phases were computed from the epochs of the observations expressed in Julian date and using the P and t0 values found with the transit fit. Table 9. Properties of the new extrasolar planet WTS-1b. Parameter Value HJD Phase RV BS -2 400 000 55477.676 55479.666 55499.608 55513.586 55522.556 55523.542 55742.729 55760.673 55782.837 55824.722 55849.650 [km s−1] [km s−1] −1.46 ± 0.12 −0.31 ± 0.40 −1.80 ± 0.10 −0.77 ± 0.39 −2.03 ± 0.10 −2.06 ± 0.17 0.76 ± 0.51 0.87 ± 0.63 −2.09 ± 0.44 −0.39 ± 0.35 −1.70 ± 0.16 −0.13 ± 0.44 −1.97 ± 0.15 0.27 ± 0.68 −1.11 ± 0.12 −0.47 ± 0.59 −2.23 ± 0.07 −2.31 ± 0.06 0.41 ± 0.39 0.16 ± 0.56 −1.19 ± 0.06 0.18 ± 0.31 0.74 0.33 0.28 0.45 0.13 0.42 0.26 0.79 0.22 0.25 0.75 better estimate mean value, standard deviation and possi- ble bias in the sample of measured RVs. Finally, the mea- sured RVs were corrected for the Earth orbital movements and reduced to the heliocentric rest-of-frame. The phase val- ues φ were computed from the epochs of the observations, expressed in Julian date, and using the extremely well de- termined P and t0 values (relative uncertainties are of the order of 10−6 and 10−9 respectively) obtained from the pho- tometric fit (see Table 7). The data, listed in Table 8, were then fitted with a simple two parameters sinusoid of the form: RV = γ + K⋆sin(2πφ) (1) where K⋆ is the RV semi-amplitude of the host star and γ is the systemic velocity of the system. Thanks to the acquisition of two ThAr calibration exposures (before and after the science exposure, see Section 2.2.4), we detected a small drift between the ThAr lines occurring during the science exposures on November 22, 2010. In the presence of a suspected systematic trend, which could affect the measured RV value, we performed the fitting procedure excluding the data point related to that night (at φ = 0.33). The larger RV error of the data point at φ = 0.13 is due to the integration time (half hour) of the science frame which is shorter than those of all the other data points (one hour). The best fitting model (χ2 ν=1.45) was obtained for K⋆=479±34 m s−1and γ=-1 714±35 m s−1and is plotted in Figure 9 with the RV data. We imposed the orbit to be circular as the eccentricity was compatible with zero when a Keplerian orbit fit was performed (see Anderson et al. (2012) for a discussion of the rationale for this). In accordance to the RV uncertainties, a relatively loose upper limit can be plaved on the eccentricity (e < 0.1, C.L.= 95%). The fitted RV semi-amplitude implies a planet mass of 4.01±0.35 MJ assuming a host star mass of 1.2±0.1 M⊙ (see Section 3.1.3). The uncertainty on the planet mass in mainly driven by the uncertainty on the mass of the host star. As can be seen in Figure 9, the RVs related to observations performed in late 2010 and in the second half of 2011 are consistent, showing no significant long term trends in our measurements. Mp Rp Prot a e inc βimpact t0 ρp a Teq 4.01±0.35 MJ 1.49+0.16 −0.18 RJ 3.352057+1.3×10−5 −1.5×10−5 d 0.047±0.001 AU −0.7 deg < 0.1 (C.L.= 95%) 85.5+1.0 0.69+0.05 −0.09 2 454 318.7472+0.0043 1.61±0.56 g cm−3, 1.21±0.42 ρJ 1500±100 K −0.0036 HJD a We assumed a Bold albedo AB=0 and re-irradiating fraction F= 1; The HET spectra were employed also to investigate the possibility that the measured RVs are not due to true Doppler motion in response to the presence of a planetary companion. Similar RV variations can rise in case of dis- tortions in the line profiles due to stellar atmosphere os- cillations (Queloz et al. 2001). To assert that this is not our case, we used the same cross-correlation profiles pro- duced previously for the RV calculation to compute the bi- sector spans (BS hereafter) which are reported in Table 8. Following Torres et al. (2005), we measured the difference between the bisector values at the top and at the bot- tom of the correlation function for the different observation epochs. In case of contaminations, we would have expected to measure BS values consistently different from zero and a strong correlation with the measured RVs (Queloz et al. 2001; Mandushev et al. 2005). As it can be seen in the bot- tom panel of Figure 9, the measured BS do not show signif- icant deviation from zero within the uncertainties. No cor- relation was detected between the BS and the RV values. In this way, contaminations that could mimic the effect of the presence of a planet were ruled out. 4 DISCUSSION AND CONCLUSIONS In this paper we announce the discovery of a new transit- ing extrasolar planet, WTS-1b, the first detected by the UKIRT/WFCAM Transit Survey. The parameters of the planet are collected in Table 9. WTS-1b is a ∼4 MJ planet orbiting in 3.35 days a late F-star with possibly slightly sub- solar metallicity. With a radius of 1.49+0.16 −0.18 RJ , it is lo- cated in the upper part of the mass -- radius diagram of the known extrasolar planets in the mass range 3-5 MJ (see Figure 10). The parameters of the other planets are taken from www.exoplanet.eu at the time of the publication of this work. Planets with only an upper limit on the mass and/or on the radius are not shown. It is worth noting that only a cut-off of the transit depth, different for each survey, could act as a selection effect against the detection of plan- ets in this upper portion of the diagram. Larger planetary radii imply a deeper transit feature in the light curves and c(cid:13) 2012 RAS, MNRAS 000, 1 -- 15 Inflated HJ in close orbit around a late F-star 13 down the cooling contraction of a HJ even on timescales of several Gyr: a surface wind blowing across the planetary magnetic field acts as a battery that rises Ohmic dissipa- tion in the deeper layers. In Huang & Cumming (2012), the Ohmic dissipation in HJs is treated decoupling the interior of the planet and the wind zone. In this scenario, the radius evolution for an irradiated HJ planet (see their Figure 9, 3 MJ ) leads to a value consistent with our observation up to 3 Gyr. Accordingly to Fortney et al. (2008), the incident flux (the amount of energy from the host star irradiation, per unit of time and surface, that heats the surface of the planet) computed for WTS-1b (1.12±0.26 · 109 erg s−1 cm−2) as- signs it to the so called pM class of HJs. This classification considers the day-side atmospheres of the highly-irradiated HJs that are somewhat analogous to the M- and L-type dwarfs. In particular, the predictions of equilibrium chem- istry for pM planet atmospheres are similar to M-dwarf stars, where absorption by TiO, VO, H2O, and CO is promi- nent (Lodders 2002). Planets in this class are warmer than required for condensation of titanium (Ti)- and vanadium (V)-bearing compounds and will possess a temperature in- version (which could lead to a smaller inflation due to Ohmic heating accordingly to Heng (2012)) due to absorption of in- cident flux by TiO and VO molecules. Fortney et al. (2008) propose that these planets will have large day/night effective temperature contrasts and an anomalous brightness in sec- ondary eclipse at mid-infrared wavelengths. Unfortunately, the SNR in the J-band light curve, due to the faintness of the parent star WTS-1, is not high enough for such kind of detection (see Section 3.2.1). To conclude, the discovery of WTS-1b demonstrates the capability of WTS to find planets, even if it operates in a back-up mode during dead time on a queue-schedule tele- scope and despite of the somewhat randomised observing strategy. Moreover, WTS-1b is an inflated HJ orbiting a late F-star even if the project is designed to search for ex- trasolar planets hosted by M-dwarfs. Birkby et al. 2012b will present the second WTS detection, WTS-2b, a Jupiter-like planet around a cool K-star. ACKNOWLEDGMENTS We acknowledge support by RoPACS during this research, a Marie Curie Initial Training Network funded by the European Commissions Seventh Framework Programme. The United Kingdom Infrared Telescope is operated by the Joint Astronomy Centre on behalf of the Science and Technology Facilities Council of the U.K.; some of the data reported here were obtained as part of the UKIRT Service Programme. The Hobby-Eberly Telescope (HET) is a joint project of the University of Texas at Austin, the Pennsylvania State University, Stanford Univer- sity, Ludwig-Maximilians-Universitat Munchen, and Georg- August-Universitat Gottingen. The HET is named in honor of its principal benefactors, William P. Hobby and Robert E. Eberly. The 2.5m Isaac Newton Telescope and the William Herschel Telescope are operated on the island of La Palma by the Isaac Newton Group in the Spanish Observatorio del Roque de los Muchachos of the Instituto de Astrof´ısica de Canarias. We thank Calar Alto Observatory, the German- Figure 10. Mass -- Radius diagram of the known planets with a mass in the range 3-5 MJ (black dots). Labels with the related planet name are shown for an easier identification. Planets with only an upper limit on mass and/or radius are not shown. The blue dashed lines represent the iso-density curves. The green dot- dashed lines indicate the planetary radii at different ages accord- ingly to Fortney et al. (2007) (see text for details). Masses, radii of the planets are taken from www.exoplanet.eu at the time of the publication of this work, while the related uncertainties were found in the refereed publication. WTS-1b is shown in red. thus, within this mass range, larger object are more easily detectable. The properties of WTS-1b, as well as those of the other two planets present in the upper part of the dia- gram, CoRoT-2b (Alonso et al. 2008) and OGLE2-TR-L9b (Snellen et al. 2009), are not explained within standard for- mation and evolution models of isolated gas giant planets (Guillot 2005). The radius anomaly is at the ∼ 2σ level considering the stellar irradiation that retards the contraction of the planets, the distance of the planet from the host star and the age of the planet (Fortney et al. 2007). The models of Fortney and collaborators predict indeed a radius of 1.2 RJ for a 600 Myr-old planet (see their Figure 5, 3 MJ and 0.045 AU model). This radius estimate is an upper limit as 600 Myr is the lower limit on the age of the WTS-1 system due to the Li abundance (see Section 3.1.3). The radius trend shown in the figure would suggest an age for WTS-1b less than 10 Myr. The same significance on the radius anomaly is obtained considering empirical relationships coming from the fit of the observed radii as a function of the physical properties of the star-planet system, such as mass, equilibrium temperature and tidal heating (Enoch et al. 2012, eq. 10). In any case, a rapid migration of WTS-1b inward to the highly-irradiated domain after its formation seems required. Surface day/night temperature gradients due to the strong incident irradiation, are likely to generate strong wind activity through the planet atmosphere. Recently, Wu & Lithwick (2012) showed how the Ohmic heating pro- posal (Batygin & Stevenson 2010; Perna et al. 2010) can ef- fectively bring energy in the interior of the planet and slow c(cid:13) 2012 RAS, MNRAS 000, 1 -- 15 14 M. Cappetta et al. Spanish Astronomical Center, Calar Alto, jointly operated by the Max-Planck-Institut fur Astronomie Heidelberg and the Instituto de Astrof´ısica de Andaluc´ıa (CSIC), for allo- cation of director's discretionary time to this program. We thank Kitt Peak National Observatory, National Optical As- tronomy Observatory, which is operated by the Association of Universities for Research in Astronomy (AURA) under cooperative agreement with the National Science Founda- tion. This publication makes use of VOSA, developed under the Spanish Virtual Observatory project supported from the Spanish MICINN through grant AyA2008-02156. This re- search has been funded by Spanish grants AYA 2010 -- 21161 -- C02 -- 02, CDS2006 -- 00070 and PRICIT -- S2009/ESP -- 1496. This work was partly funded by the Funda¸cao para a Ciencia e a Tecnologia (FCT)-Portugal through the project PEst- OE/EEI/UI0066/2011. MC is grateful to JB for the cordial and fruitful collaboration. MC thanks A. Driutti, B. Sartoris and P. Miselli for technical support and stimulating discus- sions. NL was funded by the Ram´on y Cajal fellowship num- ber 08-303-01-02 and the national program AYA2010-19136 funded by the Spanish ministry of science and innovation. This work was co-funded under the Marie Curie Actions of the European Commission (FP7-COFUND). EM was partly supported by the CONSOLIDER-INGENIO GTC project and the project AYA2011-30147-C03-03. This research has made use of NASA's Astrophysics Data System. REFERENCES Aigrain S., Irwin M., 2004, MNRAS, 350, 331 Alonso R., Auvergne M., Baglin A., Ollivier M., Moutou C., Rouan D., Deeg H. J., Aigrain S., Almenara J. M., Barbieri M., Barge P. Benz W., Bord´e P., Bouchy F., de La Reza R., Deleuil M., Dvorak R., Erikson A., Fridlund M., Gillon M., Gondoin P., Guillot T., Hatzes A., H´ebrard G., Kabath P., Jorda L., Lammer H., L´eger A., Llebaria A., Loeillet B., Magain P., Mayor M., Mazeh T., Patzold M., Pepe F., Pont F., Queloz D., Rauer H., Shporer A., Schneider J., Stecklum B., Udry S., Wuchterl G., 2008, A&A, 482L, 21A Amores E. B., L´epine J. R. D., 2005, AJ, 130, 659 Anderson D. R., Collier Cameron A., Gillon M., Hellier C., Jehin E., Lendl M., Maxted P. F. L., Queloz D., Smalley B., Smith A. M. S., Triaud A. H. M. J., West R. G., Pepe F., Pollacco D., S´egransan D., Todd I., Udry S., 2012, MNRAS, 422, 1988 Baraffe I., Chabrier G., Allard F., Hauschildt P. H., 1998, A&A, 337, 403B Baraffe I., Chabrier G., Barman T., 2010, arXiv, 1001, 3577 Barbuy B., 1982, PhD thesis, Universit´e de Paris VII Barbuy B., Perrin M. N., Katz D., Coelho P., Cayrel R., Spite M., Van't Veer-Menneret C., 2003, A&A, 404, 661 Batygin K., Stevenson D. J., 2010, ApJ, 714, 238B Bayo A., Rodrigo C., Barrado y Navascu´es D., Solano E., Guti´errez R., Morales-Calder´on M., Allard F., 2008, A&A, 492, 277B Bayo A., submitted, 2012 Birkby J., Hodgkin S., Pinfield D., WTS consortium, 2011, ASPC, 448, 803B Mooij E., Goulding N., CruzP., Stoev h., Cappetta M., Palle E., Barrado D., Saglia R. P., Mart´ın E., Pavlenko Y., 2012a, MNRAS accepted, arXiv, 1206, 2773B Birkby et al. 2012b, in prep Bodenheimer P., Lin D. N. C., Mardling R. A., 2001, ApJ, 548, 466B Bodenheimer P., Laughlin G., Lin D. N. C., 2003, ApJ, 592, 555B Burrows A., Hubeny I., Budaj J., Hubbard W. B., 2007, ApJ, 661, 502B Cayrel R., Perrin M. N., Barbuy B., Buser R., 1991, A&A, 247, 108 Canuto V. M., Mazzitelli I., 1991, ApJ, 370, 295C Casali M., Adamson A., Alves de Oliveira C., Almaini O., Burch K., Chuter T., Elliot J., Folger M., Foucaud S., Hambly N., Hastie M., Henry D., Hirst P., Irwin M., Ives D., Lawrence A., Laidlaw K., Lee D., Lewis J., Lunney D., McLay S., Montgomery D., Pickup A., Read M., Rees N., Robson I., Sekiguchi K., Vick A., Warren S., Woodward B., 2007, A&A, 467, 777C Castelli F., Gratton R.G., Kurucz R.L., 1997, A&A, 318, 841 Castelli F., Kurucz R.L., 1997, Modelling of Stellar Atmo- spheres, 210, 20P Chabrier G., Baraffe I., 2007, ApJ, 661L, 81C Claret A., Bloemen S., 2011, A&A, 529, 75 Coelho P., Barbuy B., Melndez J., Schiavon R. P., Castilho B. V., 2005, A&A, 443, 735C Collier Cameron A., Wilson D. M., West R. G., Hebb L., Wang X. B., Aigrain S., Bouchy F., Christian D. J., Clark- son W. I., Enoch B., Esposito M., Guenther E., Haswell C. A., H´ebrard G., Hellier C., Horne K., Irwin J., Kane S. R., Loeillet B., Lister T. A., Maxted P., Mayor M., Moutou C., Parley N., Pollacco D., Pont F., Queloz D., Ryans R., Skillen I., Street R. A., Udry S., Wheatley P. J., 2007, MNRAS, 380, 1230 Covey K. R., Ivezi ., Schlegel D., Finkbeiner D., Padman- abhan N., Lupton R. H., Agueros M. A., Bochanski J. J., Hawley S. L., West A. A., Seth A., Kimball A., Gogarten S. M., Claire M., Haggard D., Kaib N., Schneider D. P., Sesar B., 2007, AJ, 134, 2398 Demory B. O., Seager S., 2011, ApJS, 197, 12D Enoch B., Collier Cameron A., Horne K., 2012, A&A, 540, 99 Fossati L., Ryabchikova T., Bagnulo S., Alecian E., Grun- hut J., Kochukhov O., Wade G., 2009, AAP, 503, 945 Fossati L., Ryabchikova T., Shulyak D. V., Haswell C. A., Elmasli A., Pandey C. P., Barnes T. G., Zwintz K. 2011, MNRAS, 417, 495 Fortney J. J., Marley M. S., Barnes J. W., 2007, ApJ, 659, 1661 Fortney J. J., Lodders K., Marley M. S., Freedman R. S., 2008, ApJ, 678, 1419 Fuhrmann K., Axer M., Gehren T., 1993, AAP, 271, 451 Girardi L., Bressan A., Bertelli G., Chiosi C., 2000, A&AS, 141, 371G Goulding N. T., et al., 2012, in prep. Guillot T., 2005, Ann Rev EPS, 33, 493G Heng K., 2012, ApJ, 748, L17 Hodgkin S. T., Irwin M. J., Hewett P. C., Warren S. J., Birkby J., Nefs B., Hodgkin S., Kov´acs G., Sipocz B., Pin- field D., Snellen I., Mislis D., Murgas F., Lodieu N., de 2009, MNRAS, 394, 675H Huang X., Cumming A., 2012, arXiv, 1207, 3278 c(cid:13) 2012 RAS, MNRAS 000, 1 -- 15 Irwin J., Irwin M., Aigrain S., Hodgkin S., Hebb L., Moraux Chem. Ref. Data, 27, 1275 Inflated HJ in close orbit around a late F-star 15 Ryabchikova T., Piskunov N., Savanov I., Kupka F., Malanushenko V., 1999, A&A, 343, 229R Ryabchikova T. A., Fossati L., Shulyak D. V., 2009, AAP, 506, 203 Seager S., Mallen-Ornelas G., 2003, ApJ, 585, 1038 Seager S., Kuchner M., Hier-Majumder C. A., Militzer B., 2007, ApJ, 669, 1279S Sekiguchi M., Fukugita M., 2000, AJ, 120, 1072S Sestito, P. & Randich, S. 2005, A&A, 442, 615 Showman A. P., Guillot T., 2002, A&A , 385, 166 Shulyak D., Tsymbal V., Ryabchikova T., Sttz Ch., Weiss W. W., 2004, A&A, 428, 993S Sipocz B., et al., 2012, in prep. Skrutskie M. F., Cutri R. M., Stiening R., Weinberg M. D., Schneider S., Carpenter J. M., Beichman C., Capps R., Chester T., Elias J., Huchra J., Liebert J., Lonsdale C., Monet D. G., Price S., Seitzer P., Jarrett T., Kirk- patrick J. D., Gizis J. E., Howard E., Evans T., Fowler J., Fullmer L., Hurt R., Light R., Kopan E. L., Marsh K. A., McCallon H. L., Tam R., Van Dyk S. and Wheelock S., 2006, AJ, 131, 1163 Snellen I. A. G., Koppenhoefer J., van der Burg R. F. J., Dreizler S., Greiner J., de Hoon M. D. J., Husser T. O., Kruhler T., Saglia R. P., Vuijsje F. N., 2009, A&A, 497, 545 Stepi´en K., 1988, ApJ, 335, 907 Torres G., Konacki M., Sasselov D. D., Jha S., 2005, ApJ, 619, 558 Torres G., J. Andersen, A. Gim´enez, 2009, A&A, 18, 67 Tsymbal V., 1996, ASPC, 108, 198T Tull R. G., 1998, SPIE, 3355, 387 Valenti J. and Fischer D. A., 2005, ApJ, 622, 1102 Vogt S. S., Allen S. L., Bigelow B. C., Bresee L., Brown B., Cantrall T., Conrad A., Couture M., Delaney C., Epps H. W., Hilyard D., Hilyard D. F., Horn E., Jern N., Kanto D., Keane M. J., Kibrick R. I., Lewis J. W., Osborne J., Pardeilhan G. H., Pfister T., Ricketts T., Robinson L. B., Stover R. J., Tucker D., Ward J., Wei M. Z., 1994, SPIE, 2198, 362V Willmarth D., Barnes J., A user's guide to reducing Echelle spectra with IRAF, 1994 Wolff S. C., Boesgaard A. M., Simon T., 1986, ApJ, 310, 360 Wright E. L., Eisenhardt P. R. M., Mainzer A. K., Ressler M. E., Cutri R. M., Jarrett T., Kirkpatrick J. D., Padgett D., McMillan R. S., Skrutskie M., Stanford S. A., Cohen M., Walker R. G., Mather J. C., Leisawitz D., Gautier T. N., McLean I., Benford D., Lonsdale C. J., Blain A., Mendez B., Irace W. R., Duval V., Liu F., Royer D., Hein- richsen I., Howard J., Shannon M., Kendall M., Walsh A. L., Larsen M., Cardon J. G., Schick S., Schwalm M., Abid M., Fabinsky B., Naes L., Tsai C., 2010, AJ, 140, 1868 Wu Y., Lithwick Y., 2012, arXiv, 1202, 0026 Youdin A. N., Mitchell J. L., 2010, ApJ, 721, 1113Y Zapolsky H. S., Salpeter E. E., 1969, ApJ...158..809Z E., 2007, MNRAS, 375, 1449I Irwin M., Lewis J., 2001, New Astronomy Reviews, 45, 105 Jackson B., Greenberg R., Barnes R., 2008, ApJ, 678, 1396J Jester S., Schneider D. P., Richards G. T., Green R. F., Shmidt M., Hall P. B., Strauss M. A., Vanden Berk D. E., Stoughton C., Gunn J. E., Brinkmann J., Kent S. M., Smith J. A., Tucker D. L., Yanny B., 2005, AJ, 130, 873J Johnson D. R. H., Soderblom D. R., 1987, AJ, 93, 864 Kochukhov O., 2007, Spectrum synthesis code SYNTH3 Kov´acs B., Hodgkin S., Sipocz B., Pinfield D., Barrado D., Birkby J., Cappetta M., Cruz P., Koppenhofer J., Mart´ın E., Murgas F., Nefs B., Saglia R. P., Zendejas J., 2012, MNRAS submitted Kupka F., Piskunov N., Ryabchikova T. A., Stempels H. C., Weiss W. W., 1999, A&AS, 138, 119 Kurucz R. L., 1993, Stells surface structure, 523 Leconte J., Chabrier G., 2012, A&A, 540A, 20L Laughlin G., Crismani M., Adams F. C., 2011, ApJ, 729, 7 Leggett S. K., 1992, ApJSS, 82, 351 Lodders K., ApJ, 577, 974L Liu X, Burrows A., Ibgui L., 2008, ApJ, 687, 1191L Lin D. N. C., Bodenheimer P., Richardson D. C., 1996, Nature, 380, 606L McMahon R. G., Walton N. A., Irwin M. J., Lewis J. R., Bunclark P. S., Jones D. H., 2001, New Astronomy Re- views, 45, 97 Mandel K., Agol E., 2002, ApJ, 580L, 171M Mandushev G., Torres G., Latham D. W., Charbonneau D., Alonso R., White R.Y., Stefanik R. P., Dunham E. W., Brown T.M., ODonovan F. T., 2005, ApJ, 621, 1061 Mart´ın E. L., 1997, A&A, 321, 492 Mart´ın E. L., Spruit H. C., Tata R., 2011, A&A, 535A, 50M Mashonkina L., 2011, mast.conf, 314M de Mooij E. J. W., Brogi M., de Kok R. J., Koppenhoefer J., Nefs S. V., Snellen I. A. G., Greiner J., Hanse J., Heins- broek R. C., Lee C. H., van der Werf P. P., 2012, A&A, 538, 46 Munn J. A., Monet D. G., Levine S. E., Canzian B., Pier J. R., Harris H. C., Lupton R. H., Ivez´ı Z., Hindsley R. B., Hennessy G. S., Schneider D. P.k Brinkmann J., 2004, AJ, 127, 3034M Munn J. A., Monet D. G., Levine S. E., Canzian B., Pier J. R., Harris H. C., Lupton R. H., Ivezi´Z., Hindsley R. B., Hennessy G. S., Schneider D. P.k Brinkmann J., 2008, AJ, 136, 895M Nash T., 1996, sube.conf, 477N Perna, R., Menou, K., Rauscher, E. 2010, ApJ, 724, 313 Piskunov N. E., Kupka F., Ryabchikova T. A., Weiss W. W., Jeffery C. S., 1995, A&AS, 112, 525P Pont F., Zucker S., Queloz D., 2006, MNRAS, 373, 231 Queloz D., Henry G. W., Sivan J. P., Baliunas S. L., Beuzit J. L., Donahue R. A., Mayor M., Naef D., Perrier C., Udry S., 2001, A&A, 379, 279 Ramsey L. W., Adams. M. T., Barnes T. G., Booth J. A., Cornell M. E., Fowler J. R., Gaffney N. I., Glaspey J. W., Good J. M., Hill G. J., Kelton P. W., Krabbendam V. L., Long L., MacQueen P. J., Ray F. B., Ricklefs R. L., Sage J., Sebring T. A., Spiesman W. J., Steiner M., 1998, SPIE, 3352, 34 Rosman K. J. R. & Taylor P. D. P. 1998, Journal Phys. c(cid:13) 2012 RAS, MNRAS 000, 1 -- 15
1709.09687
1
1709
2017-09-27T18:19:05
Paleohydrology on Mars constrained by mass balance and mineralogy of pre-Amazonian sodium chloride lakes
[ "astro-ph.EP", "physics.geo-ph" ]
Chloride-bearing deposits on Mars record high-elevation lakes during the waning stages of Mars' wet era (mid-Noachian to late Hesperian). The water source pathways, seasonality, salinity, depth, lifetime, and paleoclimatic drivers of these widespread lakes are all unknown. Here we combine reaction-transport modeling, orbital spectroscopy, and new volume estimates from high-resolution digital terrain models, in order to constrain the hydrologic boundary conditions for forming the chlorides. Considering a T = 0 degrees C system, we find: (1) individual lakes were >100 m deep and lasted decades or longer; (2) if volcanic degassing was the source of chlorine, then the water-to-rock ratio or the total water volume were probably low, consistent with brief excursions above the melting point and/or arid climate; (3) if the chlorine source was igneous chlorapatite, then Cl-leaching events would require a (cumulative) time of >10 yr at the melting point; (4) Cl masses, divided by catchment area, give column densities 0.1 - 50 kg Cl/m^2, and these column densities bracket the expected chlorapatite-Cl content for a seasonally-warm active layer. Deep groundwater was not required. Taken together, our results are consistent with Mars having a usually cold, horizontally segregated hydrosphere by the time chlorides formed.
astro-ph.EP
astro-ph
Paleohydrology on Mars constrained by mass balance and mineralogy of pre- Amazonian sodium chloride lakes M. Melwani Daswani1, and E. S. Kite1 1 Department of the Geophysical Sciences, University of Chicago, 5734 S. Ellis Avenue Chicago, Illinois 60637, USA Corresponding author: Mohit Melwani Daswani ([email protected]) Key Points:  Halite lake deposits suggest very limited water-rock interaction on Early Mars, different to the extensive alteration evident at Gale Crater  Lakes were > 100 m deep and lasted > (101 – 103) yr  Punctuated high rates of volcanism raising temperatures above freezing could have supplied the chlorine for the salt deposits 1 Abstract Chloride-bearing deposits on Mars record high-elevation lakes during the waning stages of Mars' wet era (mid-Noachian to late Hesperian). The water source pathways, seasonality, salinity, depth, lifetime, and paleoclimatic drivers of these widespread lakes are all unknown. Here we combine reaction-transport modeling, orbital spectroscopy, and new volume estimates from high-resolution digital terrain models, in order to constrain the hydrologic boundary conditions for forming the chlorides. Considering a T = 0 °C system, we find: (1) individual lakes were >100 m deep and lasted decades or longer; (2) if volcanic degassing was the source of chlorine, then the water-to-rock ratio or the total water volume were probably low, consistent with brief excursions above the melting point and/or arid climate; (3) if the chlorine source was igneous chlorapatite, then Cl-leaching events would require a (cumulative) time of >10 yr at the melting point; (4) Cl masses, divided by catchment area, give column densities 0.1 – 50 kg Cl/m2, and these column densities bracket the expected chlorapatite-Cl content for a seasonally-warm active layer. Deep groundwater was not required. Taken together, our results are consistent with Mars having a usually cold, horizontally segregated hydrosphere by the time chlorides formed. 1. Introduction Ancient mineral deposits on planetary surfaces can serve as records of past climatic and environmental conditions. Discrete chloride mineral bearing sedimentary units occur on mid- Noachian to early Hesperian (3.9 – 3.5 Gyr old) crust on the southern highlands of Mars (Figure 1), and have been variously interpreted as a result of evaporation of ponds and lakes fed by surface runoff (Hynek et al., 2015; Osterloo et al., 2008) or groundwater upwelling (El-Maarry et al., 2013, 2014; Osterloo et al., 2008; Ruesch et al., 2012), or possibly a combination of both (Glotch et al., 2016; Osterloo et al., 2010). Here we use a novel, physically and chemically self-consistent method to constrain the paleohydrology of Mars when the chloride-bearing deposits formed by combining: (1) reaction-transport geochemical modeling of aqueous alteration on the surface of Mars, (2) mass balance and geological constraints on the origin of the chlorine in the chloride-bearing deposits, and (3) geomorphologic analyses of the deposits and the basins in which they were emplaced. This combination of geochemical modeling and basin analysis is enabled by recently published constraints from remote sensing and laboratory work on the chlorides (Glotch et al., 2016). 2 Figure 1. Map showing the location of the chloride-bearing deposits (black polygons) observed on the surface of Mars (Osterloo et al., 2010). The locations of the four sites studied here are shown in red. In the background, a dust cover index map overlays the shaded relief map, which uses MOLA topography (D. E. Smith et al., 2001). The dust cover index is a measure of the amount of silicate dust obscuring the surface to orbital spectroscopy, making use of the emissivity measured at 1350 – 1400 cm-1 by the Mars Global Surveyor Thermal Emission Spectrometer. Higher numbers mean less dust cover (Ruff & Christensen, 2002). As detected by thermal emission spectroscopy from orbit (THermal EMission Imaging System, THEMIS; (Christensen et al., 2004)), the chloride-bearing deposits (hereafter CBDs) tend to occur in local lows or basins (Osterloo et al., 2010), appear to be thick (on the order of meters (Hynek et al., 2015; Osterloo et al., 2010)), and often exhibit vertical polygonal fractures that have been interpreted as desiccation cracks (El-Maarry et al., 2013, 2014, Osterloo et al., 2008, 2010). Occasionally, the CBDs occur in inverted channels and/or infilling small craters (Osterloo et al., 2010). As such, the deposits appear to be paleoplayas or paleolakes, and inconsistent with salt efflorescence forming thin surficial crusts (Osterloo et al., 2008). While the chloride-bearing sinuous ridges are suggestive of surface runoff as the mechanism for transporting the brines which led to the CBDs, some regions that lack fluvial valley networks (e.g., S Noachis Terra) have abundant CBDs (Osterloo et al., 2010). (This does not exclude a correlation with runoff, because surface runoff does not necessarily result in channel formation, and small >3.5 Gyr channels might no longer be visible.) As an alternative to runoff as a water source, hydrothermal brines could have upwelled at topographic lows and subsequently evaporated, as has been interpreted for the origin of the halogen enrichment (~ 2 wt. % Cl compared to ~ 0.5 wt. % Cl in less altered basalts) analyzed by the Mars Exploration Rover (MER) Spirit at Home Plate, Gusev Crater(Schmidt et al., 2008). However, at the CBD sites, no evidence has been reported for an associated hydrothermal mineralogical assemblage (Osterloo et al., 2010), although this may have been obfuscated by e.g., wind erosion. Additionally, a widespread process of deep groundwater upwelling is inconsistent with the observation that some CBDs are found in local topographic lows at high elevation, while topographic lows at lower elevations often lack CBDs (Osterloo et al., 2010). While the chloride-bearing depositional facies are scattered across Mars, their distinctive geomorphology and local topographic setting suggests a single formation mechanism. However, these data offer no constraints on the duration, intensity, and number of wet events that led to CBD formation (Osterloo et al., 2010). A key provenance constraint is the non- 3 detection of other evaporite minerals (sulfates, carbonates, silica, other halides) in close proximity, overlain by, or in "association" with the CBDs (Osterloo et al., 2010). In fact, new emissivity scatter modeling and laboratory experimental validation of THEMIS spectra shows that the CBDs are composed of 10 – 25 vol. % halite (Glotch et al., 2016); with <5 wt. % gypsum or calcite (Ye & Glotch, 2016). Assuming no outflow, this signifies that the molar Cl/S ratio of the fluid that filled the lakes was ≲11 (for a molar volume of 74.7 cm3 for gypsum and 27.0 cm3 for halite). Evaporation of saline bodies of water on Earth typically forms sequences of evaporitic minerals defined by the fluid chemistry and the solubilities of salts present (Eugster & Hardie, 1978); broadly similar sequences are predicted from the evaporation of fluids on Mars (Tosca & McLennan, 2006). Hence, the fluids that formed the CBDs must have contained little S compared to Cl. A cartoon schematic of how the CBDs could have been formed is shown in Figure 2. Runoff and groundwater discharge into a topographic low would have ponded a fluid. Evaporative concentration would have then precipitated chloride minerals. CBD catchment areas, and lower limits on CBD volumes, can be accurately mapped today by combining digital terrain models (DTMs) and orbiter thermal emission spectroscopy (e.g., Hynek et al., 2015). This is because landscape modification by wind erosion has been generally subdued since the formation of the CBDs; the overall topography of the Southern Highlands has changed little since the CBDs formed (Nimmo & Tanaka, 2005). This enables a mass balance analysis that is not possible for more ancient aqueous deposits whose watersheds are less well preserved (Murchie et al., 2009). Brine fractionation is not investigated in this study, but it leads to paleohydrology conclusions similar to those of our T = 0 °C investigation. NaCl brines with eutectic points below the freezing point of water could have been responsible for the chloride enrichment in the observed deposits (Burt & Knauth, 2003; Clark et al., 2005). In this case, the absence of sulfate salts at the CBDs (Glotch et al., 2016; Osterloo et al., 2010) could be the result of brine fractionation since chloride brines have lower freezing points than sulfate brines (Marion, 2001; e.g., Reeburgh & Springer-Young, 1983; Toner & Sletten, 2013). Figure 2. Cartoon depicting the geological context of chloride-bearing lakes on Mars. Fluids flow as surface runoff and groundwater from left to right. At (A) the fluids react with the basin rocks and leach anions and cations from the rocks, including Cl and S deposited on the surface by volcanic gases. At (B), the salt-bearing fluid discharges at a topographic low, and at (C), salts precipitate after the ponding fluid has evaporated. Regardless of the physical process that led to the formation of the facies, the geochemical source of chlorine and cations (namely sodium, based on laboratory spectra and spectral 4 modeling; (Glotch et al., 2016)) has not been constrained. Possible origins for the chlorine could be: weathering and leaching of the drainage basin rock (section 2.1) (e.g., Eugster, 1980; Warren, 2010), volcanic outgassing (section 2.2) (Glotch et al., 2016; Osterloo et al., 2008; Tosca & McLennan, 2006; Zolotov & Mironenko, 2016), reworking (i.e., dissolution and transport) of previously formed evaporites (e.g., Salvany et al., 1994; Warren, 2010) (section 2.3), and cometary/extraterrestrial delivery by a halite-rich body (section 2.4) (e.g., H-chondrites like Zag and Monahans (Rubin et al., 2002)). Although some sodium may have been derived from volcanic degassing or exsolution from magma (e.g., Webster, 2004), most Na was probably derived from the interaction of a fluid with the martian basaltic crust (Tosca & McLennan, 2006; Zolotov & Mironenko, 2016). On Earth, the vast majority of Cl in the oceans is derived from volcanic degassing (including submarine vents) as opposed to continental weathering (Rubey, 1951; Spencer & Hardie, 1990), and maxima in Cl in ~500 km resolution Mars Odyssey Gamma Ray Spectrometer maps are spatially associated with modern Mars volcanoes (Keller et al., 2007). In contrast, Cl in terrestrial endorheic lakes is largely leached and mobilized from surrounding basin rocks (Eugster, 1980; Eugster & Hardie, 1978). Dry deposition from volcanic gases is unlikely to directly source chloride deposits on Mars. For example, the regions where pyroclastic deposits were predicted to form (Kerber et al., 2012, 2013) are not broadly coincident with the location of the chloride deposits observed by THEMIS. In section 2, we discuss the possible geological and geochemical origins of the chlorine in the CBDs and relate these to paleohydrology and paleoclimatology. In section 3, we describe the workflow we used to explore possible origins of chlorine, using compositional analyses of the martian surface, four different CBD-hosting basins (one a reevaluation of previous work by Hynek et al. (2015)), and two hydrological scenarios that potentially fed the lakes where the CBDs are found: surface runoff, and groundwater discharge. In section 4, we present our results, and finally, in light of the results we discuss the implications on Mars' paleoclimate and paleohydrology in section 5. 2. Physicochemical controls on plausible chlorine sources Hydrologic scenarios for the CBD forming event(s) can be parametrized using three variables: (1) water-to-rock mass ratio (hereafter W/R; a unitless measurement of the mass of water reacting with rock, divided by the total mass of rock that has reacted with water; see Reed (1997, 1998) for a detailed discussion), (2) the duration of the individual warming event (or events) that allowed liquid water to exist and form the Cl-bearing pools (τ, in Mars years, where 1 Mars year ≈ 5.94 × 107 s), and (3) the Total Water Volume (TWV, in kg H2O) flowing through the basin and discharging into the topographic low that will form a lake. W/R is linked to the mobility of elements and the 'open' vs 'closed' behavior in aqueous alteration systems with respect to chemical exchange with different reservoirs (e.g., Ehlmann et al., 2011). W/R controls the reaction path progress between the crust and the fluid. High W/R conditions occur at the interface between the fluid and the rock (e.g., fractures and the atmosphere-rock interface), where a dilute fluid comes in contact with the rock surface, and little rock is dissolved. Low W/R occurs where fluid chemistry is dominated by solutes leached from the rocks, e.g., at the end of a long groundwater travel path where much of the water permeating through the rock has been consumed to form secondary minerals (e.g., phyllosilicates in a weathering profile). W/R also controls the formation and fractionation of brines: chloride salts are usually more soluble than sulfate salts and require less water per unit mass to dissolve, so e.g., low W/R could form a chloride-only brine from a mixed chloride 5 and sulfate-bearing reactant whereas high W/R would put both salts in solution. Similarly, the evaporation of water from a mixed chloride-sulfate brine will cause sulfate salts to precipitate before chloride salts (e.g. Eugster, 1980; Tosca & McLennan, 2006). Our second hydrologic variable is τ. Climate models suggest that Mars in the Late Noachian to Early Hesperian had an equilibrium surface temperature below the freezing point of water on average (e.g., Wordsworth, 2016). Assuming this is correct, the availability of liquid water for water-rock interactions is controlled by the duration (τ) of the individual event or events that allowed temperatures above the freezing point of water (see Sections 2.1 and 2.2). We consider that "short" durations are always < 0.5 Mars years, such that the water required to form the chloride deposits could have been sourced from seasonal melting of snow and/or ice. We assume availability of snow or ground ice for episodic melting, consistent with the geomorphologic (e.g., Fassett & Head, 2008, 2011) and isotopic (e.g., Mahaffy et al., 2015; Villanueva et al., 2015) evidence that early Mars had abundant water. Total Water Volume (TWV) influences the geomorphology and the degree of chemical interaction between the rock and the fluid. High TWV typically allows larger amounts of rock to be dissolved, leading to the formation of an assemblage of secondary minerals formed from the rock-derived solutes. A low TWV would preclude deep lakes but might permit shallow pools. Low TWV allows only the most soluble minerals in the precursor rock to be dissolved, altering both fluid chemistry and the resultant mineralogical assemblage. 2.1. Basalt weathering as a source of chlorine Martian basalts are enriched in Cl by a factor of ~ 2.5 compared to terrestrial basalts and mantle rocks (Filiberto & Treiman, 2009). Cl is incompatible, so it is enriched in late-stage volatile-rich minerals (apatite and amphibole) formed in crystallizing magmas, or degassed (e.g. Aiuppa, 2009). Basalts containing these minerals will release Cl upon weathering by alteration fluids. Chlorapatite (Ca5(PO4)3Cl) could therefore act as a source of Cl (Adcock et al., 2013; Guidry & Mackenzie, 2003) for the brines forming the CBDs. As an illustration, we consider a single wet event that lasts just long enough to dissolve apatite at the surface. The dissolution rate of chlorapatite depends on pH, mass and surface area of the apatite grains. For soil that is close to the surface, in equilibrium with a 60 mbar pCO2 atmosphere (pH ≈ 4.5), and assuming that the apatite grain diameter in the martian regolith is similar to the apatite size found in martian meteorites (10-5 – 10-4 m in the basaltic breccia Northwest Africa 7034 (Wittmann et al., 2015)), we calculate that chlorapatite in the regolith dissolves completely in 0.04 – 0.4 Mars years, using a dissolution rate of 4.2 × 10-9 mol chlorapatite · m-2 · s-1 derived from Adcock et al. (2013). (See Appendix A for further details.) A single wet event of this duration suggests a weathering depth controlled by vertical diffusion of a top-down warming pulse through the surface soil/regolith: (1) where L is the depth reached by a warming pulse originating at the atmosphere-regolith interface (m), κ is the thermal diffusivity of the regolith (assumed here to be typical for silicates, i.e. 7 × 10-7 m2 s-1) and τ is the time (in seconds) of a single unfreezing event (e.g., Turcotte & Schubert, 2002, sec. 4.15). (We ignore the latent heat associated with melting a body of ice.) Thus, we calculate 3 – 10 m as the implied depth of unfreezing in the regolith in a single event lasting 0.04 – 0.4 Mars years. In this scenario, apatite at the top of the weathering profile will be completely dissolved and apatite at the bottom will have just 6 32.2L started dissolving. Assuming the composition of basalts encountered by the MER rover Spirit at Gusev crater (McSween et al., 2004, 2006) is representative of the composition of the martian surface elsewhere on Mars, on average, martian surface rocks contain ~ 0.15 wt. % Cl (Table 1), or 2.5 – 4.35 kg Cl · m-3 (using densities of 1650 kg · m-3 for Mars 200 µm grain sand and 2900 kg · m-3 for Mars basalt from Mellon et al. (2008)). Therefore, a priori, it appears that apatite dissolution could potentially supply significant chlorine to form the CBDs via melting of ice/snow within a season. The exact Cl requirements will depend on the amount of Cl in the CBDs, and on the ratio of catchment area to CBD volume. Figure 3a shows a matrix of outcomes when considering primary igneous apatite weathering as the source of Cl for the CBDs. While apatite dissolution is rapid, the regolith is mainly basaltic, and infiltrating fluids would be buffered to higher pH, which would decrease the dissolution rate of apatite (e.g., Adcock et al., 2013). Short durations of warming (τ) may not thaw a sufficient depth in the active layer to mobilize sufficient anions and cations. For medium durations of warming, fluids will be able to dissolve sufficient basalt to meet the cation and anion mass requirements of the CBDs, but at high W/R and low TWV, the mass of the ponding fluids would be insufficient and too dilute to form the CBDs. This scenario could correspond to the melting of a limited amount of near-surface ice. Low W/R would prevent high concentrations of S (relative to Cl) from reaching the lake, whereas at intermediate W/R, S in solution derived from dissolving primary sulfides in the basalt would form secondary minerals along the reaction path and prevent S-enriched brines from reaching the ponding site. At high W/R, however, S would not precipitate along the reaction path and would pond with Cl (along with carbonate if the W/R is high enough) (e.g., Eugster & Hardie, 1978) at the lake site. Therefore, this parameter combination is unlikely, because of the non-detection of sulfates and carbonates associated with the CBDs. 7 Figure 3. Possible sources of chlorine in the chloride-bearing deposits, showing permitted and excluded combinations of water-to-rock ratio (W/R; unitless), duration of the warming event(s) above the freezing point of water (τ, Mars years), and total water volume (TWV; kg H2O). Green areas are permitted scenarios, pink areas are excluded, and the orange area is disfavored but not excluded (see text). 8 2.2. Evaluating a volcanic source for chlorine Volcanic gases from the Noachian to the mid-Hesperian most likely contained HCl (M. L. Smith et al., 2014; e.g., Wänke et al., 1994). Cl sourced from volcanic gases could build up in the shallow regolith via (1) dry deposition of Cl-bearing molecules or mineral phases (subsequently remobilized by a fluid), and (2) wet deposition, in which chlorine-bearing gases are dissolved in atmospheric water, and then react with the shallow regolith. The mantle source regions of the shergottite-nakhlite-chassignite (SNC) martian meteorites would have contained 25 ± 8 ppm Cl and 56 ± 71 ppm H2O, i.e., they were on average drier, but halogen enriched compared to Earth (Filiberto et al., 2016; Filiberto & Treiman, 2009). (These results supersede previous estimates, which suggested 8 ppm HCl and 0.5 wt. % H2O were present in martian pre-eruptive basaltic magmas (Craddock & Greeley, 2009).) We use the estimated volatile content of the parental melt of the Shergotty meteorite to calculate the amount of chlorine degassed to the surface over time. Shergotty is a basalt derived from 10 – 15 % partial melting of a Light Rare Earth Element (LREE) enriched mantle region (Stolper & McSween, 1979) which contained 12 – 23 ppm Cl and 36 – 73 ppm H2O (McCubbin et al., 2016). The partial melt would have contained 363 – 484 ppm H2O and 116 – 155 ppm Cl (McCubbin et al., 2016), while the bulk meteorite contains ~ 108 ppm Cl and ~ 280 ppm H2O (Lodders, 1998). The solubility of volatiles in magma controls their release into the gas phase. For basalt melts containing ≤ 1 wt. % H2O, as was probably the case for Shergotty, the solubility of chloride is highly dependent on the concentration of cations in the melt able to complex with Cl (e.g., Al, Na, Ca, Mg) (Webster et al., 1999). Assuming negligible loss of cations during and after fractional crystallization of Shergotty, and whole rock composition reported by Lodders (1998), we calculate a Cl- solubility of 1.0 wt. % at 2 kbar and 0.75 wt. % at 1 bar. The value we use for chlorine concentration in the ~3.0 – 3.9 Gyr mantle is a lower limit because the 1.35 – 0.17 Gya (Nyquist et al., 2001) SNCs sample a young mantle and degassing has removed Cl from the mantle reservoir over time. Volcanic degassing is proportional to magma flux. Despite high early magma production rates, the volume of intrusive production in the crust could be a factor of 3 – 750 times the extrusive basalts (Black & Manga, 2016; Filiberto et al., 2014; Greeley & Schneid, 1991; Lillis et al., 2009). Gaseous HCl is the main carrier of Cl in volcanic gases (e.g., Aiuppa et al., 2009). HCl abundance is controlled by a poorly understood mechanism involving vapor- melt partitioning dependent on melt and fluid compositions, temperature, pressure, redox state of the melt, crystallization, partial melting and/or "open" versus "closed" modes of degassing (Aiuppa, 2009; Aiuppa et al., 2009; Edmonds et al., 2009; Métrich & Wallace, 2009). Absent a mechanistic understanding, our best guide is given by semi-empirical models using melt inclusion analyses of erupted terrestrial samples, solubility experiments, and S/Cl relationships observed in venting volcanic gases (Aiuppa, 2009; Edmonds et al., 2009; Pyle & Mather, 2009; Webster et al., 1999). We adopted separate molar partition coefficients from magma into the gas phase for Cl (DCl) for intrusive (DCl = 0 – 0.25) and extrusive (DCl = 0.9) bodies, consistent with published semi-empirical models (Appendix B). Figure 4 shows the calculated mass of accumulated Cl degassed on Mars over the range of ages of CBD-hosting terrain (3.96 to 3.0 Gyr ago (Osterloo et al., 2010)). We use the solubility of Cl in a Shergotty-like melt (see above), intrusive to extrusive ratios of 3 (maximal outgassed Cl) and 10 (minimal outgassed; consistent with intrusive martian gabbro NWA 6963 (Filiberto et al., 2014)), DCl appropriate for intrusives and extrusives, and a 9 crustal production model specific to Mars' composition of (Kiefer, 2016 pers. comm. Kiefer et al., 2015). Figure 4. Cumulative chlorine degassed by crust formation on Mars over the range of ages of CBD-hosting terrain (3.96 to 3.0 Gyr ago (Osterloo et al., 2010)). I/E = intrusive to extrusive ratio, DCl = gas-melt molar partition coefficient for chlorine. See section 2.2 and Appendix B for details. In the minimal degassing scenario (I/E = 10, and no Cl degassed from intrusive bodies), volcanic degassing between 3.96 and 3.0 Gyr ago could produce sufficient chlorine in < 1 Myr (Figure 4) to account for the mass of the observed global inventory of CBDs, assuming the volcanic Cl is all deposited in the catchments of the observed CBDs, which contain ~ 1.85 × 1013 kg Cl if the CBDs are 4 m thick (Hynek et al., 2015), cover a surface area of ~ 1.41 × 104 km2 (Osterloo et al., 2010) and contain 10 – 25 vol. % NaCl (density = 2165 kg m-3) (Glotch et al., 2016). At 3 Gyr ago, the cumulative minimal-scenario degassed Cl over the planet would have been ~ 300 kg/m2 (Figure 4). Figure 3b shows a matrix of outcomes when considering volcanic degassing as the source of chlorine, based on the hydrological parameters described at the beginning of section 2. For short durations of the warming event(s), volcanic volatiles deposited in the shallow regolith 10 could be mobilized easily by fluids, but cations in low concentration in volcanic gases (Na, K, Mg, Ca, Fe) required to ultimately form chloride salts would need to be sourced from surface rocks/soil. For a short warming event, low TWV coupled with low W/R could result from near surface melting of small amounts of ice reacting with near-surface volcanic Cl and local rock (e.g., seasonal melting). In this case, transport of the fluid and its solutes into a pond would not occur, since water-rock reaction would consume the small amount of water locally, and any residual water would freeze or evaporate at the end of the warming event. The melting of a large volume of subsurface pore ice could result in high TWV and low W/R. The high TWV would prevent the fluid from being completely consumed by reactions, but the low W/R fluids will contain large cation, Si and S concentrations from the dissolution at the rock-pore interface. If melting and draining occurs repeatedly, then the cation, Si and S concentration may exceed that demanded by the available volcanic Cl to form the CBDs, and sulfates or clay minerals may co-occur with chlorides. Both these outcomes are in tension with observations. For warm periods on the order of a year, at low W/R, one or a few wetting events with low TWV could cause the required fluid mass and composition to pond at topographic lows. (Sulfates could also precipitate far from the lakes.) On the other hand, a persistently warm/wet climate (high TWV, or many wetting events at low TWV) could potentially mobilize the soil in the regolith before Cl from volcanic degassing can accumulate in the soil, since volcanic Cl degassing is relatively slow (Figure 4). At high W/R, regardless of the TWV, sulfates and chlorides would co-occur at the site of evaporite precipitation. This is because high W/R reaction of surface soils/rocks would not only dissolve and transport the chlorine and sulfur-bearing phases deposited from volcanic gases, but also weather primary igneous sulfides in the near-subsurface, and transport S species downstream. Since the chloride deposits are not "associated" with sulfates (e.g., Osterloo et al., 2010), this scenario must be ruled out. The dissolution rate of basaltic chlorapatite and the amount of chlorapatite in the martian crust control the upper limit for the duration of the warming event(s) in a volcanic Cl-source scenario, since extended periods of liquid water availability could cause Cl derived from chlorapatite dissolution to dilute the signal of volcanically-derived Cl in the CBDs. Chlorapatite dissolves in 0.04 – 0.4 Mars years, corresponding to a weathering depth of 3.0 – 10 m (see section 2.1). This means that in a low-end Cl degassing scenario (I/E = 10, with only extrusive bodies degassing), crustal production on Mars in the timescale relevant to the CBDs produced on average 3.2 × 10-7 ± 2.1 × 10-7 kg Cl m-2 yr-1 (Figure 4), which would outweigh the Cl present in basaltic apatite in 7.9 – 13.8 Myr (4.3 – 7.5 × 106 Mars years) of volcanic degassing. This time is much shorter than the range of formation ages inferred for the CBDs (Osterloo et al., 2010). Furthermore, apatite dissolution would have been hindered as the pH of the fluid initially in equilibrium with the atmosphere was buffered by basalt. Therefore, the dissolution of primary minerals requires a more realistic approach, which we carry out with geochemical models (section 3). 2.3. Reworking (dissolution and transport) of previously formed evaporites. The low-elevation, low-latitude sulfate deposits of Mars have been suggested to be reworked Noachian evaporites (Milliken et al., 2009; Zolotov & Mironenko, 2016). However, reworking of pre-existing massive evaporites is unlikely to be the source of the Cl for the 11 observed CBDs. For example, the chlorides are often found in perched basins on high elevations, far from any plausible ancient marine basin. 2.4. Chlorine from the sky: Meteoritic delivery of chlorine? Readily soluble Cl-bearing phases might be supplied to the shallow regolith by meteorites. The observed global inventory of NaCl from the CBDs is 1.4 – 14 km3 if the CBDs are 1 – 4 m thick (Hynek et al., 2015) and 10 – 25 vol. % NaCl (Glotch et al., 2016). Rare H chondrites Zag and Monahans contain percent levels of extraterrestrial halite (Rubin et al., 2002) – the cumulative volume of Zag-like impactors delivering NaCl would have to be at least 28 – 563 km3 (for 5 vol % – 1 vol % NaCl in the impactors). This averages out to 0.03 – 0.11 kg Cl/m2 if it was delivered solely across Noachian and Hesperian highland terrains (6.79 × 107 km2 (Tanaka et al., 2014)). We believe meteoritic delivery (or "astrosedimentation"(Hesselbrock & Minton, 2017)) of Cl to form the CBDs is unlikely. One reason is the much greater efficiency of volcanism (section 2.2). Another is that the CBDs postdate basin-forming impacts on Mars (Robbins et al., 2013; Toon et al., 2010). Ongoing in-situ isotopic analyses of Mars Cl (Farley et al., 2016) might allow the meteoritic-source hypothesis to be tested. 3. Methods 3.1. Geochemical 1D flow-through and flush modeling To better understand the origin of the chloride-bearing deposits, we used reaction-transport modeling (e.g., van Berk & Fu, 2011; Bridges et al., 2015). Specifically, we used program CHIM-XPT (Reed, 1998) to compute simulations of geochemical weathering of the Mazatzal basalt analyzed by Spirit at Gusev crater with Alpha Particle X-Ray Spectrometer (APXS) and Mössbauer (Table 1) (McSween et al., 2004, 2006). Mazatzal is representative of the basaltic crustal composition of Mars (McSween et al., 2009). The model input composition we used was obtained from a brushed and abraded surface rock in which the interference of Mars dust and soil was minimized, was corrected for the sulfur and chlorine allocthonous to the rock, and is consistent with measured basaltic martian meteorite compositions (McSween et al., 2004, 2006). Additionally, our model composition input is insensitive to local elemental redistribution (basalts at the Gusev plains have experienced limited, isochemical weathering (Haskin et al., 2005; Hurowitz et al., 2006)) since our model input uses the whole rock bulk composition, rather than the calculated inferred mineralogy (Table 1). A detailed explanation of the modeling techniques is given by Reed (1998). We use the updated thermodynamic database SolthermBRGM, which includes data for low temperature geochemistry from the BRGM Thermoddem database (Blanc et al., 2012) among other sources. In the 1D flow-through model, water in equilibrium with a 60 mbar pCO2 + 0.15 mbar pO2 atmosphere (~ 10 × and 17 × the modern atmospheric CO2 and O2 partial pressure, respectively; Mahaffy et al. (2013)) permeates the atmosphere/rock interface and traverses the basaltic crust. The fluid equilibrates with the parcel of rock it is in contact with, but is out of equilibrium with the preceding parcels of rock (from which it has been fractionated). As the fluid moves to equilibrate with successive parcels of rock (Figure 5a), the fluid is consumed in water-rock reactions along the fluid path and the system evolves from high W/R to low W/R. The actual W/R of the inflowing fluid at the CBD lakes is unknown, so we vary the W/R in the model and compare the resulting model output to observed mineralogy. 12 Figure 5. Schematic of the two types of geochemical reaction-transport models used in our study. (a) 1D flow-through model used to study the basalt weathering hypothesis for the origin of the CBDs. Fresh water equilibrated with the atmosphere reacts and equilibrates with a parcel of basalt. Porosity of the rocks in the flow path changes as secondary minerals precipitate and rock is dissolved. Water is consumed in water-rock reactions along the path, decreasing W/R as the fluid moves through to the next parcel of fresh unreacted basalt. (b) Flush model used to study the volcanic origin hypothesis for the chlorine in the CBDs. Fresh water equilibrated with the atmosphere reacts with surface basalt + Cl and S from volcanic dry deposition. A proportion (a fixed percentage of the rock's porosity volume, here chosen as 10 vol. %) of fluid equilibrated with the rock exits the system. The same rock parcel is 13 reacted with fresh water in subsequent steps. Flushing forms alteration minerals and dissolves the rock, while altering the porosity of the system (allowing more or less water to enter in subsequent steps) and the composition of the exiting fluid. Flush models were used to constrain the total amount of water required to produce a fluid with a high Cl to S ratio, in accordance with observations that the chloride deposits are not associated with sulfates or other evaporites (e.g., Glotch et al., 2016; Osterloo et al., 2010). In flush models, a parcel of rock in contact with the atmosphere is flushed in consecutive "wetting events" by a fresh fluid equilibrated with the atmosphere (Figure 5b). The amount of fluid allowed to enter and leave the rock is controlled in the first "wetting event" by the desired W/R, and in subsequent events by the resulting porosity. We chose to allow 10 % pore volume space of the equilibrated fluid to exit the parcel of rock after each wetting event, and to infuse the rock with 10 % pore volume space of fresh fluid. Initial W/R ratios chosen were: 1, 5, 10, 50, 100, 500 and 1000. Porosity evolved in our flush models as minerals with different densities dissolved and precipitated, and solutes were removed from the system with fluid extracted at every step. 3.2. Depositional basin analyses Four CBD sites previously catalogued by Osterloo et al. (2010) were selected for this study: (1) the deposit studied by Hynek et al. (2015) west of Miyamoto Crater, (2) Terra Sirenum, (3) west of Knobel Crater, and (4) southeast of Bunnik Crater (Figure 1, Supplementary Table 1). Sites were selected on the basis of HiRISE stereopair coverage. At each site, CBDs were mapped using high-resolution orthorectified HiRISE image data (25 cm/pixel (McEwen et al., 2007)) and digital terrain models (DTMs; 1 m/pixel) produced by David P. Mayer using Ames Stereo Pipeline (ASP; (Moratto et al., 2010)) and the University of Chicago ASP scripts (Mayer & Kite, 2016). Additionally, orthorectified CTX images (~ 6 m/pixel (Malin et al., 2007)) and CTX DTMs (24 m/pixel) were used to manually delineate watersheds and the extent of the paleolakes encompassing the CBDs. The MOLA (D. E. Smith et al., 2001) gridded elevation product (~ 463 m/pixel) was used to determine maximum depths of the paleolakes where HiRISE and CTX DTM coverage was not available. We computed the thicknesses of the CBDs using a three-step procedure. First, we identified points in HiRISE orthoimages on top of the deposits and underneath the deposits, using craters excavating through the CBDs (as in Hynek et al. (2015)), erosional features, and the edges of the CBDs (Figure 6). Second, we subtracted the elevations of the bottom of the deposits from the top using the HiRISE DTMs to calculate local deposit thicknesses. Finally, we interpolated across irregularly-spaced thickness measurements using an inverse distance weighting function in ArcGIS in order to estimate thickness across each CBD. 14 Figure 6. Cartoon depicting chloride-bearing deposit measurements. Using HiRISE DTMs, the chloride-bearing deposit (wavy lines) thickness was calculated using erosional windows (green dotted line, e.g., craters) into the substrate below the chloride-bearing deposit, and at the edges of the deposit (green stars). The measured thickness points were interpolated (see section 3.2) to obtain an overall thickness. The minimum lake extent (red long dashed line) and volume was calculated based on the maximum height of the chloride-bearing deposit, and the maximum lake extent (blue short dashed line) and volume was calculated based on the pour-point elevation of the lake. Where CBDs within a basin lay outside our HiRISE DTM coverage (e.g., three HiRISE DTMs at Terra Sirenum), the mean of the interpolated thicknesses on each of the HiRISE DTMs was used to assign the thickness of the CBDs outside of the HiRISE DTMs. Average extents and thicknesses of the CBDs are reported in Table 2. Minimum and maximum lake volumes were calculated using the Cut/Fill tool on ArcGIS, where the lake bottom topography derived from the MOLA DTM (or a CTX DTM in the case of the CBD west of Miyamoto Crater) was subtracted from a constant elevation raster that defined the lake depth at (1) its minimum extent, from the maximum elevation of the CBD, and (2) at its maximum extent, from the maximum elevation of the lake overflow/tipping point (Figure 6). Calibrated daytime thermal infrared emission images from THEMIS (~ 100 m/pixel; Christensen et al., (2004)) were used to identify the extent of the CBDs. Where THEMIS multispectral data were available, decorrelation stretch (DCS) images were produced with spectral bands 8, 7, and 5 ("8/7/5"), 9/6/4 and 6/4/2 mapped to red, green and blue channels respectively , as described by Osterloo et al. (2010) to allow improved identification and mapping of CBDs. 4. Results 4.1. Geochemical models 4.1.1. 1D flow-through model for basalt weathering As the fluid reacted with larger masses of fresh basalt, the fluid composition evolved from high W/R to low W/R, and secondary minerals were precipitated along the reaction-transport path (Figure 7, and Supplementary Table 2 for details on specific mineral species). Cl concentration increased in the fluid with decreasing W/R as it was dissolved from the reactant basalt and not incorporated into secondary mineral precipitates. Cl/S ratios above the minimum Cl/S calculated to be in the CBDs (~ 11) occurred only at W/R ≤ 2.1. 15 The most noticeable change to the fluid composition occurred between W/R 6000 to 500, where most of the C in solution was consumed (Figure 7a) to form carbonates (Figure 7c) (siderite formed at W/R = 6000 – 2500, ankerite at W/R = 5000 – 700 and calcite at W/R = 700 – 350), thus binding cations (Fe and Ca, Figure 7b) leached from the reactant rock by the fluid at earlier stages (Figure 7a). At moderate to low W/R, the main mineral sinks for cations (Si, Al, Mn, Fe, Mg, Ca, K and Na) and water were phyllosilicates (kaolinite, montmorillonite, minnesotaite, chlorite, sepiolite, saponite and greenalite), and zeolites (clinoptilolite, phillipsite and chabazite) (Figure 7b-c). At lower W/R (W/R < 280) the cement calcium silicate hydrate (CSH) consumed Ca, silica and water, whereas at very low W/R (W/R < 10), hematite was the main stable Fe-bearing phase. Phosphate was precipitated out of the solution as hydrous phosphate minerals stable at low temperature (vivianite, MgHPO4, MnHPO4, and Ca4H(PO4)3·3H2O (octocalcium phosphate), a metastable precursor to authigenic apatite (Gunnars et al., 2004; Oxmann & Schwendenmann, 2015)). Sulfur was mainly taken up by pyrite, which lead to a fluid composition with increasing Cl/S at W/R ≤ 10 (Figure 7a). 16 Figure 7. Composition of the fluid along the 1D reaction-transport flow path through Mars basalt. Fluid initially equilibrated with the Mars atmosphere (see section 3.1) permeates the basalt and is consumed by water-rock reactions as it traverses through more fresh basalt, i.e., from high water-to-rock ratio (W/R) to low W/R. Note that in (a) and (b) the Y-axis shows molal concentration (moles ion/(kg H2O)), whereas in (c), the Y-axis shows the concentration in grams of a particular group of minerals per gram of added fresh basalt, normalized per kg of water. (At low W/R, more mineral mass is formed per gram of added reactant because the fluid contains high amounts of solutes.) For simplicity, species in solution (e.g., CaCl2, NaCl, HCl) are represented by the sum of their components (e.g., CaCl2 = Ca + 2Cl, NaCl = Na + Cl, HCl = H + Cl). (a) –pH, anion and silicon concentration in the solution. (b) Cation concentration in the solution. (c) Secondary minerals formed (see Supplementary Table 2 for the specific mineral species and formulae). 17 4.1.2. Flush model for volcanic deposit reworking W/R and TWV control the molar Cl/S concentration ratio of the fluid produced by flushing a rock of Mazatzal's composition (Figure 8). For W/R and TWV combinations that stabilize sulfates and sulfides, sulfur is fixed in mineral phases while the Cl was flushed out. Cl/S was always below the relevant ratio for the CBDs (~ 11) at initial W/R > 400, and with flushing by > 400 kg of H2O. Cl/S ≥ 11 occurred in the initial wetting events, when less than ~ 15 kg TWV had flushed through the basalt. Above ~ 15 kg TWV flushed, sulfur released from the dissolution of relatively insoluble pyrite decreased Cl/S in the solution. The larger abundance (a factor of ~ 2) of S (in pyrite) compared to Cl (in chlorapatite) in the host basalt (Table 1) resulted in the continued release of S into the fluid with continued flushing, while Cl was depleted. At initial W/R < 280, molar Cl/S in solution was initially high because thaumasite (Ca3Si(OH)6(CO3)(SO4)·12H2O) was stable in addition to pyrite, thereby fixing S while allowing Cl to stay in solution. After flushing >15 kg H2O through the basalt at an initial W/R < 280, thaumasite was entirely dissolved and could no longer act as a sink for S, resulting in a decreased Cl/S in the resulting alteration fluid. At low initial W/R (< 10), thaumasite stability increased with additional fresh water (equilibrated with the martian atmosphere) flushed through the basalt, increasing the Cl/S in the resulting pore and alteration fluid. Specifically, S dissolved from pyrite in the initial wetting event (but retained in the pore fluid) precipitated as thaumasite in subsequent wetting events. Thaumasite dissolved with further flushing of fresh water, thereby decreasing fluid Cl/S. A window of relatively high initial W/R and high TWV permitted Cl/S ≥ 11 (Figure 8). In this W/R and TWV regime, the precipitation of thermodynamically stable sulfide minerals captured S from the solution, thereby increasing the relative proportion of Cl in the solution. This peak in fluid Cl/S ratio occurred after flushing 150 – 170 kg TWV at moderate W/R (W/R = 300) and up to 250 – 370 kg TWV at low W/R (W/R = 1) (Figure 8). 18 Figure 8. Molar Cl/S ratio of the fluid resulting from alteration of Mars basalt containing Cl and S derived from volcanic degassing, as a function of total amount of water reacting with the basalt (TWV), and water-to-rock ratio. Bivariate interpolation was used for values that were not explicitly modeled in CHIM-XPT. The black line follows Cl/S = 11, which is the hard lower limit of the Cl/S ratio of the chloride-bearing deposits, based on volume estimations from spectroscopy and assuming any S present in the deposits is in gypsum (see text). Water-to-rock ratios and amounts of water resulting in fluids with Cl/S < 11 are excluded, so we can place constraints on paleohydrology (see Discussion). 4.1.3. Basin and chloride-bearing deposit physical parameters Table 2 shows the calculated thicknesses and extents of the CBDs studied here, the areas of the basins in which they are found, and the amount of chlorine in the watersheds that was required to form the deposits. All CBDs are light-toned, and are most easily identifiable where polygonal features (which have been interpreted as chloride-bearing desiccation cracks (Osterloo et al., 2008, 2010) probably containing phyllosilicates (El-Maarry et al., 2013, 2014)), occur on the surface e.g., Figure 9a. At the site near Miyamoto crater, we note the same contiguous CBD identified in the previous study by Hynek et al. (2015) in the local depression, as well as fluvial channels incising the north and east basin flanks leading towards the topographic low, and a lake pour 19 point into a large outflow channel in the south. However, our study of the extent, mass and basin area of the CBD differs from the previous study (Table 2). While Hynek et al. (2015) determined an average thickness of 4 m for the chloride-bearing deposit based on the compositional change of the material excavated by a crater near the center of the deposit, we note that the thickness of the chloride-bearing layer is laterally variable (as for other sites; Figure 9) from 0.3 to 4.4 m (average: 1.5 m). The slope of the deposit near Miyamoto crater is very flat (maximum < 1º from NNE to SSW), so the difference in thickness across the deposit is related to the topography of the material below the deposit. However, the appearance of the deposits at the studied sites differs: CBDs west of Knobel crater (Figure 9c) form non-contiguous raised ridges, so the variable thickness of the deposit (from 0.3 to 13.7 m, average: 8 m) may be related to post-depositional modification, or may be related to dune-like materials underneath indurated chlorides (El-Maarry et al., 2013). Terra Sirenum contained the largest chloride-bearing deposits we studied here (~ 9.5 × 108 m3) while the deposits west of Knobel crater were the smallest volumetrically (~ 4.2 × 106 m3). By multiplying the calculated deposit volumes by the volume of NaCl estimated to be in the deposits (10 – 25 %; Glotch et al. (2016)) and assuming an NaCl density of 2165 kg m-3, we calculated the mass of Cl in the deposits and the minimum concentration of Cl per catchment surface area required in the watersheds to form the deposits. The Cl requirements are minima because aeolian processes have eroded the CBDs. Deposit volumes did not vary linearly with watershed areas, so the required Cl concentrations varied considerably (Table 2). West of Knobel crater, the small CBDs are found in a large watershed, leading to low Cl concentrations required in the basin (0.16 – 0.40 kg Cl m-2), whereas the watershed at Terra Sirenum was only slightly larger, but its significantly more voluminous CBD required much higher Cl column abundance in the rocks throughout the basin (19.3 – 48.2 kg Cl m-2). The CBD lakes we have studied are distinct in comparison to the large crater lakes on Mars, but are also too deep to have been playa lakes. We have not identified sizeable and clear sub- lacustrine deltaic deposits at the end of numerous incising canyons at the CBD lakes, as found in the southern rim of Gale crater (Palucis et al., 2016) and southwestern Melas Chasma (Williams & Weitz, 2014). The spatial resolution of the orbital dataset precludes identifying small diagenetic features and evidence of late-stage alteration mineralogy (e.g., fracture-filled veins and concretions) as identified by the Mars Science Laboratory at Yellowknife Bay in Gale Crater (Bridges et al., 2015; Grotzinger et al., 2015; McLennan et al., 2014). Clear morphologic evidence of multievent stages of wetting (e.g., several lake stands) was not observed with orbital data sets at the CBD sites we studied, only a single maximum lake stand and a putative lower stand based on the height and extent of the CBDs we mapped were discernible at each of the sites (see the maximum and minimum lake extents defined in Figure 6). Lake spillover points are visible at the CBD sites near Miyamoto crater and at Terra Sirenum, and a sinuous raised ridge at Terra Sirenum containing chlorides and leading from SW to a major CBD in the NE is most likely an inverted inlet channel (Figure 9b). These features together with the depth of the lakes (Table 2) are more consistent with lacustrine settings rather than playa lake environments previously proposed (El-Maarry et al., 2013, 2014; Glotch et al., 2010; Osterloo et al., 2008, 2010; Ruesch et al., 2012). Outlet channels are also evident at the sites west of Knobel Crater and southeast of Bunnik Crater, although breaching points are not obvious. 20 Figure 9. Mapped chloride-bearing deposit sites. Deposit thicknesses (in m) are labelled. (a) Site near Miyamoto crater, previously analyzed by Hynek et al. (2015), (b) site at Terra Sirenum, (c) site west of Knobel Crater, (d) site southeast of Bunnik crater. 21 5. Discussion Osterloo et al. (2010) reported a total global inventory of ~ 1.4 × 104 km2 of CBDs on the surface of Mars. Assuming an average deposit thickness of 4 m (from our observations, Table 2, combining the three high-confidence sites west of Miyamoto Crater, west of Knobel Crater and Terra Sirenum), the deposits contain an equivalent of 1.2 × 1013 – 3.1 × 1013 kg NaCl (7.4 × 1012 – 1.9 × 1013 kg Cl) if the deposits contain 10 – 25 vol. % NaCl. Water volumes required to form the CBDs via groundwater alteration of near-surface basalt varied from site to site (Table 3), depending on the W/R of the fluid discharging into the lake and the amount of NaCl estimated to be in the deposits. We calculate a minimum of 3.1 × 108 m3 H2O (0.09 m water column) was required to form the chloride-bearing deposit West of Knobel crater assuming a concentrated fluid (5.09 × 10-2 moles Cl kg H2O-1 at W/R = 1) discharged at the ponding site and that the deposit contains 10 vol. % NaCl. The highest amount of water required to form the CBDs was calculated for Terra Sirenum, where, if the fluid was more dilute (2.31 × 10-2 moles Cl kg H2O-1 at W/R = 2) and the deposit contains 25 vol. % NaCl, 3.8 × 1011 m3 H2O (58.9 m water column) was necessary (Table 3). Similarly, the mass of rock required to be leached to produce the observed deposits was calculated, given the observed concentration of Cl in the Mars basalt we used as the source of Cl (0.15 wt. % Cl, see section 3.1 and Table 1). Translated across the watersheds, we calculated the depth to which the surface basalt needed to be weathered (Table 3) to obtain the Cl estimated to be in the CBDs. The depth of weathering ranged from 6 cm (West of Knobel crater) to 19.5 m (Terra Sirenum) across the basins (Table 3). Our calculated minimum times for top-down unfreezing of chlorapatite-bearing regolith (Table 3 and Figure 10) ranged from less than 1 sol (West of Knobel crater), to up to 1.7 Mars years (Terra Sirenum). 22 Figure 10. Range of weathering depths required at the chloride-bearing deposit sites to produce the deposits by weathering of the basalt in their respective basins. Weathering depths "W of Miyamoto C [1]" were calculated using the basin analyses by Hynek et al. (2015) and depths at "W of Miyamoto C [2]" were calculated with our own analyses. In light of these timescales, seasonal melting of ice or snow (Clow, 1987; Kite et al., 2013) could have been the source of water to leach Cl from basalt to form the deposits. The low TWV required (0.09 – 0.49 m water column) to leach the amount of Cl calculated to be in the CBDs West of Knobel crater is consistent with a short (< 1 Mars year) timescale for the wet event(s). Other sites that require tens of meters of water (e.g., up to 59 m water column at Terra Sirenum) would require decades to hundreds of Mars years of precipitation at low W/R (flushing at W/R > 400 consistently yields low Cl/S (Figure 8), inconsistent with the observed surface mineralogy) to produce the calculated volumes of chlorides. Similarly low W/R (< 2) are inferred from Alpine glacial meltwater transport of solutes (Brown et al., 1996; Fairchild, Killawee, Hubbard, et al., 1999; Fairchild, Killawee, Sharp, et al., 1999), where TWV may be high, but also in hypersaline pore fluids not directly sourced from precipitation or groundwater in Antarctic soils (Levy et al., 2012). W/R in sites of carbonate cement formation in shallow burial depths (< 100 m) by meteoric waters are typically too high (W/R > 10 (Meyers, 1989)). Paleolake depths from our DTMs range from 61 – 111 m (SE of Bunnik crater) to 122 – 223 m (Terra Sirenum). The conservative estimates are based on the topographic range of the chloride deposits, and (given that the chloride deposits are only a few meters thick) rule out a playa lake environment similar to that inferred at Meridiani Planum (McLennan et al., 2005). This depth is sufficient to protect the lake interiors from both UV light and galactic cosmic radiation (Hassler et al., 2014). Such deep lakes likely had lifetimes of decades or more (see 23 also Hynek et al., 2015). The net aridity (ratio of evaporation to precipitation) can be estimated from the ratio of lake area to catchment area (Matsubara et al., 2011); three of our four sites fall in the range between present-day Death Valley and present-day Western Nevada. Precipitation rates on early Mars were likely < 10 m/yr (Toon et al., 2010; Wordsworth, 2016), and the catchment-averaged rainfall/snowmelt columns for lake-filling range from 27 m to 195 m, so the lakes were unlikely to have filled in a single season. Further, the lakes could not have evaporated in a single season – because energetic limits on evaporation imply net evaporation rates < 1m/yr (Irwin et al., 2015). Perennial lakes can be sustained by latent heat import at annual average temperatures well below the freezing point, however (McKay et al., 1985), so the inference of deep, long-lived lakes is consistent with seasonal melting. Assuming an igneous chlorapatite source for the Cl, the concentration of Cl derived from the 1D flow-through model (5.09 × 10-2 moles Cl kg H2O-1 at W/R = 1; Figure 7) sets a lower bound on the mass of water pumped through the catchment. Assuming 10 % chloride content, and dividing by the areas of the catchments, we obtain:  > 1.9 m rainfall/snowmelt for West of Miyamoto crater (> 7.1 m using the catchment and CBD measurements from Hynek et al. (2015));  > 10.7 m rainfall/snowmelt for Terra Sirenum;  > 0.1 m rainfall/snowmelt for West of Knobel crater;  > 1.7 m rainfall/snowmelt for SE of Bunnik crater. Energetic limits on snowmelt (e.g., Kite et al., 2013) strongly imply that snowmelt rates are < 1 m/yr. These results imply minimum lake lifetimes of tens of years for Terra Sirenum, or less for other sites. These lake lifetime lower limits are shorter than those obtained for other Hesperian lakes from sediment transport considerations (Irwin et al., 2015; Palucis et al., 2016; Williams & Weitz, 2014). However, the geochemical constraint applies only to the minimum cumulative lifetime of the Cl dissolution/leaching events. Individual lake flooding events may not be recorded by the chloride-bearing deposits if the active layer is stripped off of Cl, e.g., in early warm events (Halevy et al., 2011), or if runoff reacted little with the regolith. Assuming by contrast a volcanic source for the Cl, all the data are satisfied by a single regolith-leaching event, forming saline lakes of modest duration (decades – thousands of years). Radiative forcing by a SO2-rich atmosphere, produced by punctuated higher than average outgassing rates concurring with volcanic paroxysm in the Late Noachian – Early Hesperian, has been argued to permit sustained surface T > 273 K for hundreds of years (Halevy & Head, 2014). Assuming instantaneous effusion rates of 105 – 106 m3/s (Halevy & Head, 2014), and melt Cl contents and extrusive degassing rates used here (see section 2.2), we calculate 8 × 10-3 – 0.1 kg Cl m-2 yr-1 supplied globally to the shallow regolith from extrusive paroxysm, which would satisfy the mass balance constraints of the CBDs in Terra Sirenum in < 6000 yr, and West of Knobel Crater in < 50 yr. This paroxysm is unrealistically rapid; however, Northern Plains lava flooding, or ongoing Tharsis volcanism, could each provide a sufficient degassing volume. 24 Because the chlorides postdate all basin-forming impacts on Mars, an impact-induced wet event is not expected (Toon et al., 2010) to produce enough precipitation (27 m – 195 m) to fill the lakes. While the inferred paleolake and CBD depths, and inflow and outflow channels containing chlorides suggest lacustrine, non-playa-like settings, the lack of submerged deltaic deposits and multiple lake stands suggests a decreased role for fluvial sediment transport compared to large crater lakes like Gale Crater (e.g., Palucis et al., 2016). Finally, we note that the low W/R required to produce the high Cl/S ratio brines from the aqueous alteration of Mars basalt, and Mars basalt + volcanic deposits, leads to the parallel formation of a phyllosilicate assemblage (Figure 7), namely saponite and chlorite minerals (Supplementary Table 2). Our model output is consistent with the presence of phyllosilicate minerals spatially associated with the CBDs. The spatial association of phyllosilicates with the CBDs has been established in ~ 30 % of the CBD sites studied previously (e.g., El- Maarry et al., 2013), and appears to support the development of desiccation cracks much like the fractured terrains seen at the CBD sites (El-Maarry et al., 2013, 2014). 6. Conclusions Combining constraints from physical (topographic) and chemical (mineralogic) observations increases the science value of both. Our study combines reaction transport modeling, orbital spectroscopy, and new volume estimates from high-resolution digital terrain models, in order to constrain the hydrologic boundary conditions for forming the chlorides. Considering a T = 0 °C system, we find:  Individual lakes lasted decades or longer, given the long evaporation times corresponding to their >100 m depths.  For a volcanic HCl source of chlorine, either the water-to-rock ratio, or the total water volume, or both, were probably not high. Violation of these limits would lead to precipitation of sulfates in the lakes, but sulfates are not observed in the lake deposits. Modest W/R and/or TWV is consistent with brief and/or arid climate excursions above the melting point, and consistent with punctuated high rates of volcanism.  For an igneous chlorapatite source of Cl, minimum duration of runoff was on the order of a decade.  Cl masses, divided by catchment area, give column densities 0.2 – 48 kg/m2, and these column densities bracket the expected chlorapatite-Cl content for a seasonally- warm active layer. Mining apatite from greater depths (deep-sourced groundwater) is not required.  Previous work has shown that for T < 0 °C, brine fractionation can separate Cl and S (Marion, 2001; e.g., Reeburgh & Springer-Young, 1983). Brine fractionation within the lakes is inconsistent with our observations, but brine fractionation of groundwater could immobilize S in the subsurface and yield Cl-enriched fluids to be concentrated in lakes.  Taken together, our results are consistent with Mars having a usually cold (Wordsworth, 2016), horizontally segregated hydrosphere (Gaidos & Marion, 2003; Head, 2012) by the time chlorides formed. Taliks (unfrozen zones) beneath the chloride lakes may have locally connected the surface and deep hydrosphere 25 (Andersen et al., 2002; Mikucki et al., 2015). Below-freezing temperatures or very limited W/R interaction are suggested (Ehlmann et al., 2011). Our results are consistent with pervasive leaching confined to a near-surface active layer and limited to a small percentage of Noachian-Hesperian geologic history. We rule out the following:  A wet, warm climate would leach S alongside Cl, contrary to the observed paucity of S minerals in the lakes. Either low T (groundwater brine fractionation), or low W/R, or low total water volume, can satisfy the "low S/Cl" constraint.  Playa lakes are ruled out, by the fairly constant thickness (a few m) of observed chloride-bearing deposits over an occasionally much greater topographic range, lake depths (> 100 m), and lake spillover features.  Because the chlorides postdate all basin-forming impacts on Mars, an impact-induced wet event is not expected (Toon et al., 2010) to produce enough precipitation (27 m – 195 m) to fill the lakes.  Deep weathering caused by infiltrating precipitation (corresponding to a prolonged warm climate (e.g., Andrews-Hanna & Lewis, 2011)) would be expected to produce thick chloride deposits, which are not observed. However, our results do not disfavor warm climates at other locations, or at earlier times. Appendix A: The duration of chlorapatite dissolution The total mass or total volume of chlorapatite in martian meteorites or regolith does not affect the dissolution rate of the chlorapatite per se – the mass of chlorapatite per surface area of the chlorapatite grains does. Under equal conditions, smaller grains dissolve more rapidly than larger grains even if the mass of the sum of small grains equals the mass of the sum of large grains, because smaller grains expose more mineral surface area than larger grains. The dissolution rate formula from Adcock et al. (2013) defines the dissolution rate (R) in moles/m2/s. To illustrate, the number of moles of apatite per surface area varies with grain size: Where r is the radius of the spherical apatite grain. For example, for a hypothetical chlorapatite grain size diameter of 1 cm, the surface area of the spherical grain is 3.14×10-4 m2, with a volume of 5.24×10-7 m3. The average density of chlorapatite is 3175 kg/m3, so each kg of chlorapatite would contain 602 spherical grains (1 cm diameter each). The total surface area of 1 cm diameter spherical grains adding up to 1 kg of chlorapatite would be ~ 0.19 m2/kg. The molar mass of chlorapatite is 523.8 g/mol, so the surface area per mole of chlorapatite is ~ 9.90×10-2 m2/mol, or ~ 10.1 moles of chlorapatite per m2. For a calculated dissolution rate of 4.22×10-9 moles apatite/m2/s from Adcock et al. (2013) at a pH = 4.5 (for water in equilibrium with a pCO2 = 60 mbar), 1 cm diameter grains of chlorapatite would take 2.4×109 s to dissolve, or > 40 Mars years. 26 apatite kg5238.0apatite molapatitem4apatitegrain apatitegrain apatitem34apatitemapatite kg317522333rr Appendix B: Determination of the gas-melt partition coefficient of Cl (DCl). At low crustal pressures, the abundance of HCl released into gases from a magma appears to be strongly dependent on the closed- vs. open-system behavior of degassing, leading Cl to be partitioned into Cl-bearing minerals (apatite, amphiboles, ...) in open systems, and removed from the melt as gas or exsolved fluid in closed, slow-cooling systems (Aiuppa et al., 2009). The effect of confining pressure on Cl degassing from a melt is either not well constrained or highly variable: Edmonds et al. (2009) found that Cl degassing was inefficient for intraplate basaltic melts in the Kīlauea Volcano, Hawai'i at pressures > 10 bar, whereas basalt melts in subocean ridges and rifts appear to degas at ≲ 50 bar, regardless of whether Cl is supersaturated (Schilling et al., 1980). Black et al. (2012) noted, however, that since the gas/melt partitioning coefficient of Cl (DCl) increased with Cl concentration in the melt (Webster et al., 1999), low Cl degassing efficiency from the Kīlauea Volcano (Edmonds et al., 2009) reflected low Cl concentration in the parental melt. For the higher Cl content of Siberian Trap basalts, Black et al. (2012) used degassing efficiencies from other terrestrial flood basalt provinces to impose a 25 % Cl degassing efficiency from intrusive melts, and 36 – 63 % Cl degassing from extrusive melts. Sulfur-rich magmas tend to favor HCl partitioning into fluids in closed systems; in openly degassing systems (i.e., gas bubbles segregate from the melt; Métrich and Wallace (2009)), sulfur tends to be less soluble than Cl in melts, so S/Cl ratios decrease in the gas phase as degassing progresses (Aiuppa et al., 2002). At its simplest, S/Cl in the gas phase of openly degassing systems can be described as follows: (1) where is the resulting molar ratio of S/Cl in the gas phase, is the initial molar ratio in the parental magma, Ds and DCl are the gas-melt molar partition coefficients for S and Cl respectively, and RS is the residual molar fraction of S in the melt (from 1 initially, to 0 where all S has been degassed) (Aiuppa et al., 2002). The average ratio of partition coefficients (DS/DCl = 9) has been empirically determined from terrestrial volcanoes (Aiuppa, 2009). However, DCl varies strongly with pressure and open- versus closed-system degassing behavior (Edmonds et al., 2009), and experimental determination of Cl solubility in basaltic melts revealed a DCl range of 0.9 – 6 (Webster et al., 1999). In our model we do not take into account the coupled H2O fraction in the melt and its partition into the crystallizing melt as in Edmonds et al. (2009). For the determination of the amount of Cl degassed over time on Mars (Figure 4), we imposed a minimum degassing of 0 for intrusive bodies, and a maximum DCl of 0.25 (consistent with degassing from sills in the Siberian Traps (Black et al., 2012)). For extrusive bodies, we impose a DCl of 0.9, consistent with experimental solubility data (Webster et al., 1999). Acknowledgements We thank David P. Mayer for producing the HiRISE and CTX DTMs, for helpful discussions, and for help with GIS. We are grateful for comments and suggestions from 27 SCl1SClSmeltparent gasClSClSDDRDDgasClSmeltparent ClS reviewers John Bridges and M. R. El-Maarry, which improved the manuscript. We thank Walter Kiefer, Timothy Glotch and Cheng Ye for data that helped shape this work, and we also acknowledge discussions with and insights from Justin Filiberto, Benjamin Black, Mark Reed, Caleb Fassett, Thomas Bristow, Jim Palandri, Tim Bowling and Monica Grady. We acknowledge the use of University of Chicago Research Computing Center computing resources ("Midway" cluster). The data used are listed in the references, tables and Supplementary Information. This work was supported by NASA grant NNX16AG55G. References Adcock, C. T., Hausrath, E. M., & Forster, P. M. (2013). Readily available phosphate from minerals in early aqueous environments on Mars. Nature Geoscience, 6(10), 824–827. https://doi.org/10.1038/ngeo1923 Aiuppa, A. (2009). Degassing of halogens from basaltic volcanism: Insights from volcanic gas observations. Chemical Geology, 263(1–4), 99–109. https://doi.org/10.1016/j.chemgeo.2008.08.022 Aiuppa, A., Federico, C., Paonita, A., Pecoraino, G., & Valenza, M. (2002). S, Cl and F degassing as an indicator of volcanic dynamics: The 2001 eruption of Mount Etna. Geophysical Research Letters, 29(11), 54–1–54–4. https://doi.org/10.1029/2002GL015032 Aiuppa, A., Baker, D. R., & Webster, J. D. (2009). Halogens in volcanic systems. Chemical Geology, 263(1–4), 1–18. https://doi.org/10.1016/j.chemgeo.2008.10.005 Andersen, D. T., Pollard, W. H., McKay, C. P., & Heldmann, J. (2002). Cold springs in permafrost on Earth and Mars. Journal of Geophysical Research: Planets, 107(E3), 4–1. https://doi.org/10.1029/2000JE001436 Andrews-Hanna, J. C., & Lewis, K. W. (2011). Early Mars hydrology: 2. Hydrological evolution in the Noachian and Hesperian epochs. Journal of Geophysical Research: Planets, 116(E2), E02007. https://doi.org/10.1029/2010JE003709 van Berk, W., & Fu, Y. (2011). Reproducing hydrogeochemical conditions triggering the formation of carbonate and phyllosilicate alteration mineral assemblages on Mars (Nili Fossae region). Journal of Geophysical Research: Planets, 116(E10), E10006. https://doi.org/10.1029/2011JE003886 28 Black, B. A., & Manga, M. (2016). The eruptibility of magmas at Tharsis and Syrtis Major on Mars. Journal of Geophysical Research: Planets, 121(6), 944–964. https://doi.org/10.1002/2016JE004998 Black, B. A., Elkins-Tanton, L. T., Rowe, M. C., & Peate, I. U. (2012). Magnitude and consequences of volatile release from the Siberian Traps. Earth and Planetary Science Letters, 317–318, 363–373. https://doi.org/10.1016/j.epsl.2011.12.001 Blanc, P., Lassin, A., Piantone, P., Azaroual, M., Jacquemet, N., Fabbri, A., & Gaucher, E. C. (2012). Thermoddem: A geochemical database focused on low temperature water/rock interactions and waste materials. Applied Geochemistry, 27(10), 2107–2116. https://doi.org/10.1016/j.apgeochem.2012.06.002 Bridges, J. C., Schwenzer, S. P., Leveille, R., Westall, F., Wiens, R. C., Mangold, N., … Berger, G. (2015). Diagenesis and clay mineral formation at Gale Crater, Mars: Gale Crater Diagenesis. Journal of Geophysical Research: Planets, 120(1), 1–19. https://doi.org/10.1002/2014JE004757 Brown, G. H., Tranter, M., & Sharp, M. J. (1996). Experimental investigations of the weathering of suspended sediment by Alpine glacial meltwater. Hydrological Processes, 10(4), 579–597. https://doi.org/10.1002/(SICI)1099-1085(199604)10:4<579::AID-HYP393>3.0.CO;2-D Burt, D. M., & Knauth, L. P. (2003). Electrically conducting, Ca-rich brines, rather than water, expected in the Martian subsurface. Journal of Geophysical Research: Planets, 108(E4), 8026. https://doi.org/10.1029/2002JE001862 Christensen, P. R., Jakosky, B. M., Kieffer, H. H., Malin, M. C., McSween, H. Y., Nealson, K., … Ravine, M. (2004). The Thermal Emission Imaging System (THEMIS) for the Mars 2001 Odyssey Mission. Space Science Reviews, 110(1–2), 85–130. https://doi.org/10.1023/B:SPAC.0000021008.16305.94 Clark, B. C., Morris, R. V., McLennan, S. M., Gellert, R., Jolliff, B., Knoll, A. H., … Rieder, R. (2005). Chemistry and mineralogy of outcrops at Meridiani Planum. Earth and Planetary Science Letters, 240(1), 73–94. https://doi.org/10.1016/j.epsl.2005.09.040 29 Clow, G. D. (1987). Generation of liquid water on Mars through the melting of a dusty snowpack. Icarus, 72(1), 95–127. https://doi.org/10.1016/0019-1035(87)90123-0 Craddock, R. A., & Greeley, R. (2009). Minimum estimates of the amount and timing of gases released into the martian atmosphere from volcanic eruptions. Icarus, 204(2), 512–526. https://doi.org/10.1016/j.icarus.2009.07.026 Edmonds, M., Gerlach, T. M., & Herd, R. A. (2009). Halogen degassing during ascent and eruption of water-poor basaltic magma. Chemical Geology, 263(1–4), 122–130. https://doi.org/10.1016/j.chemgeo.2008.09.022 Ehlmann, B. L., Mustard, J. F., Murchie, S. L., Bibring, J.-P., Meunier, A., Fraeman, A. A., & Langevin, Y. (2011). Subsurface water and clay mineral formation during the early history of Mars. Nature, 479(7371), 53–60. https://doi.org/10.1038/nature10582 El-Maarry, M. R., Pommerol, A., & Thomas, N. (2013). Analysis of polygonal cracking patterns in chloride-bearing terrains on Mars: Indicators of ancient playa settings. Journal of Geophysical Research: Planets, 118(11), 2263–2278. https://doi.org/10.1002/2013JE004463 El-Maarry, M. R., Watters, W., McKeown, N. K., Carter, J., Noe Dobrea, E., Bishop, J. L., … Thomas, N. (2014). Potential desiccation cracks on Mars: A synthesis from modeling, analogue-field studies, and global observations. Icarus, 241, 248–268. https://doi.org/10.1016/j.icarus.2014.06.033 Eugster, H. P. (1980). Geochemistry of Evaporitic Lacustrine Deposits. Annual Review of Earth and Planetary Sciences, 8, 35. https://doi.org/10.1146/annurev.ea.08.050180.000343 Eugster, H. P., & Hardie, L. A. (1978). Saline Lakes. In A. Lerman (Ed.), Lakes: Chemistry, Geology, Physics (pp. 237–293). New York, NY: Springer New York. Retrieved from http://dx.doi.org/10.1007/978-1-4757-1152-3_8 Fairchild, I. J., Killawee, J. A., Sharp, M. J., Spiro, B., Hubbard, B., Lorrain, R. D., & Tison, J.-L. (1999). Solute generation and transfer from a chemically reactive alpine glacial–proglacial system. Earth Surface Processes and Landforms, 24(13), 1189–1211. https://doi.org/10.1002/(SICI)1096-9837(199912)24:13<1189::AID-ESP31>3.0.CO;2-P 30 Fairchild, I. J., Killawee, J. A., Hubbard, B., & Wolfgang Dreybrodt. (1999). Interactions of calcareous suspended sediment with glacial meltwater: a field test of dissolution behaviour. Chemical Geology, 155(3–4), 243–263. https://doi.org/10.1016/S0009-2541(98)00170-3 Farley, K. A., Martin, P., Archer Jr., P. D., Atreya, S. K., Conrad, P. G., Eigenbrode, J. L., … Sutter, B. (2016). Light and variable 37Cl/35Cl ratios in rocks from Gale Crater, Mars: Possible signature of perchlorate. Earth and Planetary Science Letters, 438, 14–24. https://doi.org/10.1016/j.epsl.2015.12.013 Fassett, C. I., & Head, J. W. (2008). The timing of martian valley network activity: Constraints from buffered crater counting. Icarus, 195(1), 61–89. https://doi.org/10.1016/j.icarus.2007.12.009 Fassett, C. I., & Head, J. W. (2011). Sequence and timing of conditions on early Mars. Icarus, 211(2), 1204–1214. https://doi.org/10.1016/j.icarus.2010.11.014 Filiberto, J., Gross, J., Trela, J., & Ferré, E. C. (2014). Gabbroic Shergottite Northwest Africa 6963: An intrusive sample of Mars. American Mineralogist, 99(4), 601. https://doi.org/10.2138/am.2014.4638 Filiberto, J., Gross, J., & McCubbin, F. M. (2016). Constraints on the water, chlorine, and fluorine content of the Martian mantle. Meteoritics & Planetary Science, 51(11), 2023–2035. https://doi.org/10.1111/maps.12624 Filiberto, J., & Treiman, A. H. (2009). Martian magmas contained abundant chlorine, but little water. Geology, 37(12), 1087–1090. https://doi.org/10.1130/G30488A.1 Gaidos, E., & Marion, G. (2003). Geological and geochemical legacy of a cold early Mars. Journal of Geophysical Research: Planets, 108(E6), 5055. https://doi.org/10.1029/2002JE002000 Glotch, T. D., Bandfield, J. L., Wolff, M. J., Arnold, J. A., & Che, C. (2016). Constraints on the composition and particle size of chloride salt-bearing deposits on Mars. Journal of Geophysical Research: Planets. https://doi.org/10.1002/2015JE004921 Glotch, T. D., Bandfield, J. L., Tornabene, L. L., Jensen, H. B., & Seelos, F. P. (2010). Distribution and formation of chlorides and phyllosilicates in Terra Sirenum, Mars. Geophysical Research Letters, 37(16), L16202. https://doi.org/10.1029/2010GL044557 31 Greeley, R., & Schneid, B. D. (1991). Magma Generation on Mars: Amounts, Rates, and Comparisons with Earth, Moon, and Venus. Science, 254(5034), 996–998. Grotzinger, J. P., Gupta, S., Malin, M. C., Rubin, D. M., Schieber, J., Siebach, K., … Wilson, S. A. (2015). Deposition, exhumation, and paleoclimate of an ancient lake deposit, Gale crater, Mars. Science, 350(6257). https://doi.org/10.1126/science.aac7575 Guidry, M. W., & Mackenzie, F. T. (2003). Experimental study of igneous and sedimentary apatite dissolution. Geochimica et Cosmochimica Acta, 67(16), 2949–2963. https://doi.org/10.1016/S0016-7037(03)00265-5 Gunnars, A., Blomqvist, S., & Martinsson, C. (2004). Inorganic formation of apatite in brackish seawater from the Baltic Sea: an experimental approach. Marine Chemistry, 91(1–4), 15–26. https://doi.org/10.1016/j.marchem.2004.01.008 Halevy, I., Fischer, W. W., & Eiler, J. M. (2011). Carbonates in the Martian meteorite Allan Hills 84001 formed at 18 ± 4 °C in a near-surface aqueous environment. Proceedings of the National Academy of Sciences, 108(41), 16895–16899. https://doi.org/10.1073/pnas.1109444108 Halevy, I., & Head, J. W. (2014). Episodic warming of early Mars by punctuated volcanism. Nature Geosci, 7(12), 865–868. https://doi.org/doi:10.1038/ngeo2293 Haskin, L. A., Wang, A., Jolliff, B. L., McSween, H. Y., Clark, B. C., Des Marais, D. J., … Soderblom, L. (2005). Water alteration of rocks and soils on Mars at the Spirit rover site in Gusev crater. Nature, 436(7047), 66–69. https://doi.org/10.1038/nature03640 Hassler, D. M., Zeitlin, C., Wimmer-Schweingruber, R. F., Ehresmann, B., Rafkin, S., Eigenbrode, J. L., … MSL Science Team. (2014). Mars' Surface Radiation Environment Measured with the Mars Science Laboratory's Curiosity Rover. Science, 343(6169). https://doi.org/10.1126/science.1244797 Head, J. W. (2012). Mars Planetary Hydrology: Was the Martian Hydrological Cycle and System Ever Globally Vertically Integrated? In Lunar and Planetary Science Conference (Vol. 43, p. 2137). Houston, TX: Lunar and Planetary Institute. 32 Herkenhoff, K. E., Golombek, M. P., Guinness, E. A., Johnson, J. B., Kusack, A., Richter, L., … Gorevan, S. (2008). In situ observations of the physical properties of the Martian surface. In J. Bell III (Ed.), The Martian Surface - Composition, Mineralogy, and Physical Properties (p. 451). Hesselbrock, A. J., & Minton, D. A. (2017). An ongoing satellite-ring cycle of Mars and the origins of Phobos and Deimos. Nature Geoscience, 10(4), 266–269. https://doi.org/doi:10.1038/ngeo2916 Hurowitz, J. A., McLennan, S. M., Tosca, N. J., Arvidson, R. E., Michalski, J. R., Ming, D. W., … Squyres, S. W. (2006). In situ and experimental evidence for acidic weathering of rocks and soils on Mars. Journal of Geophysical Research: Planets, 111(E2), E02S19. https://doi.org/10.1029/2005JE002515 Hynek, B. M., Osterloo, M. K., & Kierein-Young, K. S. (2015). Late-stage formation of Martian chloride salts through ponding and evaporation. Geology, 43(9), 787–790. https://doi.org/10.1130/G36895.1 Irwin, R. P., Lewis, K. W., Howard, A. D., & Grant, J. A. (2015). Paleohydrology of Eberswalde crater, Mars. Geomorphology, 240, 83–101. https://doi.org/10.1016/j.geomorph.2014.10.012 Keller, J. M., Boynton, W. V., Karunatillake, S., Baker, V. R., Dohm, J. M., Evans, L. G., … Williams, R. M. S. (2007). Equatorial and midlatitude distribution of chlorine measured by Mars Odyssey GRS. Journal of Geophysical Research, 112(E3). https://doi.org/10.1029/2006JE002679 Kerber, L., Head, J. W., Madeleine, J.-B., Forget, F., & Wilson, L. (2012). The dispersal of pyroclasts from ancient explosive volcanoes on Mars: Implications for the friable layered deposits. Icarus, 219(1), 358–381. https://doi.org/10.1016/j.icarus.2012.03.016 Kerber, L., Forget, F., Madeleine, J.-B., Wordsworth, R., Head, J. W., & Wilson, L. (2013). The effect of atmospheric pressure on the dispersal of pyroclasts from martian volcanoes. Icarus, 223(1), 149–156. https://doi.org/10.1016/j.icarus.2012.11.037 Kiefer, W. S. (2016, November 4). Mars Crustal production from mid-Noachian to mid-Hesperian. 33 Kiefer, W. S., Filiberto, J., Sandu, C., & Li, Q. (2015). The effects of mantle composition on the peridotite solidus: Implications for the magmatic history of Mars. Geochimica et Cosmochimica Acta, 162, 247–258. https://doi.org/10.1016/j.gca.2015.02.010 Kite, E. S., Halevy, I., Kahre, M. A., Wolff, M. J., & Manga, M. (2013). Seasonal melting and the formation of sedimentary rocks on Mars, with predictions for the Gale Crater mound. Icarus, 223(1), 181–210. https://doi.org/10.1016/j.icarus.2012.11.034 Levy, J. S., Fountain, A. G., Welch, K. A., & Lyons, W. B. (2012). Hypersaline "wet patches" in Taylor Valley, Antarctica. Geophysical Research Letters, 39(5), L05402. https://doi.org/10.1029/2012GL050898 Lillis, R. J., Dufek, J., Bleacher, J. E., & Manga, M. (2009). Demagnetization of crust by magmatic intrusion near the Arsia Mons volcano: Magnetic and thermal implications for the development of the Tharsis province, Mars. Journal of Volcanology and Geothermal Research, 185(1–2), 123–138. https://doi.org/10.1016/j.jvolgeores.2008.12.007 Lodders, K. (1998). A survey of shergottite, nakhlite and chassigny meteorites whole-rock compositions. Meteoritics & Planetary Science, 33(S4), A183–A190. https://doi.org/10.1111/j.1945-5100.1998.tb01331.x Mahaffy, P. R., Webster, C. R., Atreya, S. K., Franz, H., Wong, M., Conrad, P. G., … MSL Science Team. (2013). Abundance and Isotopic Composition of Gases in the Martian Atmosphere from the Curiosity Rover. Science, 341(6143), 263–266. https://doi.org/10.1126/science.1237966 Mahaffy, P. R., Webster, C. R., Stern, J. C., Brunner, A. E., Atreya, S. K., Conrad, P. G., … Wray, J. J. (2015). The imprint of atmospheric evolution in the D/H of Hesperian clay minerals on Mars. Science, 347(6220), 412. https://doi.org/10.1126/science.1260291 Malin, M. C., Bell, J. F., Cantor, B. A., Caplinger, M. A., Calvin, W. M., Clancy, R. T., … Wolff, M. J. (2007). Context Camera Investigation on board the Mars Reconnaissance Orbiter. Journal of Geophysical Research: Planets, 112(E5), E05S04. https://doi.org/10.1029/2006JE002808 34 Marion, G. M. (2001). Carbonate mineral solubility at low temperatures in the Na-K-Mg-Ca-H-Cl- SO4-OH-HCO3-CO3-CO2-H2O system. Geochimica et Cosmochimica Acta, 65(12), 1883– 1896. https://doi.org/10.1016/S0016-7037(00)00588-3 Matsubara, Y., Howard, A. D., & Drummond, S. A. (2011). Hydrology of early Mars: Lake basins. Journal of Geophysical Research: Planets, 116(E4), E04001. https://doi.org/10.1029/2010JE003739 Mayer, D. P., & Kite, E. S. (2016). An Integrated Workflow for Producing Digital Terrain Models of Mars from CTX and HiRISE Stereo Data Using the NASA Ames Stereo Pipeline. In Lunar and Planetary Science Conference (Vol. 47, p. 1241). Houston, TX: Lunar and Planetary Institute. Retrieved from http://www.lpi.usra.edu/meetings/lpsc2016/pdf/1241.pdf McCubbin, F. M., Boyce, J. W., Srinivasan, P., Santos, A. R., Elardo, S. M., Filiberto, J., … Shearer, C. K. (2016). Heterogeneous distribution of H2O in the Martian interior: Implications for the abundance of H2O in depleted and enriched mantle sources. Meteoritics & Planetary Science, 51(11), 2036–2060. https://doi.org/10.1111/maps.12639 McEwen, A. S., Eliason, E. M., Bergstrom, J. W., Bridges, N. T., Hansen, C. J., Delamere, W. A., … Weitz, C. M. (2007). Mars reconnaissance orbiter's high resolution imaging science experiment (HiRISE). Journal of Geophysical Research E: Planets, 112(5). https://doi.org/10.1029/2005JE002605 McKay, C. P., Clow, G. D., Wharton, R. A., & Squyres, S. W. (1985). Thickness of ice on perennially frozen lakes. Nature, 313(6003), 561–562. https://doi.org/10.1038/313561a0 McLennan, S. M., Anderson, R. B., Bell, J. F., Bridges, J. C., Calef, F., Campbell, J. L., … MSL Science Team. (2014). Elemental Geochemistry of Sedimentary Rocks at Yellowknife Bay, Gale Crater, Mars. Science, 343(6169). https://doi.org/10.1126/science.1244734 McLennan, S. M., Bell III, J. F., Calvin, W. M., Christensen, P. R., Clark, B. C., de Souza, P. A., … Yen, A. (2005). Provenance and diagenesis of the evaporite-bearing Burns formation, Meridiani Planum, Mars. Earth and Planetary Science Letters, 240(1), 95–121. https://doi.org/10.1016/j.epsl.2005.09.041 35 McSween, H. Y., Taylor, G. J., & Wyatt, M. B. (2009). Elemental Composition of the Martian Crust. Science, 324(5928), 736. https://doi.org/10.1126/science.1165871 McSween, H. Y., Arvidson, R. E., Bell, J. F., Blaney, D., Cabrol, N. A., Christensen, P. R., … Zipfel, J. (2004). Basaltic Rocks Analyzed by the Spirit Rover in Gusev Crater. Science, 305(5685), 842. https://doi.org/10.1126/science.3050842 McSween, H. Y., Wyatt, M. B., Gellert, R., Bell, J. F., Morris, R. V., Herkenhoff, K. E., … Zipfel, J. (2006). Characterization and petrologic interpretation of olivine-rich basalts at Gusev Crater, Mars. Journal of Geophysical Research: Planets, 111(E2), E02S10. https://doi.org/10.1029/2005JE002477 Mellon, M. T., Fergason, R. L., & Putzig, N. E. (2008). The thermal inertia of the surface of Mars. In J. Bell (Ed.), The Martian Surface (pp. 399–427). Cambridge, UK: Cambridge University Press. https://doi.org/10.1017/CBO9780511536076.019 Métrich, N., & Wallace, P. J. (2009). Volatile Abundances in Basaltic Magmas and Their Degassing Paths Tracked by Melt Inclusions. Reviews in Mineralogy and Geochemistry, 69(1), 363. https://doi.org/10.2138/rmg.2008.69.10 Meyers, W. J. (1989). Trace element and isotope geochemistry of zoned calcite cements, Lake Valley Formation (Mississippian, New Mexico): insights from water-rock interaction modelling. Sedimentary Geology, 65(3), 355–370. https://doi.org/10.1016/0037-0738(89)90034-1 Mikucki, J. A., Auken, E., Tulaczyk, S., Virginia, R. A., Schamper, C., Sørensen, K. I., … Foley, N. (2015). Deep groundwater and potential subsurface habitats beneath an Antarctic dry valley. Nature Communications, 6, 6831. https://doi.org/doi:10.1038/ncomms7831 Milliken, R. E., Fischer, W. W., & Hurowitz, J. A. (2009). Missing salts on early Mars. Geophysical Research Letters, 36(11), L11202. https://doi.org/10.1029/2009GL038558 Moratto, Z. M., Broxton, M. J., Beyer, R. A., Lundy, M., & Husmann, K. (2010). Ames Stereo Pipeline, NASA's Open Source Automated Stereogrammetry Software. In Lunar and Planetary Science Conference (Vol. 41, p. 2364). Houston, TX: Lunar and Planetary Institute. Retrieved from http://adsabs.harvard.edu/abs/2010LPI....41.2364M 36 Murchie, S. L., Mustard, J. F., Ehlmann, B. L., Milliken, R. E., Bishop, J. L., McKeown, N. K., … Bibring, J.-P. (2009). A synthesis of Martian aqueous mineralogy after 1 Mars year of observations from the Mars Reconnaissance Orbiter. Journal of Geophysical Research: Planets, 114(E2), E00D06. https://doi.org/10.1029/2009JE003342 Nimmo, F., & Tanaka, K. (2005). Early crustal evolution of Mars. Annual Review of Earth and Planetary Sciences, 33(1), 133–161. https://doi.org/10.1146/annurev.earth.33.092203.122637 Nyquist, L. E., Bogard, D. D., Shih, C.-Y., Greshake, A., Stöffler, D., & Eugster, O. (2001). Ages and Geologic Histories of Martian Meteorites. In Chronology and Evolution of Mars (pp. 105– 164). Springer, Dordrecht. https://doi.org/10.1007/978-94-017-1035-0_5 Osterloo, M. M., Hamilton, V. E., Bandfield, J. L., Glotch, T. D., Baldridge, A. M., Christensen, P. R., … Anderson, F. S. (2008). Chloride-Bearing Materials in the Southern Highlands of Mars. Science, 319(5870), 1651–1654. https://doi.org/10.1126/science.1150690 Osterloo, M. M., Anderson, F. S., Hamilton, V. E., & Hynek, B. M. (2010). Geologic context of proposed chloride-bearing materials on Mars. Journal of Geophysical Research, 115(E10), E10012. https://doi.org/10.1029/2010JE003613 Oxmann, J. F., & Schwendenmann, L. (2015). Authigenic apatite and octacalcium phosphate formation due to adsorption–precipitation switching across estuarine salinity gradients. Biogeosciences, 12(3), 723–738. https://doi.org/10.5194/bg-12-723-2015 Palucis, M. C., Dietrich, W. E., Williams, R. M. E., Hayes, A. G., Parker, T., Sumner, D. Y., … Newsom, H. (2016). Sequence and relative timing of large lakes in Gale crater (Mars) after the formation of Mount Sharp. Journal of Geophysical Research: Planets, 121(3), 472–496. https://doi.org/10.1002/2015JE004905 Pyle, D. M., & Mather, T. A. (2009). Halogens in igneous processes and their fluxes to the atmosphere and oceans from volcanic activity: A review. Chemical Geology, 263(1–4), 110– 121. https://doi.org/10.1016/j.chemgeo.2008.11.013 37 Reeburgh, W. S., & Springer-Young, M. (1983). New measurements of sulfate and chlorinity in natural sea ice. Journal of Geophysical Research: Oceans, 88(C5), 2959–2966. https://doi.org/10.1029/JC088iC05p02959 Reed, M. H. (1997). Hydrothermal alteration and its relationship to ore fluid composition. In H. L. Barnes (Ed.), Geochemistry of hydrothermal ore deposits (3rd ed., pp. 303–365). New York: Wiley. Reed, M. H. (1998). Calculation of simultaneous chemical equilibria in aqueous-mineral-gas systems and its application to modeling hydrothermal processes. In J. P. Richards (Ed.), Techniques in Hydrothermal Ore Deposits Geology (Vol. 10, pp. 109–124). Littleton, CO: Society of Economic Geologists, Inc. Robbins, S. J., Hynek, B. M., Lillis, R. J., & Bottke, W. F. (2013). Large impact crater histories of Mars: The effect of different model crater age techniques. Icarus, 225(1), 173–184. https://doi.org/10.1016/j.icarus.2013.03.019 Rubey, W. W. (1951). Geologic history of seawater: An attempt to state the problem. Geological Society of America Bulletin, 62(9), 1111–1148. https://doi.org/10.1130/0016- 7606(1951)62[1111:GHOSW]2.0.CO;2 Rubin, A. E., Zolensky, M. E., & Bodnar, R. J. (2002). The halite-bearing Zag and Monahans (1998) meteorite breccias: Shock metamorphism, thermal metamorphism and aqueous alteration on the H-chondrite parent body. Meteoritics & Planetary Science, 37(1), 125–141. https://doi.org/10.1111/j.1945-5100.2002.tb00799.x Ruesch, O., Poulet, F., Vincendon, M., Bibring, J.-P., Carter, J., Erkeling, G., … Reiss, D. (2012). Compositional investigation of the proposed chloride-bearing materials on Mars using near- infrared orbital data from OMEGA/MEx. Journal of Geophysical Research: Planets, 117(E11), E00J13. https://doi.org/10.1029/2012JE004108 Ruff, S. W., & Christensen, P. R. (2002). Bright and dark regions on Mars: Particle size and mineralogical characteristics based on Thermal Emission Spectrometer data. Journal of Geophysical Research: Planets, 107(E12), 2–1. https://doi.org/10.1029/2001JE001580 38 Salvany, J. M., Munoz, A., & Perez, A. (1994). Nonmarine evaporitic sedimentation and associated diagenetic processes of the southwestern margin of the Ebro Basin (lower Miocene), Spain. Journal of Sedimentary Research, 64(2a), 190. https://doi.org/10.1306/D4267D52-2B26- 11D7-8648000102C1865D Schilling, J.-G., Bergeron, M. B., Evans, R., & Smith, J. V. (1980). Halogens in the Mantle Beneath the North Atlantic [and Discussion]. Philosophical Transactions of the Royal Society of London. Series A, Mathematical and Physical Sciences, 297(1431), 147. https://doi.org/10.1098/rsta.1980.0208 Schmidt, M. E., Ruff, S. W., McCoy, T. J., Farrand, W. H., Johnson, J. R., Gellert, R., … Schroeder, C. (2008). Hydrothermal origin of halogens at Home Plate, Gusev Crater. Journal of Geophysical Research: Planets, 113(E6), E06S12. https://doi.org/10.1029/2007JE003027 Smith, D. E., Zuber, M. T., Frey, H. V., Garvin, J. B., Head, J. W., Muhleman, D. O., … Sun, X. (2001). Mars Orbiter Laser Altimeter: Experiment summary after the first year of global mapping of Mars. Journal of Geophysical Research: Planets, 106(E10), 23689–23722. https://doi.org/10.1029/2000JE001364 Smith, M. L., Claire, M. W., Catling, D. C., & Zahnle, K. J. (2014). The formation of sulfate, nitrate and perchlorate salts in the martian atmosphere. Icarus, 231, 51–64. https://doi.org/10.1016/j.icarus.2013.11.031 Spencer, R. J., & Hardie, L. A. (1990). Control of seawater composition by mixing of river waters and mid-ocean ridge hydrothermal brines. In R. J. Spencer & I.-M. Chou (Eds.), Fluid-Mineral Interactions: A Tribute to H.P. Eugster. (pp. 409–419). San Antonio, TX: The Geochemical Society. Retrieved from https://www.geochemsoc.org/files/2914/1261/1780/SP-2_409- 420_Spencer.pdf Stolper, E., & McSween, H. Y. (1979). Petrology and origin of the shergottite meteorites. Geochimica et Cosmochimica Acta, 43(9), 1475–1498. https://doi.org/10.1016/0016-7037(79)90142-X Tanaka, K. L., Robbins, S. J., Fortezzo, C. M., Skinner Jr., J. A., & Hare, T. M. (2014). The digital global geologic map of Mars: Chronostratigraphic ages, topographic and crater morphologic 39 characteristics, and updated resurfacing history. Planetary and Space Science, 95, 11–24. https://doi.org/10.1016/j.pss.2013.03.006 Toner, J. D., & Sletten, R. S. (2013). The formation of Ca-Cl-rich groundwaters in the Dry Valleys of Antarctica: Field measurements and modeling of reactive transport. Geochimica et Cosmochimica Acta, 110, 84–105. https://doi.org/10.1016/j.gca.2013.02.013 Toon, O. B., Segura, T., & Zahnle, K. (2010). The Formation of Martian River Valleys by Impacts. Annual Review of Earth and Planetary Sciences, 38(1), 303–322. https://doi.org/10.1146/annurev-earth-040809-152354 Tosca, N. J., & McLennan, S. M. (2006). Chemical divides and evaporite assemblages on Mars. Earth and Planetary Science Letters, 241(1–2), 21–31. https://doi.org/10.1016/j.epsl.2005.10.021 Turcotte, D. L., & Schubert, G. (2002). Geodynamics. (D. L. Turcotte, Ed.) (2nd ed). Cambridge ; New York: Cambridge University Press. Villanueva, G. L., Mumma, M. J., Novak, R. E., Käufl, H. U., Hartogh, P., Encrenaz, T., … Smith, M. D. (2015). Strong water isotopic anomalies in the martian atmosphere: Probing current and ancient reservoirs. Science, 348(6231), 218. https://doi.org/10.1126/science.aaa3630 Wänke, H., Dreibus, G., & Wright, I. P. (1994). Chemistry and Accretion History of Mars [and Discussion]. Philosophical Transactions of the Royal Society of London. Series A: Physical and Engineering Sciences, 349(1690), 285–293. https://doi.org/10.1098/rsta.1994.0132 Warren, J. K. (2010). Evaporites through time: Tectonic, climatic and eustatic controls in marine and nonmarine deposits. Earth-Science Reviews, 98(3–4), 217–268. https://doi.org/10.1016/j.earscirev.2009.11.004 Webster, J. D. (2004). The exsolution of magmatic hydrosaline chloride liquids. Chemical Geology, 210(1–4), 33–48. https://doi.org/10.1016/j.chemgeo.2004.06.003 Webster, J. D., Kinzler, R. J., & Mathez, E. A. (1999). Chloride and water solubility in basalt and andesite melts and implications for magmatic degassing. Geochimica et Cosmochimica Acta, 63(5), 729–738. https://doi.org/10.1016/S0016-7037(99)00043-5 40 Williams, R. M. E., & Weitz, C. M. (2014). Reconstructing the aqueous history within the southwestern Melas basin, Mars: Clues from stratigraphic and morphometric analyses of fans. Icarus, 242, 19–37. https://doi.org/10.1016/j.icarus.2014.06.030 Wittmann, A., Korotev, R. L., Jolliff, B. L., Irving, A. J., Moser, D. E., Barker, I., & Rumble, D. (2015). Petrography and composition of Martian regolith breccia meteorite Northwest Africa 7475. Meteoritics & Planetary Science, 50(2), 326–352. https://doi.org/10.1111/maps.12425 Wordsworth, R. D. (2016). The Climate of Early Mars. Annual Review of Earth and Planetary Sciences, 44(1), 381–408. https://doi.org/10.1146/annurev-earth-060115-012355 Ye, C., & Glotch, T. D. (2016). VNIR and MIR Spectral Features and Detection Limits of Minor Phases in Chloride-bearing Mineral Mixtures. In AGU Fall Meeting Abstracts (p. P21C– 2123). San Francisco, CA: American Geophysical Union. Retrieved from http://adsabs.harvard.edu/abs/2016AGUFM.P21C2123Y Zolotov, M. Y., & Mironenko, M. V. (2016). Chemical models for martian weathering profiles: Insights into formation of layered phyllosilicate and sulfate deposits. Icarus, 275, 203–220. https://doi.org/10.1016/j.icarus.2016.04.011 41 Tables Table 1. Composition of the reactant rock used in the geochemical models, based on Mazatzal basalt analyzed by Spirit at Gusev Crater, using APXS and Mössbauer (McSween et al., 2004, 2006). TiO2 and Cr2O3 exist in low concentration in Mazatzal (combined ~ 1 wt. %) and are not included because of lacking Ti and Cr minerals in the SOLTHERM thermodynamic database, and their low solubility. Component SiO2 Al2O3 Fe2O3 FeO MnO CaO Na2O K2O P2O5 FeS Cl Total Rock composition (wt. %) 46.22 10.88 2.14 17.08 0.44 8.35 2.66 0.11 0.64 0.84 0.15 100 Mineral (CIPW) Plagioclase (An43.7) Orthoclase Diopside Hypersthene Olivine (Fo52.7) Magnetite Apatite Pyrite Total Wt. % 39.98 0.65 16.49 6.03 31.63 3.1 1.48 0.64 100 42 Published in JGR: Planets Table 2. Calculated parameters of the chloride-bearing deposits studied here. The density of NaCl used here is 2165 kg/m3. Hynek et al. (2015) determined a minimum lake volume of 3.59 × 1010 m3 for the CBD lake site close to Miyamoto crater which we take as a maximum value, because we determine minimum lake volumes from the height of the CBDs at present (see Section 3.2 and Figure 6). West of Miyamoto Crater This study Hynek et al. (2015) Terra Sirenum West of Knobel Crater SE of Bunnik Crater Basin area (m2) 8.35E+8 1.23E+9 6.45E+9 3.47E+9 9.83E+9 Maximum lake depth (m) 195.4 NA 223 155 111 Mean lake depth at maximum capacity ± 1σ (m) 75.4 ± 22.1 NA 122.3 ± 32.6 98.7 ± 25.2 61.0 ± 17.1 Minimum lake area (m2) 1.35E+8 NA 2.76E+9 3.05E+8 3.98E+8 Maximum lake area (m2) 2.91E+8 NA 4.67E+9 8.53E+8 1.19E+9 Minimum lake volume (m3) Maximum lake volume (m3) 5.35E+9 NA 1.26E+11 9.56E+9 2.58E+10 2.19E+10 3.59E+10 5.72E+11 8.42E+10 7.25E+10 Mean deposit thickness (m) 1.50 4.00 3.24 7.95 8.16 Chloride-bearing deposit area (m2) Chloride-bearing deposit volume (m3) Basin area/deposit area ratio Volume of NaCl if CBD = 10 vol. % NaCl (m3) Mass of NaCl if CBD = 10 vol. % NaCl (kg) 1.43E+7 2.98E+7 2.93E+8 5.32E+5 2.85E+7 2.15E+7 1.19E+8 9.48E+8 4.23E+6 2.33E+8 58 41 22 6522 39 2.15E+6 1.19E+7 9.48E+7 4.23E+5 2.33E+7 4.65E+9 2.58E+10 2.05E+11 9.16E+8 5.04E+10 43 Published in JGR: Planets 2.82E+9 1.57E+10 1.25E+11 5.56E+8 3.06E+10 3.38 12.8 19.3 0.16 3.11 5.37E+6 2.98E+7 2.37E+8 1.06E+6 5.82E+7 1.16E+10 6.46E+10 5.13E+11 2.29E+9 1.26E+11 7.06E+9 3.92E+10 3.11E+11 1.39E+9 7.64E+10 8.45 32.0 48.2 0.40 7.77 Mass of Cl if CBD = 10 vol. % NaCl (kg) Mass of Cl/basin area if CBD = 10 vol. % NaCl (kg/m2) Volume of NaCl if CBD = 25 vol. % NaCl (m3) Mass of NaCl if CBD = 25 vol. % NaCl (kg) Mass of Cl if CBD = 25 vol. % NaCl (kg) Mass of Cl/basin area if CBD = 25 vol. % NaCl (kg/m2) 44 Published in JGR: Planets Table 3. Minimum and maximum volumes of water and masses of rock leached in the watersheds of the chloride-bearing deposits, as a result of basalt alteration with groundwater. At w/r = 1 the fluid contained 5.09 × 10-2 (moles Cl) · (kg H2O-1), at w/r = 2, the Cl concentration was 2.31 × 10-2 (moles Cl) · (kg H2O-1) (Fig. 7a). The density of poorly consolidated basalt (composition reported in Table 1) was assumed to be 1650 kg m-3 (similar to Mars sand analog with up to 20 % moisture; (Herkenhoff et al., 2008)). We assumed a thermal diffusivity typical for silicates (7 × 10-7 m2 s-1) to calculate the time required for a thermal wave to penetrate from the surface to the required depth of weathering. West of Miyamoto Crater This study Hynek et al. (2015) Terra Sirenum West of Knobel Crater SE of Bunnik Crater Minimum (10 vol. % NaCl in deposits) Volume of H2O at deposit site at w/r = 1 (m3) Water column over catchment at w/r = 1 (m) TWV at w/r = 1/Minimum lake volume ratio TWV at w/r = 1/Maximum lake volume ratio Volume of H2O at deposit site at w/r = 10 (m3) Water column over catchment at w/r = 2 (m) TWV at w/r = 2/Minimum lake volume ratio TWV at w/r = 2/Maximum lake volume ratio Total rock mass weathered (kg) Rock mass weathered per m2 in the basin (kg/m2) 1.57E+9 8.69E+9 6.91E+10 3.08E+8 1.69E+10 1.88 7.10 10.7 0.09 1.72 0.29 NA 0.55 0.03 0.66 0.07 0.24 0.12 0.004 0.23 1.85E+10 1.03E+11 8.16E+11 3.64E+9 2.00E+11 4.13 15.62 23.56 0.20 3.79 0.64 NA 1.21 0.07 1.45 0.16 0.53 0.27 0.01 0.51 1.88E+12 1.04E+13 8.30E+13 3.70E+11 2.04E+13 2.25E+3 8.53E+3 1.29E+4 1.07E+2 2.07E+3 45 Depth of rock weathered throughout the watershed (m) Time required for thermal wave to penetrate to required depth of weathering (Mars years) Maximum (25 vol. % in deposits) Volume of H2O at deposit site at w/r = 1 (m3) Water column over catchment at w/r = 1 (m) TWV at w/r = 1/Minimum lake volume ratio TWV at w/r = 1/Maximum lake volume ratio Total rock mass weathered (kg) Water column over catchment at w/r = 2 (m) TWV at w/r = 2/Minimum lake volume ratio TWV at w/r = 2/Maximum lake volume ratio Rock mass weathered per m2 in the basin (kg/m2) Depth of rock weathered throughout the watershed (m) Time required for thermal wave to penetrate to required depth of weathering (Mars years) Published in JGR: Planets 1.37 5.17 7.80 0.06 1.26 0.01 0.12 0.27 < 1 sol 0.01 3.91E+9 2.17E+10 1.73E+11 7.70E+8 4.24E+10 4.69 17.7 26.8 0.22 4.31 0.73 NA 1.37 0.08 1.64 0.18 0.61 0.30 0.01 0.58 4.71E+12 2.61E+13 2.08E+14 9.26E+11 5.09E+13 10.32 39.05 58.91 0.49 9.48 1.61 NA 3.03 0.18 3.62 0.39 1.33 0.67 0.02 1.29 5.64E+3 2.13E+4 3.22E+4 2.67E+2 5.18E+3 3.42 12.92 19.49 0.16 3.14 0.05 0.75 1.70 < 1 sol 0.04 46
1710.09776
1
1710
2017-10-26T15:57:33
Constraints on Super-Earths Interiors from Stellar Abundances
[ "astro-ph.EP" ]
Modeling the interior of exoplanets is essential to go further than the conclusions provided by mean density measurements. In addition to the still limited precision on the planets' fundamental parameters, models are limited by the existence of degeneracies on their compositions. Here we present a model of internal structure dedicated to the study of solid planets up to ~10 Earth masses, i.e. Super-Earths. When the measurement is available, the assumption that the bulk Fe/Si ratio of a planet is similar to that of its host star allows us to significantly reduce the existing degeneracy and more precisely constrain the planet's composition. Based on our model, we provide an update of the mass-radius relationships used to provide a first estimate of a planet's composition from density measurements. Our model is also applied to the cases of two well-known exoplanets, CoRoT-7b and Kepler-10b, using their recently updated parameters. The core mass fractions of CoRoT-7b and Kepler-10b are found to lie within the 10-37% and 10-33% ranges, respectively, allowing both planets to be compatible with an Earth-like composition. We also extend the recent study of Proxima Centauri b, and show that its radius may reach 1.94 Earth radii in the case of a 5 Earth masses planet, as there is a 96.7% probability that the real mass of Proxima Centauri b is below this value.
astro-ph.EP
astro-ph
Accepted for publication in The Astrophysical Journal Preprint typeset using LATEX style emulateapj v. 12/16/11 7 1 0 2 t c O 6 2 . ] P E h p - o r t s a [ 1 v 6 7 7 9 0 . 0 1 7 1 : v i X r a CONSTRAINTS ON SUPER-EARTHS INTERIORS FROM STELLAR ABUNDANCES B. Brugger1, O. Mousis1, M. Deleuil1, and F. Deschamps2 Accepted for publication in The Astrophysical Journal ABSTRACT Modeling the interior of exoplanets is essential to go further than the conclusions provided by mean density measurements. In addition to the still limited precision on the planets' fundamental param- eters, models are limited by the existence of degeneracies on their compositions. Here we present a model of internal structure dedicated to the study of solid planets up to ∼10 Earth masses, i.e. Super-Earths. When the measurement is available, the assumption that the bulk Fe/Si ratio of a planet is similar to that of its host star allows us to significantly reduce the existing degeneracy and more precisely constrain the planet's composition. Based on our model, we provide an update of the mass-radius relationships used to provide a first estimate of a planet's composition from den- sity measurements. Our model is also applied to the cases of two well-known exoplanets, CoRoT-7b and Kepler-10b, using their recently updated parameters. The core mass fractions of CoRoT-7b and Kepler-10b are found to lie within the 10–37% and 10–33% ranges, respectively, allowing both planets to be compatible with an Earth-like composition. We also extend the recent study of Proxima Cen- tauri b, and show that its radius may reach 1.94 R⊕ in the case of a 5 M⊕ planet, as there is a 96.7% probability that the real mass of Proxima Centauri b is below this value. Keywords: Earth - planets and satellites: composition - planets and satellites: individual (CoRoT- 7b, Kepler-10b, Proxima Centauri b) - planets and satellites: interiors 1. INTRODUCTION The huge diversity of discovered worlds, in terms of physical and orbital parameters, led to enlarge the planet population with new families. Among them, Super- Earths and sub-Neptune bodies fill the gap between the terrestrial planets and the giant gaseous planets that compose our solar system. Measurements of the mass and radius of an exoplanet, mostly obtained from radial velocity and transit methods, respectively, allow to de- rive the body's mean density. This quantity gives a rough estimate of the planet's bulk composition, whether the mean density is closer to that of the Earth (5.51 g/cm3) or to that a gaseous planet like Jupiter (1.33 g/cm3). To better constrain the composition of exoplanets, that is the distribution of elements inside these bodies, interior models have been developed, based on our knowledge of the properties of the Earth and other solar system bodies (Valencia et al. 2006; Sotin et al. 2007; Seager et al. 2007; Zeng & Seager 2008; Rogers & Seager 2010; Dorn et al. 2015). Models of planetary interiors inherently present a degeneracy issue as two planetary bodies with different compositions may have the same mass and radius. For instance, the same set of mass and radius allows interior models to generate both a planet displaying properties similar to those of the Earth (silicate mantle surrounding a relatively small metal core) and a planet possessing a larger core with a smaller mantle surrounded by a thick water layer. In this paper, we present an interior model derived from those developed for the Earth, and able to han- dle compositions as various as those of small planets and 1 Aix Marseille Univ, CNRS, LAM, toire ([email protected]) d'Astrophysique de Marseille, Marseille, Labora- France 2 Institute of Earth Sciences, Academia Sinica, 128 Academia Road Sec. 2, Nangang, Taipei 11529, Taiwan large satellites of the solar system (Mercury-like to ocean worlds). This model is limited to the case of dense solid planets with possible addition of water, and does not consider planets that harbour thick gaseous atmospheres made of H/He. Using a more appropriate equation of state compared to previous studies, we provide up-to- date mass-radius relationships for planets with masses under 20 M⊕. Coupled to an adapted numerical scheme, our model explores the parameter space for the possible compositions of solid planets to minimize the set of com- positions consistent with the measured physical proper- ties. We aim at breaking the aforementioned degener- acy by incorporating the Fe/Si bulk ratio of the investi- gated planet into our code. This ratio helps constraining the size of the metal core inside the planet, a parameter mostly concerned by this degeneracy. The numerical model is described in Section 2. Sec- tion 3 presents our results concerning the investigation of the interiors of two well-known low-mass exoplanets, namely CoRoT-7b and Kepler-10b, based on the lat- est estimates of their physical parameters, and assuming they do not contain gaseous envelopes. We also investi- gate the interior of Proxima Centauri b by considering a larger mass range than the minimum mass considered in a previous study (Brugger et al. 2016). Section 4 is dedicated to discussion and conclusions. 2. MODEL 2.1. Planet internal structure Our model is based on the approach described by Sotin It assumes a fully differentiated planet et al. (2007). with several shells (or layers). Reflecting the terrestrial planets in our solar system, the three main layers consist of a metallic core, a silicate mantle, and a hydrosphere. In our model, the mantle and the hydrosphere can be divided into two sublayers each, leading to planets that 2 Table 1 Compositional parameters and surface conditions for the Earth (Stacey 2005; Sotin et al. 2007) Parameter Value Description xcore xwater falloy (cid:16) Mg (cid:17) (cid:17) (cid:16) Fe Si P Si P Mg# Tsurf (K) Psurf (bar) 0.325 Core mass fraction 0.0005 Water mass fraction 13% Fraction of alloy in the core 1.131 Corrected Mg/Si ratio 0.986 Corrected Fe/Si ratio 0.9 288 1 Mg number Surface temperature Surface pressure Figure 1. Schematic view of the different concentric layers that compose our interior model: metallic core, lower and upper silicate mantles, high-pressure water ice, and liquid water. Depending on the mass of the water layers, the upper mantle or the high pressure ice layer may be absent. can be made of up to a total of five concentric layers (see Figure 1), which are from the center: 1. The core. In the Earth, the core is essentially composed of iron, along with smaller fractions of other metals (as nickel or sulfur). Here we assume that it consists in a single layer formed of a mixture of pure iron (Fe) and iron alloy (FeS). 2. The lower mantle. Here, the elements Fe, Mg, Si and O form the silicate rocks. The lower mantle corresponds to a region of high pressure (Dziewon- ski & Anderson 1981) in the phase diagram of sil- icates, which can be in the forms of bridgman- ite (Mg,Fe)SiO3 and ferro-periclase (Mg,Fe)O (also known as perovskite and magnesiowustite, respec- tively). 3. The upper mantle. This layer is made of the same elements as in the lower mantle, but in the form of olivine (Mg,Fe)2SiO4 and ortho-pyroxene enstatite (Mg,Fe)2Si2O6 because of a lower pres- sure (Dziewonski & Anderson 1981). 4. The high-pressure water ice. As for the mantle, the hydrosphere divides into two layers: because some planets may include a significant amount of water, a layer of water ice VII can form at pressures reaching several GPa (Frank et al. 2004). Note that if the ice VII layer is thick enough the pres- sure at its bottom may be too large for olivine and enstatite to exist, in which case the upper mantle would be absent. 5. The liquid water. On top of ice VII, water is in liquid form, provided that the surface conditions of the planet are close to the Earth's values. These five layers, represented in Figure 1, allow to model the interiors of planets with various compositions, from terrestrial (i.e. fully rocky) planets like the Earth or Mercury, to ocean planets that possess a massive amount of water (like the icy moons of the jovian and saturnian systems). The differences between two planetary compo- sitions are determined by the masses and sizes of these layers. In practice, the model's necessary inputs cor- respond to the masses of the three main layers (core, mantle, and hydrosphere), since their distribution into the distinct sublayers can be computed from the phase change laws of the different materials. These inputs can be expressed in terms of xcore, xmantle and xwater, corre- sponding to fractions of the planet's total mass for the core, mantle and hydrosphere, respectively. From mass conservation, we get xmantle = 1 − xcore − xwater, allow- ing one to derive the planet's interior from the a pri- ori knowledge of its total mass and the values of the core mass fraction (CMF) xcore and water mass fraction (WMF) xwater. (cid:17) (cid:16) Mg Three additional parameters set the distribution of all chemical species in the different layers: the fraction of alloy in the core falloy, the overall Mg/Si ratio of the planet , and the amount of iron present in the silicate mantles Mg# ≡ (cid:16) Mg , describing the (cid:17) Si P Mg+Fe Mantle level of differentiation of the planet (Sotin et al. 2007). In the absence of compositional data for the host stars of investigated planets, we use those derived from the Earth (see Table 1). Once the composition and the mass of the planet are set, an iterative process solves the canonical equations for gravitational acceleration g, pressure P , temperature T , and density ρ, until convergence is reached, and provides a planet radius. This process, as well as the convergence conditions, are detailed in the Appendix. 2.2. Equations of state An equation of state (EOS) gives the dependence of density to pressure and temperature. It is specific to each material through its thermodynamic and elastic parame- ters. Most EOS are obtained by fitting the measurements of ρ versus P and T in laboratory experiments. This fit is then extrapolated outside the measurement range to reach pressure values existing inside the Earth and more massive planets. Sotin et al. (2007) used the well-known Birch- Murnaghan EOS, in its third-order development (see Ap- pendix), where the thermal dependency of the pressure P is directly incorporated in the coefficients. The third- CoreIronIron alloyUpper mantleLower mantleLiquid waterHigh pressure iceBridgmaniteFerro-periclaseEnstatiteOlivine order Birch-Murnaghan EOS (hereafter BM3 EOS) is often used because it is particularly fast for computa- tions. However this EOS is limited to low pressure val- ues (P < 150–300 GPa; Seager et al. 2007) and to tem- peratures of the same orders of magnitude as those en- countered inside the Earth's mantle (Sotin et al. 2007). Above these values, the results from this EOS may devi- ate from theory (Valencia et al. 2009). Therefore, Sotin et al. (2007) used the BM3 EOS in the upper mantle and the liquid water layer, where the pressure remains low. In the deeper layers, the authors used the Mie-Gruneisen- Debye (MGD) formulation as a replacement for BM3, which better describes the behavior of the used materials. The MGD EOS dissociates static pressure and thermal pressure, by adding a thermal Debye correction to the isothermal Mie-Gruneisen formulation (see Appendix). In our model, we replace the MGD EOS by another EOS already used by Valencia et al. (2007a), namely the Vinet EOS (see Appendix; Vinet et al. 1989). As for the MGD EOS, the Vinet EOS separates static and thermal pressures with the use of the thermal Debye correction. By doing so, the Vinet EOS presents the same advantages as the MGD formulation in comparison to the BM3 EOS. However, although these three EOS have the same valid- ity range in pressure, the Vinet EOS has been shown to better reproduce experimental data than other EOS, and to better extrapolate at pressures higher than ∼100 GPa (Hama & Suito 1996; Cohen et al. 2000). Therefore, it is well-adapted to the modeling of Super-Earths, where such high pressures can easily be reached. We however keep the BM3 formula within the hydrosphere, since the pressure and density remain relatively low in the upper two layers. 2.3. Water model We consider two phases of water: ice VII and liquid water. For planets with large amounts of water, ice VII appears at pressures higher than ∼1 GPa. The condi- tions for the existence of liquid water are based on a simplified version of the phase diagram of water (see Fig- ure 2). The liquid–ice VII transition law is taken from Frank et al. (2004) who fitted a larger range of pressures (from 3 to 60 GPa) than previous works. The remaining phase change laws are taken from Wagner & Pruss (2002) and the website of the International Association for the Properties of Water and Steam (IAPWS)3. 2.4. Exploration of the ternary diagram (cid:16) Mg (cid:17) P Si In our approach, the planet's mass is first set to its measured value. The compositional parameters falloy, , and Mg# do not have a significant influence on the planet's radius compared to its mass, the CMF and the WMF (Sotin et al. 2007; Valencia et al. 2006). They are then fixed by default to the Earth's values. The CMF and WMF are then the remaining free param- eters as they cannot be measured. They are both varied within the [0–1] range of values and linked by the rela- tion xcore + xmantle + xwater = 1. The parameter space formed by the variation of these three variables is repre- sented by a ternary diagram (see Figure 4) displaying the 3 http://www.iapws.org/relguide/MeltSub.html 3 Figure 2. Simplified version of the water phase diagram used in the model to compute the boundaries between the different water layers. Black dots are triple points of water, and red curves delimit the phase transitions fitted from experimental data. Light grey lines represent the phase boundaries not considered in our work. mass fractions of the three main layers forming an exo- planet: core, mantle, and hydrosphere (or water) (Va- lencia et al. 2007b). Each point on the ternary diagram corresponds to a unique planet composition given by the pair (CMF,WMF). For instance, an Earth-like com- position corresponds to (CMF,WMF) = (32.5%,0.05%) (Stacey 1992, 2005), and a Mercury-like composition to (CMF,WMF) = (68%,0%), even if these values are still under debate (Schubert et al. 1988; Harder & Schubert 2001; Spohn et al. 2001; Stacey 2005). Mercury is com- pletely dry, and the Earth's WMF is close to zero, as for all terrestrial planets in our solar system. Several moons of Jupiter and Saturn are not dry, and present a significant water amount, like Titan, with (CMF,WMF) = (0%,50%) (Tobie et al. 2006). A numerical scheme allows the model to explore the entire domain of planetary compositions formed by the ternary diagram, which produces a colormap of com- puted planet radii (see Figure 3). Isoradius lines drawn on this colormap illustrate the degeneracy existing in models of internal structure, as two planets may have the same mass and radius but not the same composition (i.e. they are located at different points in the diagram). 2.5. Physical limitations on planetary compositions As suggested by Valencia et al. (2007b), an upper limit of 65% can be set on the CMF, assuming that the bulk Fe/Si ratio of the considered planet is protosolar. These authors also limit the possible values of the WMF to 77% at maximum, according to measurements on cometary compositions. In our case, we lower this value to 50%, based on our knowledge of the interiors of large icy satel- lites such as Titan (Tobie et al. 2006). The areas on the ternary diagram corresponding to these restrictions are thus shaded (see Figure 3). Interestingly, Mercury lies inside this exclusion region but its current state could result from a post-formation alteration such as mantle evaporation due to strong melting (Cameron 1985) or a giant impact during the early phases of its evolution 10010210410610810101012 0 200 400 600 800 1000Ice IhOtherVaporLiquidSupercriticalIce VIIPressure (Pa)Temperature (K) 4 Figure 3. Colormap of the planet's radius on the ternary diagram, as a function of its composition for a given mass (here MP = 1M⊕). The computed planet radii range from 0.79 R⊕ to 1.40 R⊕. The black lines represent isoradius curves with a separation of 0.1 R⊕. Note that the point denoting the Earth's composition does not lay on the 1 R⊕ isoradius curve (see Appendix). (Benz et al. 1988). Moreover, one must not forget that the planets considered in this work are made from a lim- ited number of materials. Planets harbouring a thick gaseous atmosphere can easily be larger than the maxi- mum value allowed with water, however we consider in this study only dense solid planets (with possible addi- tion of liquid water) without thick gaseous atmospheres. These restrictions on the ternary diagram are based on our current knowledge of the solar system bodies, and they may not be suitable for application to all exoplan- etary systems, as they could have had different planet formation conditions. Morever, these restrictions do not break the degeneracy existing on a planet's composition. This would however be possible if the Fe/Si ratio of the planet is known, since it is strongly related to the CMF value. With our assumptions, the Fe/Si ratio of a planet is independent from its mass, and is essentially governed by the CMF and WMF (see Appendix). Thus, as for the radius, we can draw isolines of constant Fe/Si ra- tios in the ternary diagram (see Figure 4). The intersec- tion of the isoline of a planet's Fe/Si with its isoradius curve would then be the only composition allowed for this planet. The Fe/Si and Mg/Si bulk ratios of an exo- planet cannot be directly measured but their host star's values can provide a good approximation (Thiabaud et al. 2015). Here, we assume that the Fe/Si ratios of the exoplanet and its host star are similar to break the de- generacy on the planet's composition. 3. RESULTS 3.1. Mass-radius relationships Our model allows the computation of RP as a function of MP for different compositions. Comparing these val- ues to the measured physical parameters then provides a first estimate of the exoplanet's composition. Here we chose six typical compositions, namely 100% water, 50% mantle–50% water, 100% mantle, Earth-like, Mercury- like, and 100% core (see Figure 5). As explained in Sec- tion 2.5, planets with extreme compositions like 100% Figure 4. Isolines of constant planetary Fe/Si ratio in the ternary diagram. The Earth's value is shown in red. Si (cid:17) (cid:16) Mg water or 100% core are unlikely to form, however the mass-radius curves corresponding to these compositions provide good theoretical markers for the validation of fundamental parameters of detected exoplanets. If an exoplanet is located under the 100% core curve, it chal- lenges its mass and radius estimates, as no planetary body can form with a density higher than pure iron. On the other hand, a planet located beyond the 100% water curve cannot be composed of the materials used here, and most likely harbours a significant gaseous atmosphere. The parameters falloy, , and Mg# are taken equal to the Earth's value (see Table 1). We explore planetary masses up to 20 M⊕, which corresponds to the upper limit of validity of the Vinet EOS (pressures of the or- der of 1–10 TPa; Hama & Suito 1996). Placing some well-known exoplanets on Figure 5 provides indications of their possible compositions. Kepler-10c, a ∼14 M⊕ planet (Weiss et al. 2016), lies above the line correspond- ing to a 100% mantle composition, i.e. the least dense fully rocky composition, and probably harbours a signif- icant fraction of water. On the other hand, planets like Kepler-10b and CoRoT-7b, with the latest estimates of their fundamental parameters, are compatible with fully rocky compositions, and even with an Earth-like com- position. A more detailed study of these exoplanets is provided in Section 3.2. P P P P Previous studies found that, for an Earth-like compo- sition, the planet radius grows proportional to M 0.274 (Sotin et al. 2007) or M 0.267–0.272 (Valencia et al. 2006) for MP ∈ [1–10]M⊕. When considering ocean planets (with a 50% WMF) the power exponent re- mains similar, but the radius increases faster (RP = 1.262 M 0.275 ; Sotin et al. 2007) because of the lower density of the water phases. For a Mercury-like com- position, the power exponent becomes ∼0.3 (Valencia et al. 2006). Here, when fitting our mass-radius curves, we obtain RP = 1.00 M 0.270 for an Earth-like com- position and MP ∈ [1–10] M⊕. The power exponent is lowered for ocean planets (50% mantle–50% water) (RP = 1.27 M 0.261 ) and for Mercury-like planets (RP = 0.90 M 0.268 ). The value close to 0.3 found P P P WaterCoreMantleEarthMercuryRadius (R⊕)Mass (M⊕)1.00 0.7 0.8 0.9 1 1.1 1.2 1.3 1.4 1.50.91.0Rmin1.11.21.3RmaxWaterCoreMantleEarthMercury0.1260.51.01.52.02.55.010.0Earth 5 in the case of a ternary diagram. We further reduce this set of compositions by considering the Fe/Si ratio of the planet's host star. From this data, we provide the ranges of plausible values for the CMF and WMF allowed in the planet, considering the uncertainties on MP, RP, and on the stellar Fe/Si ratio, assuming that the planet does not harbour a thick gaseous atmosphere. 3.2.1. CoRoT-7b CoRoT-7b is the first detected Super-Earth with known mass and radius. The discovery of this planet was reported by L´eger et al. (2009) who found a ra- dius of 1.68 ± 0.09 R⊕ and an orbital period of 0.85359 ± 5 10−5 day. Its mass was obtained shortly after by Queloz et al. (2009), who derived a value of 4.8 ± 0.8 M⊕ using radial velocity measurements, and bestowed the Super-Earth status to CoRoT-7b from its derived mean density. Meanwhile, Valencia et al. (2010) performed a detailed study of CoRoT-7b's in- terior and composition, as well as of the mass loss re- sulting from the proximity of the planet to its host star (0.0172 ± 0.00029 AU; Queloz et al. 2009). They con- cluded that the planet could not retain a gaseous at- mosphere made of H/He because of stellar irradiation, implying the exclusion of such a layer. However, they found that the values of MP and RP could still be con- sistent with a solid planet, provided that its iron con- tent is significantly depleted compared to the Earth. An Earth-like composition was only reached with a 1σ in- crease in mass (MP = 5.6 M⊕) and a 1σ decrease in radius (RP = 1.59 R⊕) compared to the central values. The physical parameters of CoRoT-7b have been re- fined by Barros et al. (2014) and Haywood et al. (2014), updating the values of radius and mass to 1.585 ± 0.064 R⊕ and 4.73 ± 0.95 M⊕, respectively. Us- ing these new values and assuming that the planet is only made of solid materials (with the possible addition of liquid water), we have re-evaluated here the internal structure of CoRoT-7b. For this, we have explored the domain of planetary compositions corresponding to one of CoRoT-7b's possible formation scenarios, i.e. in situ formation (or formation close to the star). Two of the compositional parameters that are not represented in our diagrams (falloy and Mg#) are by default taken equal to the Earth's values. The surface temperature of the planet is taken equal to the estimated equilibrium temperature of CoRoT-7b (1756 ± 27 K; Barros et al. 2014), whereas the surface pressure is set to 1 bar to mimic the presence of a light atmosphere. The Fe, Mg and Si elemental abundances in CoRoT-7 have been derived from high resolution spectroscopy by Bruntt et al. (2010). These authors also determined the abundances of Al, Ca, and Ni, among many other refrac- tory elements, in CoRoT-7. Here, following the approach of Sotin et al. (2007), these three elements are added to Fe, Mg, and Si and used to correct the Fe/Si and Mg/Si bulk ratios employed in the model (see Table 1). In the case of the Earth, Fe, Mg, Si, O, and S only account for 95% of the planet's mass, whereas the addition of Al, Ca, and Ni enlarges this fraction to more than 99% (Morgan & Anders 1980; All`egre et al. 1995). From these consid- erations, we get corrected Fe/Si and Mg/Si ratios equal to 0.826 ± 0.419 and 1.036 ± 0.614 for the bulk compo- sition of CoRoT-7b, respectively. Contrary to the Fe/Si Figure 5. From top to bottom, mass-radius curves for different planet compositions: 100% water, 50% mantle–50% water, 100% mantle, Earth-like, Mercury-like, and 100% core. Solar system planets that have a mass in the explored range are shown. Three exoplanets, with the measurement errors on their physical param- eters, are also shown: CoRoT-7b with original (1) and updated (2) values (see Section 3.2), Kepler-10b, and Kepler-10c (Dumusque et al. 2014; Weiss et al. 2016). The data used for this figure are available online (see text). by Valencia et al. (2006) is only retrieved within the 0.1– 1 M⊕ range. Overall, we obtain lower values compared to Valencia et al. (2006) and Sotin et al. (2007). This discrepancy originates from the extrapolation differences between the Vinet EOS and the BM3 EOS, since the lat- ter results in an overestimation of the planet radius. Be- cause the power laws detailed here do not provide perfect fits of the mass-radius curves represented in Figure 5, we recommend the use of the machine-readable table avail- able online. 3.2. Compositions of Super-Earths In the following, when investigating the composition of an exoplanet without inclusion of water, we draw a colormap of the computed planet radii as a function of composition and of the planet's measured mass range. In the case where we consider the presence of water, we draw instead three ternary diagrams corresponding to the minimum (MP − δMP), central (MP), and maximum (MP + δMP) values of its mass range, as a complete in- vestigation of this range cannot be represented in two dimensions. To explore the impact of the uncertainties δRP on the planet radius RP, we draw the three isoradius curves RP−δRP, RP, and RP+δRP on each diagram. On one diagram, the domain included within the RP − δRP and RP + δRP curves corresponds to the set of composi- tions allowed for this planet, and for the considered mass 0.5 1 1.5 2 2.5 0.1 1 10Radius (R⊕)Mass (M⊕)100% Water50% Mantle-50% Water100% MantleEarthMercury100% CoreMarsVenusEarthUranusNeptuneKepler-10bKepler-10cCoRoT-7b(2)CoRoT-7b(1) 6 Figure 6. Computed planet radii for CoRoT-7b as a function of composition (X-axis, corresponding to the CMF as we assume WMF=0) and mass (Y-axis, from the 1σ-range inferred by Hay- wood et al. (2014)). Also shown are the isoradius curves denoting the planet radius measured by Barros et al. (2014) with the 1σ extreme values. The Fe/Si ratio assumed for CoRoT-7b, with its associated uncertainties, delimit a line and an area represented in red. Figure 7. Computed planet radii for Kepler-10b as a function of composition (X-axis, corresponding to the CMF as we assume WMF=0) and mass (Y-axis, from the 1σ-range inferred by Weiss et al. (2016)). Also shown are the isoradius curves denoting the planet radius measured by Dumusque et al. (2014) with the 1σ extreme values. The Fe/Si ratio assumed for Kepler-10b, with its associated uncertainties, delimit a line and an area represented in red. ratio whose full range is investigated through the model, only the central Mg/Si value is used as an input param- eter because this latter does not significantly impact the computed radius (Sotin et al. 2007). As for the terrestrial planets in the solar system, CoRoT-7b may have formed in situ, or at least inside the snow line of the protoplanetary disk that surrounded the host star. In this region, the temperature in the protoplanetary disk was too high for water to condense, leading to formation of fully rocky planets. Figure 6 shows the result of the exploration of CoRoT-7b's com- positional parameter space, if we assume completely dry compositions only (i.e., WMF=0). It illustrates the fact that an Earth-like composition is easier to achieve than with the previous planet's parameters. This particular composition can be obtained for all planetary masses be- tween 4.73 and 5.68 M⊕, considering the 1σ uncertainty on the planet's radius. The conclusion of Valencia et al. (2010), i.e. the fact that CoRoT-7b needs to be de- pleted in iron to explain the measured parameters, is no longer required with the updated parameters, since now the planet can present a CMF as large as that of the Earth. Continuing with the assumption of a dry compo- sition, we show that CoRoT-7b's CMF may vary between 0 and 50%, depending on the uncertainties on the fun- damental parameters. However, when our estimate of the Fe/Si ratio is considered in the planet, this range becomes limited to 13–37%. 3.2.2. Kepler-10b Kepler-10b is the first rocky planet that has been de- tected by the Kepler mission (Batalha et al. 2011). While it is in many aspects comparable to CoRoT-7b, its ra- dius (1.47+0.03−0.02 R⊕; Dumusque et al. (2014)) and mass (3.72 ± 0.42 M⊕; Weiss et al. (2016)) were measured with a better precision. Weiss et al. (2016) performed a study of Kepler-10b's interior, showing that a fully rocky composition is compatible with the measurements of mass and radius. They also derived a CMF of the planet within the 0.17 ± 0.12 range in the solid case. Here, we use the host star's chemical composition derived by Santos et al. (2015), and compute the corresponding Fe/Si and Mg/Si ratios. Unlike the case of CoRoT-7, the Al, Ca, and Ni abundances have not been measured in Kepler-10. We then find Fe/Si = 0.708 ± 0.375 and Mg/Si = 1.230 ± 0.595. Figure 7 shows the improved precision on Kepler-10b's radius, as the domain of pos- sible compositions delimited by the 1σ isoradius lines is more restrained for this planet than for CoRoT-7b. As for CoRoT-7b, we investigate the possibility that Kepler-10b formed inside the snow line (see Figure 7). In the case of a fully rocky planet with MP = 3.72 M⊕, the CMF of Kepler-10b varies between 12 and 26%, namely roughly within the upper half of the range of Weiss et al. (2016). With a 1σ decrease on the planet's mass, this range becomes 0–13%. On the other hand, with a 1σ in- crease of the mass, the CMF becomes 24–38%, a range of values becoming closer to those estimated for the Earth or Venus. In particular, a fully rocky Kepler-10b with MP = 4.14 M⊕ and RP = 1.47 R⊕ presents the same CMF as for the Earth (32%). When considering the lim- itations imposed by the Fe/Si ratio, the CMF associated to ±1σ in mass is now reduced to 10–13% and 24–33%, respectively. The overall upper limit on the CMF is thus much closer to the Earth's value. The range evaluated for the central mass is not affected by the consideration of the planet's Fe/Si ratio. We find the possible CMF range of Kepler-10b to be in the range 10–33%, which is in good agreement with the results from Weiss et al. (2016). However, their re- sults ruled out an Earth-like composition for Kepler-10b, whereas we show here that this composition is possible. Kepler-10b appears to be one of the best cases for the study of exoplanetary composition, thanks to both the 0 0.2 0.4 0.6 0.8 1Core mass fraction 3.8 4 4.2 4.4 4.6 4.8 5 5.2 5.4 5.6Planet mass (M⊕) 1.1 1.2 1.3 1.4 1.5 1.6 1.7 1.8Radius (R⊕)EarthMercury1.5211.5851.649 0 0.2 0.4 0.6 0.8 1Core mass fraction 3.3 3.4 3.5 3.6 3.7 3.8 3.9 4 4.1Planet mass (M⊕) 1 1.1 1.2 1.3 1.4 1.5 1.6Radius (R⊕)EarthMercury1.451.471.50 7 bodies (see Section 2.5). With MP = 2 M⊕, we obtain Rmin = 1.10 R⊕ for composition A, and Rmax = 1.53 R⊕ for composition B. These values do not differ significantly from the results found by Brugger et al. (2016). The planet's radius computed for MP = 5 M⊕ spans the range 1.41–1.94 R⊕ (see Figure 8). These results are in good agreement with the density-radius relationship obtained by Weiss et al. (2016) from statistical analysis of Kepler planets with known masses and radii, in the regime of planets smaller than 4 R⊕. These authors de- rive two regimes in which the mean density increases with the radius up to 1.5 R⊕ and then decreases for radii in the 1.5–4 R⊕ range. The maximum mean density is thus found to be 7.6 g/cm3 with a corresponding radius of 1.5 R⊕. In our simulations, the case of a 5 M⊕ Proxima b with composition A yields a mean density of 9.9 g/cm3, significantly higher than this statistical maximum value. In the case of composition B, we obtain MP = 5 M⊕ and RP = 1.94 R⊕, giving a mean density of 3.8 g/cm3, which agrees well with the relationship found by Weiss et al. (2016). The capability of Proxima b to retain a possi- ble liquid water ocean facing the strong stellar irradiation is uncertain (Ribas et al. 2016). However Airapetian et al. (2017) recently concluded that the planet cannot be habitable, since an Earth-like atmosphere would escape in ∼10 Myr because of the star's strong XUV flux. In the case of Proxima b being a Super-Earth with a mass up to 5 M⊕, the higher escape velocity at its surface would not be sufficient to retain the atmosphere, as this effect would be counterbalanced by a weaker magnetic field, compared to an Earth-sized planet (Airapetian et al. 2017). 4. DISCUSSION AND CONCLUSIONS We have developed an internal structure model cou- pled to a numerical scheme that explores the composi- tional parameter space of a planet. In our approach, we have assumed that the Fe/Si ratio measured in host stars is similar to that of their orbiting planets, allow- ing us to significantly decrease the degeneracy on their compositions. Applying this model to CoRoT-7b and Kepler-10b with the latest estimates of their mass and radius, we show that both planets present CMF values varying within the 10–37% and 10–33% ranges, respec- tively. The fundamental parameters of these two planets are consistent with fully rocky compositions, and espe- cially with an Earth-like composition. These results are compatible with a formation inside the snow line for both planets, are are in good agreement with those of Dorn et al. (2017) who studied the interiors of those planets using a bayesian method, and also found that they are com- patible with an Earth-like composition. They showed that the CMF of Kepler-10b's and CoRoT-7b spans a range approximately between 0 and that of the Earth. Note that we have modeled these planets assuming sur- face temperatures equal to their equilibrium tempera- tures with surface materials in solid phases. Such high surface temperatures should however lead to the forma- tion of melted silicates oceans (Schaefer & Fegley 2009), a phase whose density differs from those of the solid phases by a few percent (L´eger et al. 2011; Lebrun et al. 2013). Reporting this difference in density in our model pro- duces a difference on the computed radius by less than 1%. However, the impact of melted silicates on a planet's Figure 8. Ternary diagram of all compositions explored for a 5 M⊕ Proxima b, with the minimum and maximum computed planet radii corresponding to compositions A and B (red dots, see text), and the isoradius curves from 1.3 to 2.1 R⊕ (black curves). high precision on its mass and radius and the measure- ment of its host star's elemental abundances. Together, these two features reduce the set of compositions allowed for this planet. 3.2.3. Proxima Centauri b Proxima Centauri b, a low-mass planet orbiting the Sun's closest neighbor, was recently discovered by Anglada-Escud´e et al. (2016). The planet's radius re- mains unknown because no transit has been detected so far (Kipping et al. 2017), and the only known physical parameter is the planet's minimum mass MPsin i found to be 1.27+0.19−0.17 M⊕ (Anglada-Escud´e et al. 2016). Brug- ger et al. (2016) have performed a study of the possible interiors and compositions of Proxima b as a rocky body, with possible addition of water, assuming that sin i = 1. In their study, the computed radius of Proxima b spans the 0.94–1.40 R⊕ range, the minimum value being ob- tained in the case of a 1.10 M⊕ dry planet with a 65% CMF, and a maximum value reached when considering MP = 1.46 M⊕ for an ocean planet with 50% water in mass. Here, we extend the study of Brugger et al. (2016) by exploring the impact of a large range of values of sin i on the mass of Proxima b, still assuming that the planet does not possess a thick gaseous atmosphere. From our computations, there is a 77% probability that the planet's mass is within the range [1.27–2] M⊕, which we find to correspond to a maximum orbital inclination of 39.4◦. If the mass range is extended to [1.27–5] M⊕ (i.e. i = 14.7◦), the computed probability reaches 96.7%. Interestingly, Proxima b is located on a temperate orbit around its host star, with an equilibrium temperature es- timated at 234 K (Anglada-Escud´e et al. 2016). There- fore, liquid water can easily be stable on its surface with the presence of an Earth-like, light atmosphere, and su- percritical water is not needed to model the current state of the planet, in contrast to CoRoT-7b and Kepler-10b. Let "A" and "B" be the compositions that produce the minimum and maximum radii for a given value of MP, namely (CMF,WMF) = (0.65,0) and (0,0.5), respec- tively, according to the characteristics of solar system WaterCoreMantleEarthMercuryRadius (R⊕)Mass (M⊕)5.00 1.2 1.3 1.4 1.5 1.6 1.7 1.8 1.9 2 2.1 2.21.41.5Rmin1.61.71.81.92.0Rmax 8 radius could be more important, as other material prop- erties, namely the bulk modulus, may be strongly al- tered. In addition, the heat transfer in this layer would be modified compared to solid materials. This would re- quire, in particular, a finer description of the planet's thermal profile, which is not investigated here. Alternatively, these planets may have formed further away from their host star. In that case, water would have condensed along with silicate and metal grains, forming planets that contain a significant amount of water in liq- uid or solid phases. To reach their current orbital posi- tions, water-rich CoRoT-7b and Kepler-10b would have migrated inward from their formation region. Moreover, given their current equilibrium temperature, water on their surface would have turned into vapor and super- critical phases (for Tsurf > 647 K; Wagner & Pruss 2002) instead of liquid. Finally, we have considered the possibility that the re- cently discovered Proxima Centauri b presents a mass reaching 5 M⊕ (which covers 96.7% of the orbital incli- nations of the system), and shown that its radius may be as large as 1.94 R⊕ when considering an important amount of volatiles (50% water in mass). This value is still below the mean mass-radius relationship found by Weiss et al. (2016) for low-mass terrestrial exoplanets. A third formation scenario has been proposed by Lee et al. (2014). This model assumes a formation below the snow line with accretion of nebular gas, even for low- mass planets, which could thus harbour an H2/He atmo- sphere of a few percent in mass. In the case of CoRoT-7b, Valencia et al. (2010) showed that stellar irradiation is too strong for the planet to retain such an atmosphere. This also applies for Kepler-10b, which undergoes simi- lar conditions, with an orbital distance and a mass even smaller than those of CoRoT-7b (Dumusque et al. 2014; Weiss et al. 2016). This third formation scenario may apply to Proxima b, which could still harbour a light primary atmosphere, especially given its unknown actual mass. Modeling the interior of a gaseous Proxima b is thus needed to complete the investigation of this planet. Note however that giant gaseous planets appear to be rel- atively rare at short orbits around M dwarf stars (Bonfils et al. 2013; Tuomi et al. 2014; Dressing & Charbonneau 2015). 4.1. Diversity of planetary materials Since our model is predominantly based on the Earth's interior, the existence of the materials considered here is only verified within ranges of pressure (105–1011 Pa) and temperature (300–6000 K) that reign inside our planet. Outside these ranges, the used materials transform into different phases that should be present in the case of the investigated extrasolar planets. At the bottom of the Earth's mantle, where the pressure values are ∼120 GPa, the existence of a post-perovskite phase (∼1.5% denser than bridgmanite) has been observed (Oganov & Ono 2004; Murakami et al. 2004) and incorporated by Zeng & Sasselov (2013). They also considered another trans- formation of this new phase around 1 TPa. The impact on their results is however small compared to simpler models. The budget of light elements in the Earth's core is still a matter of discussion. In addition to sulfur, the outer core may contain oxygen and silicate up to 5% and 3.6%, respectively (Badro et al. 2015). If a substantial amount of light element (in addition to sulfur) also en- ters the composition of super-Earths' cores, these cores may be less dense than assumed in our models. As a consequence, the fraction of water required to explain the observed mass of super-Earth may be significantly reduced. The presence of silicon in particular would add a level of degeneracy to the composition of studied plan- ets, as the value of the bulk Fe/Si ratio would not be enough to set the planet's core mass fraction. Therefore, the impact of the presence of volatiles on the computa- tion of a planet's radius has to be studied thoroughly, to determine if this parameter has a significant effect on the radius of exoplanets. At the top of the mantle, on the other hand, olivine and pyroxenes may not be stable if the thickness of the outer ice layer is such that the pressure at its bottom exceeds 25 GPa (Irifune 1987). This would leave a man- tle entirely composed of bridgmanite and ferro-periclase. Alternatively, if pressure is slightly lower than 25 GPa, an upper mantle may exist but would be very thin. Con- vection within this layer would be weak or may not op- erate at all, which would strongly alter the cooling and thermal state of such ocean planets. Here, for CoRoT- 7b and Kepler-10b, we compute that the upper mantle is no longer present for WMF approaching ∼10%. A 5% WMF yields a ∼160 km thick upper mantle in both cases, whereas this value goes down to ∼15 km for 8% water in mass. In comparison, the Earth's upper mantle is about 700 km thick. In addition, as was proposed for icy moons of the solar system (Deschamps et al. 2010), ice layers may be composed of a mix of water and volatile compounds by a few percent in mass. While the pres- ence of volatiles may only slightly affect the density of the water ice VII layer (and thus the mass-radius rela- tionships), it may change physical properties more sub- stantially, in particular thermal conductivity (Hsieh & Deschamps 2015), and may add to the alteration of heat transfer from the presence of a thin upper mantle. As for silicate rocks, water ice presents high-pressure phases that have been measured through laboratory ex- periments, such as the ice VII–ice X transition around 70 GPa, ice X having a behavior still close to that of ice VII (Sotin et al. 2007). Above these pressures, new phases have been predicted by theory around 1 TPa (Mil- itzer & Wilson 2010). These pressures cannot be reached in the case of ocean planets up to 10 M⊕ (the value reached at the bottom of the ice VII layer for these plan- ets is ∼400 GPa). However they may appear in the case of icy giant planets such as Uranus or Neptune. Exoplanets may also be formed from materials differ- ent from those encountered in our own solar system. In particular, the silicate rocks we consider are based on oxygen, as our own protosolar nebula had a C/O ratio of ∼0.54 (Asplund et al. 2009). However, a small frac- tion of stars present higher C/O ratios (Brewer & Fischer 2016), suggesting the possibility of forming carbon-rich planets around such stars. For example, Madhusudhan et al. (2012) studied the interior of a possible carbon-rich 55 Cancri e, with the incorporation of pure carbon and silicon carbide, given the high C/O ratio of its host star 55 Cancri. Even if this latter ratio has been re-estimated at a lower value (0.78 ± 0.08; Teske et al. 2013), the existence of carbon-rich planets opens another branch of exoplanetary science. 4.2. Prospects As observed in Section 2.4, modeling the interior of ex- oplanets with three major main layers (here core, mantle, and water) generates a degeneracy on their relative frac- tions, as different sets of these fractions produce the same planet radius. Adding layers made of another material, such as an H/He atmosphere, would thus add a degree of degeneracy, as the isoradius curves in the compositional parameter space would become isoradius surfaces (the four-variables equivalent of the ternary diagram being a tetrahedron). Therefore, there is an important need to break the degeneracy existing with solid planets. We have shown here that this degeneracy can be significantly reduced under the assumption that the stellar Fe/Si and Mg/Si ratios are similar to those of the planet (Thiabaud et al. 2015). However, the planet-hosting stars whose ra- tios are known from high-resolution spectroscopy repre- sents a small fraction of the ∼3500 currently confirmed planetary systems4. There is thus an important need to measure the elemental abundances of these stars along with the physical and orbital parameters of the systems. By coupling these measurements to the new generation of space-born observatories such as PLATO or CHEOPS, which will supply physical and orbital parameters of ex- oplanets around bright stars with an unrivaled precision, we should be able to significantly reduce the set of com- positions allowed for a given exoplanet, and thus provide with precise ranges of its core and water mass fractions. The project leading to this publication has received funding from Excellence Initiative of Aix-Marseille Uni- versity - A*MIDEX, a French "Investissements d'Avenir" program. O.M. and M. D. also acknowledge support from CNES. We thank an anonymous referee for her/his very constructive comments that helped us strengthen our manuscript. REFERENCES Airapetian, V. S., Glocer, A., Khazanov, G. V., et al. 2017, ApJ, 836, L3 All`egre, C. J., Poirier, J.-P., Humler, E., & Hofmann, A. W. 1995, Earth and Planetary Science Letters, 134, 515 Anglada-Escud´e, G., Amado, P. J., Barnes, J., et al. 2016, Nature, 536, 437 Asplund, M., Grevesse, N., Sauval, A. J., & Scott, P. 2009, ARA&A, 47, 481 Badro, J., Brodholt, J. P., Piet, H., Siebert, J., & Ryerson, F. J. 2015, Proceedings of the National Academy of Science, 112, 12310 Barros, S. C. C., Almenara, J. M., Deleuil, M., et al. 2014, A&A, 569, A74 Batalha, N. M., Borucki, W. J., Bryson, S. T., et al. 2011, ApJ, 729, 27 Benz, W., Slattery, W. L., & Cameron, A. G. W. 1988, Icarus, 74, 516 Benz, W., Anic, A., Horner, J., & Whitby, J. A. 2007, Space Sci. Rev., 132, 189 Bonfils, X., Delfosse, X., Udry, S., et al. 2013, A&A, 549, A109 Brewer, J. M., & Fischer, D. A. 2016, ApJ, 831, 20 Brugger, B., Mousis, O., Deleuil, M., & Lunine, J. I. 2016, ApJ, 831, L16 Bruntt, H., Deleuil, M., Fridlund, M., et al. 2010, A&A, 519, A51 4 exoplanetarchive.ipac.caltech.edu 9 Cameron, A. G. W. 1985, Icarus, 64, 285 Cohen, R. E., Gulseren, O., & Hemley, R. J. 2000, American Mineralogist, 85, 338 de Pater, I., & Lissauer, J. J. 2015, Planetary Sciences, by Imke de Pater , Jack J. Lissauer, Cambridge, UK: Cambridge University Press, 2015 Deschamps, F., Mousis, O., Sanchez-Valle, C., & Lunine, J. I. 2010, ApJ, 724, 887 Dorn, C., Khan, A., Heng, K., et al. 2015, A&A, 577, A83 Dorn, C., Hinkel, N. R., & Venturini, J. 2017, A&A, 597, A38 Dressing, C. D., & Charbonneau, D. 2015, ApJ, 807, 45 Dumusque, X., Bonomo, A. S., Haywood, R. D., et al. 2014, ApJ, 789, 154 Dziewonski, A. M., & Anderson, D. L. 1981, Physics of the Earth and Planetary Interiors, 25, 297 Fortney, J. J., Marley, M. S., & Barnes, J. W. 2007, ApJ, 659, 1661 Frank, M. R., Fei, Y., & Hu, J. 2004, Geochim. Cosmochim. Acta, 68, 2781 Hama, J., & Suito, K. 1996, Journal of Physics Condensed Matter, 8, 67 Harder, H., & Schubert, G. 2001, Icarus, 151, 118 Haywood, R. D., Collier Cameron, A., Queloz, D., et al. 2014, MNRAS, 443, 2517 Hsieh, W.-P., & Deschamps, F. 2015, Journal of Geophysical Research (Planets), 120, 1697 Irifune, T. 1987, Physics of the Earth and Planetary Interiors, 45, 324 Kipping, D. M., Cameron, C., Hartman, J. D., et al. 2017, AJ, 153, 93 Lebrun, T., Massol, H., Chassefi`eRe, E., et al. 2013, Journal of Geophysical Research (Planets), 118, 1155 Lee, E. J., Chiang, E., & Ormel, C. W. 2014, ApJ, 797, 95 L´eger, A., Rouan, D., Schneider, J., et al. 2009, A&A, 506, 287 L´eger, A., Grasset, O., Fegley, B., et al. 2011, Icarus, 213, 1 Madhusudhan, N., Lee, K. K. M., & Mousis, O. 2012, ApJ, 759, L40 Militzer, B., & Wilson, H. F. 2010, Physical Review Letters, 105, 195701 Morgan, J. W., & Anders, E. 1980, Proceedings of the National Academy of Science, 77, 6973 Murakami, M., Hirose, K., Kawamura, K., Sata, N., & Ohishi, Y. 2004, Science, 304, 855 Oganov, A. R., & Ono, S. 2004, Nature, 430, 445 Queloz, D., Bouchy, F., Moutou, C., et al. 2009, A&A, 506, 303 Ribas, I., Bolmont, E., Selsis, F., et al. 2016, A&A, 596, A111 Rogers, L. A., & Seager, S. 2010, ApJ, 712, 974 Santos, N. C., Adibekyan, V., Mordasini, C., et al. 2015, A&A, 580, L13 Schaefer, L., & Fegley, B. 2009, ApJ, 703, L113 Schubert, G., Ross, M. N., Stevenson, D. J., & Spohn, T. 1988, Mercury, University of Arizona Press, 429 Seager, S., Kuchner, M., Hier-Majumder, C. A., & Militzer, B. 2007, ApJ, 669, 1279 Sotin, C., Grasset, O., & Mocquet, A. 2007, Icarus, 191, 337 Sotin, C., Jackson, J. M., & Seager, S. 2010, Exoplanets, 375 Spohn, T., Sohl, F., Wieczerkowski, K., & Conzelmann, V. 2001, Planet. Space Sci., 49, 1561 Stacey, F. D. 1992, Physics of the Earth., by Stacey, F. D.. Brookfield Press, Kenmore, Brisbane (Australia), 1992, 525 p., ISBN 0-646-09091-7 Stacey, F. D. 2005, Reports on Progress in Physics, 68, 341 Teske, J. K., Cunha, K., Schuler, S. C., Griffith, C. A., & Smith, V. V. 2013, ApJ, 778, 132 Thiabaud, A., Marboeuf, U., Alibert, Y., Leya, I., & Mezger, K. 2015, A&A, 580, A30 Tobie, G., Lunine, J. I., & Sotin, C. 2006, Nature, 440, 61 Tuomi, M., Jones, H. R. A., Barnes, J. R., Anglada-Escud´e, G., & Jenkins, J. S. 2014, MNRAS, 441, 1545 Valencia, D., O'Connell, R. J., & Sasselov, D. 2006, Icarus, 181, 545 Valencia, D., Sasselov, D. D., & O'Connell, R. J. 2007, ApJ, 656, 545 Valencia, D., Sasselov, D. D., & O'Connell, R. J. 2007, ApJ, 665, 1413 Valencia, D., O'Connell, R. J., & Sasselov, D. D. 2009, Ap&SS, 322, 135 10 Distribution of the different materials (level 1) and chemical species (level 2) in the five layers Table 2 Layer 1. Core 2. Lower mantle 3. Upper mantle 4. Ice VII 5. Liquid water Level 1 Level 2 100% Metal 1 − y1 Fe y1 FeS x2 Bridgmanite 1 − y2 y2 1 − x2 Periclase 1 − y2 y2 x3 Olivine 1 − y3 y3 1 − x3 Enstatite y3 1 − y3 FeSiO3 MgSiO3 FeO MgO Fe2SiO4 Mg2SiO4 Fe2Si2O6 Mg2Si2O6 100% Ice VII ∅ 100% Liq. H2O ∅ Valencia, D., Ikoma, M., Guillot, T., & Nettelmann, N. 2010, Weiss, L. M., Rogers, L. A., Isaacson, H. T., et al. 2016, ApJ, A&A, 516, A20 Vinet, P., Rose, J. H., Ferrante, J., & Smith, J. R. 1989, Journal of Physics Condensed Matter, 1, 1941 Wagner, W., & Pruss, A. 2002, Journal of Physical and Chemical Reference Data, 31, 387 819, 83 Zeng, L., & Seager, S. 2008, PASP, 120, 983 Zeng, L., & Sasselov, D. 2013, PASP, 125, 227 APPENDIX SUPPLEMENTARY MATERIAL Detailed composition of a planet To fix the distribution of chemical species in the different layers of the planet (e.g. Fe and FeS in the core, and the different silicate rocks in the mantles; see Section 2.1), we introduce the mole fractions xi and yi (with i ∈(cid:74)1, 5(cid:75) the in the lower and upper parts of the mantle (respectively layers 2 and 3). If(cid:0) X The mantle is assumed to be chemically homogeneous, meaning that the Fe/Si and Mg/Si mole ratios are identical i is the value of the X/Y mole ratio in number of the layer), as detailed in Table 2. The use of two variables allows to manage two levels of distribution. (cid:1) Y the layer i, then: (cid:19) (cid:18) Fe (cid:18) Mg (cid:19) Mg 2 = (cid:19) (cid:18) Fe (cid:18) Mg (cid:19) Mg = Si 2 Si 3 ⇐⇒ y2 = y3 3 =⇒ x2 = 1 − x3 2 (A1) (A2) To link the mole fractions listed in Table 2 to the composition of a planet, we define the following parameters: • falloy the fraction of iron alloy in the core; (cid:17) • (cid:16) Mg • Mg# ≡(cid:16) Mg Si P (cid:17) the planet's overall Mg/Si mole ratio; Mg+Fe the Mg number (Sotin et al. 2007). Mantle The Mg number Mg#, which reflects the amount of iron in the mantle, expresses the degree of differentiation of a planet. For instance, the Earth (Mg# = 0.9; Sotin et al. 2007) is more differentiated than Mars (Mg# = 0.7; Sotin et al. 2007) due to their difference in mass, Mars being ten times less massive. From Equations A1 and A2, and following the definitions of the three aforementioned parameters, we obtain:  (cid:16) Mg (cid:17) Si x1 = 1 x2 = Mg#/ x3 = 2(1 − x2) x4 = 1 x5 = 1 (cid:16) Mg (cid:17) Si P P and  y1 = falloy y2 = 1 − Mg# y3 = y2 y4 = 0 y5 = 0 Thus, the internal structure of a planet can be entirely described by six compositional parameters: the planet's mass MP, the core mass fraction (CMF) xcore and water mass fraction (WMF) xwater, the fraction of iron alloy in the core falloy, the Mg/Si mole ratio of the planet , and the Mg number Mg#. Equations Once the composition of a planet is fixed by the six aforementioned parameters, the model has to simulate its interior. We define a one-dimensional spatial grid, with fixed precision, that ranges from the center of the planet to above its total radius. The internal structure of a planet is governed by its gravitational acceleration g, pressure P , temperature T , and density ρ inside the body. The four quantities are computed for every point of the grid, by solving the canonical equations of internal structure for solid bodies. The gravitational acceleration is computed from the Gauss theorem: 11 (A3) with m the mass at a given radius r, and G the gravitational constant. Thus, at a radius r ∈]Ri, Ri+1] of the spatial grid: dg dr = 4πGρ − 2Gm r3 (cid:90) r (cid:18) Ri (cid:19)2 g(r) = 4πG r2 r2ρ dr + g(Ri) where Ri is the lower radius of a layer i ∈(cid:74)1, 5(cid:75) (plus R6 = RP the planet total radius), i.e. R1 = 0. The pressure P is then computed from the planet surface assuming the hydrostatic equilibrium: Ri r (A4) (cid:90) Ri+1 ρg dr = −ρg dP dr =⇒ P (r) = P (Ri+1) + as well as the temperature T , if we consider an adiabatic profile: r dT dP = γT ρΦ =⇒ dT dr = −g γT Φ =⇒ T (r) = T (Ri+1) exp via the use of the Adams-Williamson equation: with γ and Φ the Gruneisen and seismic parameters, respectively: dρ dr = − ρg Φ γ = γ0 Φ = (cid:18) ρ0 (cid:19)q ρ = dP dρ KS ρ (cid:34)(cid:90) Ri+1 r (cid:35) γg Φ dr (A5) (A6) (A7) (A8) and γ0 the reference Gruneisen parameter, ρ0 the density at ambient conditions (kg/m3), q the adiabatic power exponent, and KS the adiabatic bulk modulus (GPa). Following the approach of Sotin et al. (2007), the temperature profile is adapted to mimic the presence of thermal boundary layers at the top and bottom of each layer animated by convection, by fixing temperature drops between the different layers, whose values are taken from Earth models. Finally, the mass of the planet verifies: dm dr = 4πr2ρ =⇒ MP = (cid:90) Ri+1 5(cid:88) i=1 Ri r2ρ dr The computational scheme works as follows: the six input parameters MP, xcore and xwater, falloy, Mg# (see Section 2.1) are given, but only the latter three are used at first, since they fix the distribution of materials in the different layers. The model starts with a planet composed of the five layers (core to liquid water) with lower radii Ri fixed arbitrarily, and a homogeneous density fixed to the density of the corresponding material at ambient conditions ρ0(i). Then are computed the profiles of g, P , and T inside the planet using Equations A3–A8, followed by the profile of ρ computed using the corresponding equation of state of the material (see Section 2.2). g and ρ are computed from the center of the planet with an increasing radius, whereas the computation of P and T starts at the planet surface and is done in the opposite direction. This requires boundary conditions, namely: no central gravitational acceleration, surface pressure and temperature fixed to given values Psurf and Tsurf . These parameters allow to simulate the presence of an atmosphere, provided that its mass and height are negligible compared to MP and RP respectively (otherwise the gaseous atmosphere should be included in the model as a supplementary layer). This is the case of the Earth, where the atmosphere only accounts for 0.0001% of the planet's mass. From the profiles of these four quantities, we are then able to re-estimate the layers' lower radii Ri to fit the leftover input parameters: (cid:16) Mg (cid:17) Si P (A9) , and 12 - R1 = 0 by definition; - R2 gives the size of the core, so we fix it using the CMF: (cid:90) R2 0 ρ(x)r2(x) dx = xcoreMP (A10) - R3 is the boundary between the lower and upper mantles, which corresponds to the phase change of silicate rocks (from bridgmanite and ferro-periclase to olivine and enstatite; see Section 2.1). Following the work of Sotin et al. (2007), this phase change is well described by a Simon equation (Irifune 1987), giving: T (R3) = T0 + P (R3) − P0 a with T0 = 800 K P0 = 25· 109 Pa a = −0.0017· 109 Pa K-1 - as for the boundary between core and mantle, R4 is the limit between (upper) mantle and hydrosphere, thus we compute it using the mass fraction of the mantle (from the CMF and WMF): ρ(x)r2(x) dx = (1 − xcore − xwater)MP (A12) - R5 is then the radius at which the phase transition of ice VII to liquid water occurs, i.e. (see Section 2.3; Frank et al. 2004): P (R5) = P0 + a with  T0 = 355 K P0 = 2.17· 109 Pa a = 0.764· 109 Pa c = 4.32 (A11) (A13) (A14) (A15) (cid:90) R4 R2 (cid:20)(cid:18) T (R5) (cid:19)c − 1 (cid:21) T0 (cid:90) R6 - finally, R6 = RP the total radius of the planet, is fixed by the mass of the hydrosphere: ρ(x)r2(x) dx = xwaterMP R4 The latter two steps (computation of g, P , T , and ρ, and estimation of the Ri) are then repeated in an iterative scheme, until convergence is reached. Convergence is achieved when the changes of the Ri and the profiles of g, P , T , and ρ from one iteration to the other are lower than a fixed precision. Once the iterative process has stopped, the model is supposed to verify all input parameters of the planet, and can provide a planet total radius RP that respects the thermodynamic and elastic properties of the materials composing the planet. This computed radius can then be compared to the measured radius, if there is one known. We also have access to the interior profiles of g, P , T , and ρ, as represented on Figure 9 in the case of the Earth again. The transitions between the different layers are easily noticed through the discontinuities of the curves. As shown on Figure 9, simulating a 1 M⊕ planet with the Earth parameters (Table 1) produces a radius RP = 0.992 R⊕, i.e. ∼60 km less than the actual Earth. This less than 1% error is comparable to those obtained by other models (Sotin et al. 2007; Valencia et al. 2006), since none of these models considers the Earth's crust (a 10–50 km thick layer of low density), and also because several chemical elements are not incorporated into the modeling of the core and mantles (as Ca, Al, and Ni) for sake of simplicity. Here, to obtain a planet as big as the Earth (6371 km, with an error of 0.001%), the CMF has to be lowered to 0.286 (the value actually inferred for Venus; Stacey 2005), if all other parameters remain fixed to Earth values. Planetary Fe/Si ratio From our model, it is possible to compute output compositional parameters of the simulated planet, that were not in the set of input parameters. For instance, the Fe/Si mole ratio of the planet can be calculated using the following equation: (cid:18) Fe (cid:19) (cid:80)5 (cid:80)5 = Si P i=1 ni(Fe)i i=1 ni(Si)i with ni = Mi , where Mi is the mass of the layer i, and Mmol,i is the mean molecular mass of the material composing this layer. Here (X)i is the mole fraction of the element X in layer i. As for the planet radius, the computed Fe/Si ratio can be compared to a measured value, or provide an estimation when the latter is unknown. Mmol,i However, it is interesting to note that in our case, the Fe/Si ratio of the simulated planet can be derived analytically: 13 Figure 9. From top to bottom: gravity acceleration g, pressure P , temperature T , and density ρ profiles computed inside a planet of 1M⊕ with an Earth-like composition, once the model has reached convergence. Vertical lines show the boundaries between the different layers that compose the planet. 0 2 4 6 8 10 12 0 0.2 0.4 0.6 0.8 1 1.2 1.4Gravity acceleration (m⋅s-2)Radius (R⊕) 0 50 100 150 200 250 300 350 400 0 0.2 0.4 0.6 0.8 1 1.2 1.4Pressure (GPa)Radius (R⊕) 0 500 1000 1500 2000 2500 3000 3500 4000 4500 0 0.2 0.4 0.6 0.8 1 1.2 1.4Temperature (K)Radius (R⊕) 0 2000 4000 6000 8000 10000 12000 14000 0 0.2 0.4 0.6 0.8 1 1.2 1.4Density (kg⋅m-3)Radius (R⊕) 14 (cid:18) Fe (cid:19) Si P = (cid:18) Mg (cid:19) 1 Mg# Si P (cid:34) 1 − Mg# + M1 Mmol,1 + 2 M3 Mmol,3 M2 Mmol,2 (cid:35) (A16) Yet, there is a simple relation between the molecular masses of the mantles, namely Mmol,3 = 2Mmol,2. This comes from our assumptions that the mantle is chemically homogeneous (see Equations A1–A2), but also from the molecules present in these layers (bridgmanite, olivine, ferro-periclase, and enstatite), whose molecular masses compensate each other. Eventually, we obtain: (cid:18) Fe (cid:19) (cid:21) (cid:18) Mg (cid:20) (cid:19) (cid:17) (cid:16) Mg P Si P 1 = Si P Mg# Si 1 − Mg# + Mmol,2 Mmol,1 xcore xmantle (A17) Interestingly, the Fe/Si mole ratio of a planet does not depend on the mass of the body, only on its compositional parameters. In particular, if the parameters falloy, , and Mg# are fixed, the Fe/Si ratio only depends on the CMF and WMF of the planet, and can thus be represented in the ternary diagram "core-mantle-water" (see Section 2.5). The thermodynamic and elastic parameters used in the EOS hereafter detailed, that describe the behavior of each material composing the planet, are taken from Sotin et al. (2007) and Sotin et al. (2010). Equations of state Third-order Birch-Murnaghan (BM3) 3(cid:35)(cid:40) (cid:19) 5 (cid:18) ρ ρT,0 1 − 3 4 (4 − K(cid:48) T,0) (cid:34)(cid:18) ρ ρT,0 (cid:35)(cid:41) (cid:19) 2 3 − 1 (A18) P (ρ, T ) = 3 2 KT,0 (cid:34)(cid:18) ρ (cid:19) 7 ρT,0 with 3 −  (cid:18)(cid:90) T KT,0 = K0 + aP (T − T0) T,0 = K(cid:48) K(cid:48) ρT,0 = ρ0 exp α(t) dt α(T ) = aT + bT T − cT T −2 T0 0 (cid:19) where T0, ρ0, K0, K(cid:48) temperature derivatives of the bulk modulus, and thermal expansion coefficients, respectively. 0, aP , and {aT , bT , cT} are the reference temperature, density, bulk modulus, pressure and Mie-Gruneisen-Debye (MGD)  P (ρ, T0) = 3 2 ∆Pth = 9 γnR V θ3 (cid:17)−q (cid:16) ρ P (ρ, T ) = P (ρ, T0) + ∆Pth (cid:34)(cid:18) ρ (cid:19) 7 (cid:90) θ ρ0 T T 4 0 K0 (cid:34) 3 − (cid:18) ρ ρ0 3(cid:35)(cid:40) (cid:19) 5 (cid:90) θ T0 t3 et − 1 dt − T 4 0 0 t3 et − 1 dt 1 − 3 4 (4 − K(cid:48) 0) (cid:35) with (cid:16) ρ (cid:17)γ (cid:35)(cid:41) (cid:19) 2 (cid:34)(cid:18) ρ ρ0 3 − 1 (A19) with θ = θ0 parameters, scaling exponent, and number of atoms per chemical formula, respectively. and γ = γ0 , θ0, γ0, q, and n being the reference Debye temperature and Gruneisen ρ0 ρ0 (cid:34)(cid:18) ρ ρ0 (cid:19) 2 3 − (cid:18) ρ ρ0 Vinet 3(cid:35) (cid:19) 1 (cid:40) exp (K(cid:48) 0 − 1) 3 2 3(cid:35)(cid:41) (cid:19)− 1 (cid:34) 1 − (cid:18) ρ ρ0 P (ρ, T0) = 3K0 (A20) with addition of the thermal pressure ∆P , as for the MGD formulation.
1007.2099
1
1007
2010-07-13T13:04:59
The "Sun-climate" relationship : III. The solar flares, north-south sunspot arrea asymmetry and climate
[ "astro-ph.EP", "physics.ao-ph" ]
In this last Paper III additional evidences that the solar high energetic particles radiation with energies higher as 100 MeV (the solar cosmic rays SCR) is an very important component for the "Sun- climate" relationship are given (see also Paper I and II). The total solar irradiance (TSI) and the galactic cosmic rays (GCR) variations given an integral climate effect of cooling in sunspot minima and warming in the sunspot maxima. Unlike the both ones the powerful solar corpuscular events plays a cooling climate role during the epochs of their heigh levels. By this one subcenturial global and regional temperature quasi- cyclic changes by duration of approximately 60 years could be track during the last 150 years of instrumental climate observations . It has been also evided in the paper that this subcenturial oscilation is very important in the Group sunspot number (GSN) data series since the Maunder minimum up to the end of 20th century. Thus the solar erruptive activity effect make the total "Sun -climate" relationship essentially more complicated as it could be follow when only the TSI and GCR variations are taken into account. In this light the climate warming tendency after AD 1975 is rather by a natural as by an antropogenic origin. Most probably the last one is very close related to the general downward tendency of erruptive solar events which is superimposed over the high long term TSI levels during the last three decades (AD 1975-2007). It is evided, that the efficiency of the solar corpuscular activuty over the climate is strongly depended by the "north-south" asymmetry of the solar activity centers (as a proxy the sunspots area north-south asymmetry index A is used there). The climate cooling effect in the Northern hemisphere is most powerful during the epochs of positive values of A.
astro-ph.EP
astro-ph
Bulgarian Astronomical Journal v.13, 2010 THE “SUN - CLIMATE” RELATIONSHIP : III. THE SOLAR ERRUPTIONS, NORTH-SOUTH SUNSPOT ARREA ASSYMETRY AND CLIMATE Boris Komitov Bulgarian Academy of Sciences- Institute of Astronomy, 6003 Stara Zagora-3, POBox 39, [email protected] In this last Paper III additional evidences that the solar high energetic particles radiation with energies higher as 100 MeV (the solar cosmic rays SCR) is an very important component for the “Sun- climate” relationship are given (see also Paper I and II). The total solar irradiance (TSI) and the galactic cosmic rays (GCR) variations given an integral climate effect of cooling in sunspot minima and warming in the sunspot maxima. Unlike the both ones the powerful solar corpuscular events plays a cooling climate role during the epochs of their heigh levels. By this one subcenturial global and regional temperature quasi- cyclic changes by duration of approximately 60 years could be track during the last 150 years of instrumental climate observations . It has been also evided in the paper that this subcenturial oscilation is very important in the Group sunspot number (GSN) data series since the Maunder minimum up to the end of 20th century. Thus the solar erruptive activity effect make the total “Sun –climate” relationship essentially more complicated as it could be follow when only the TSI and GCR variations are taken into account. In this light the climate warming tendency after AD 1975 is rather by a natural as by an antropogenic origin. Most probably the last one is very close related to the general downward tendency of erruptive solar events which is superimposed over the high long term TSI levels during the last three decades (AD 1975-2007). It is evided, that the efficiency of the solar corpuscular activuty over the climate is strongly depended by the “north-south” assymetry of the solar activity centers (as a proxy the sunspots area north-south assymetry index A is used there). The climate cooling effect in the Northern hemisphere is most powerful during the epochs of positive values of A. This effect is very significant in combination with high level of the GSN-index . A strong quasi 120 year “hypercycle” has been detected in the A index during the period of AD 1821-1994. Most probably 10 the observed 120-130 yr cyclity in climate and Be continental ice core data (both “Greenland and “Antarctic” series) is related to the last one. The expected climate changes during the next decades and especially during the new solar sunspot cycle No 24 are discussed on the base of the “multiple” nature of the “Sun- climate” relationship. 1.Introduction According the most perceived point of view the “Sun-climate” relationship during the present postglacial era (Holocene, the historical time scale) is realized predominantly by the total solar irradiance (TSI) variations (Solanki , 2002; de Jaeger and Usoskin, 2006 ). The TSI index is well known since AD 1978 on the base of satellite observations (Frolich et al, 1997; Pap et al. , 2003 ) The last one during the last ~ 400 years corresponds very well to the overall sunspot activity (the International Wolf’s number Ri and the Group sunspot number (GSN or Rh)( Lean et al., 1995; Lean, 2000, 2004). There are also a significant number of theoretical (numerical) , mixed type (statistical + theoretical) or “poor” statistical studies in which the relationship “sunspot activity - solar magnetic flux -> TSI” is investigated (Lean et al,2000; Solanki et al, 2002; Krivova et al., 2007 etc.). On other hand there are evidences that an additional mechanism of indirect Sun’ s forcing over the climate due to the modulation of galactic cosmic rays (GCR) by solar wind exist. The first works in this course are still from the middle of 1970st (Dickinson, 1975). The aerosols and clouds production rates forcing in the lower atmosphere under the GCR-flux increasing during the sunspots minimuma epochs is discussed by Svenmark and Friiz-Christiensen (1997) and Yu (2002) . There are also some interesting results of Tinsley (2000), concerning the GCR-flux influence over the atmospheric electricity and circulation. It has been marked by many authors that the “overall sunspot activity - TSI -> climate” relationship is far not enough to explain the real climate dynamics during the last 400 years since AD 1610 . As it is pointed out by Thompson (1997) only 25% of the global warming effect after AD 1850 could be explained due to the TSI increasing during the this time. For the other 75% should be search for additional factors. Especially after AD 1975/80 there is a total divergence between the TSI and global temperature changes (Solanki, 2002; Usoskin et al.,2005; Lockwood and Frolich, 2007). The phenomena couldn’t to explain satisfactory even if in addition the GCR-flux is taken into account. This is why for the last 30-35 years by the opinion of many researchers the human activity is the factor , which play the dominant role for the climate changes. It has been shown in the first paper of this series (see Paper I), that the residual variations to the regressional models “sunspot activity – temperature data” both for the Northern hemisphere (AD 1610 -1979) (Moberg et al., 2005) and for the World Ocean (1856-1995) (Parker et al., 1995) are far not occasional. There are well expressed cyclic oscilations in the quasi-centurial and subcenturial range. The spectra of the last ones is more complicated in the Northern hemisphere “residual” data series ( powerful cycles by duration of 54-67 (doublet) and 120-130 years), while in the World Ocean one there is only a strong cyclic 58-63 year oscilation (doublet) as well as essentially weaker trace of 88 year one. It has been summarized finally in Paper I that there is powerful quasi- 60 year climatic cycle in the modern epoch . The last one plays a very important role in the climate , causing few waves of cooling and warming since the end of Maunder minimum , which are superimposed over the general regressional relationship “sunspot activity –temperature data” during this time. It has been also shown in Paper I that the climate warming epoch after AD 1975 up to 2005-2006 well correspond to the serial upward phase of this 60 year cycle. It has been found in the next Paper II that a very powerful quasi –60 year cycle exist both in the middle latitude aurora (MLA) (Krivsky and Pejml, 1988) as well as in the “Greenland” 10Be data series (Beer et al., 1990,1998). It has been shown that there is a very good coincidience by time between the corresponding 60-year cycle extremums in the both series. The local 60 yr maximums in MLA and 10Be series during the last 300 years since AD 1700 correspond to subcenturial temperature minimums . They are well expressed in the both studied temperature series, but essentially better in the World Ocean ones. The MLA events occurs in the upper Earth atmosphere and by this one they are strongly independent from the troposphere processes and the climate. Their primary sources are active events such as the coronal mass ejections (CME). Consequently this is related to the 60 year cycle in this series too. Due to this fact it has been concluded in Paper II that : 1 The quasi –60 year 10Be cycle is most probably by solar origin and it is caused by an yield of solar high energetic protons E> 100 MeV in the total production rate of this “cosmogenic” isotope in the stratosphere; 2. The quasi 60 year climate cycle is caused by the same one high energetic solar corpuscular events; 3. The increasing of the solar high energetic particles ( probably mainly protons) fluxes lead to the same effects in Earth atmosphere such as the galactic cosmic rays with the same energies, i.e. an increasing of the aerosols production rate and cloudness and as a final effct – to a climate cooling. The aim of this last paper (Paper III) is to make a more detailed analysis and to give additional arguments for the important role of erruptive solar processes as a climate forcing factor. A strong evidence that a very important role there the north- south assymetry of the erruptive events is also played is given. Evidences in this course that the modern climate changes are not by antropogenic origin could also given by comparuson of Mars climate conditions during the last three decades. It has been found that the total albedo of this planet has been falled down from AD 1977 up to 1999 , which correspond to a warming in order of 0.6 K only for about of two decades (). It is demonstrated that the total “Sun-climate” relationship is much more complicated as it is follows if only the TSI and GCR flux changes are taken into account. However, in the same time it is much better fit to the real observed climate variations both in the presence and in the past. 2. Data and Methods The following data sets are used in this study: - The Northern hemishpere temperature data series from AD 1610 to 1979 – the last 370 years from the data set of Moberg et al. (2005). As a “zero level” the mean temperature between AD 1961 and 1990) is used there - The World Ocean temperature data series from AD 1856 to 1995 (Parker et al.,1995). The “zero level” there is the mean temperature in AD 1940. - The middle latitude aurora (MLA) annual number data series from AD 1700 to 1900 , i.e. the last two centuries from the catalogue of Krivsky and Pejml (1988) with most certain data. The catalogue data for the period AD 1000 to 1900 are published in the National Geophysical Data Center (ftp://ftp.ngdc.noaa.gov/STP/SOLAR_DATA/AURORA). - The index of north-south sunspot arrea assymetry A between AD 1821 and 1994, which is published in the Pulkovo observatory archive The main methods , which are used there are the T-R periodogramm analysis for detecting of cycles in the time series (see Paper I ) as well as a multiple correlation-regressional analysis. 3. The results and analysis 3.1 The temperature “residuals”, middle latitude aurora and the north-south sunspots area assymetry As a next step in our study (see also the previous Paper I and II) it should be to estimate how is the possible contribution of the solar “erruptive” component in the studied temperature series . For the better discovering of the effect the correlation between the middle latitude aurora (MLA) annual numbers (Aur_N) and the temperature “residual” series (∆∆22ΘΘ ) from the “sunspot activity –temperature” regressional model (4) in Paper I for the Northern hemisphere will be studied. As it has been mentioned in Paper II the using of MLA data is limited between AD 1700 and 1900 by two circumstances, namely: 1. The very possible serious lack of data before AD 1700; 2. The basic catalogue don’t contain data for the 20th century. It should to note that as a solar erruptive activity proxy the MLA annual number Aur_N has a disadvantage – it is very crude quantitative indicator of the total energy of these processes and consequently , of the penetrating in the Earth atmosphere high energetic solar corpuscular fluxes too. However there is no better proxy for these before the era of their instrumental observations. The coincidience between the middle latitude auroral activity maximums and the subcenturial local temperature minimums as well as the opposite events too is shown on fig.1. The all epochs of positive changes (warming) by mean duration of ~25-30 years are well corresponded to MLA fadding tendencies. In contrary, the cooling tendencies are dominating during the periods of the auroral activity increasing. As a result there is a well visible quasi- subcenturial (50-70 yr) “cooling-warming” cycle in approximately antiphase to the corresponding “auroral” cycle . Thus by fig.1 the conclusions for the reversed relationship between the high energetic solar corpuscular radiation and the Northern hemisphere temperatures is confirmed and visualized. The coefficient of linear correlation between the both smoothed series is r = - 0.43. It is about 7.5 times larger as its error and correspond to very high statistical significance (the “zero hypothesiss” probability there is P < 10-6). The relationship is slightly better if a logarithmic type fitting for the relationship is used , i.e. ∆∆22ΘΘ = a*ln(AurN) + b . The correlation coefficient in this case is - 0. 45. Fig.1. Up: The the smoothing 11 year residual variations “∆∆22ΘΘ ” of Northern hemisphere temperature data (Moberg et.al.,2005) after the removing of the regressional model “Group sunspot numbers – Northern hemisphere temperature “(Paper I).The last one corrspond to the zero-level ((the dark horizontal line) ; Below : The smoothing 11 year annual MLA numbers. Because that the logarithnic relationship is less sensitive as the linear one it follow that the climate effect of the solar corpuscular events is significant if the activity of last ones is at enough high level. We have already note that the MLA annual number Aur_N is very rough proxy for this activity. It could not to estimate by Au_N how powerful are the separate MLA events , their corresponding solar sources and the corpuscular fluxes. This is why the so founded coefficient of correlation could be much better as the obtained above if a better proxy for the erruptive activity is used. Consequently the value of 20-25% from the total variance of the residual series, caused by the erruptive solar factor should be taken as the possible lower limit of the same one. Unfortunately there is no better proxy for the solar erruptive activity with enough long series before AD 1900 as the Aur_N . On other hand this relatively weak value of r indicate that there a very important factor (or factors), most probably, is no taken into account. Is the last one connected to the Sun or it is by terrestrial origin? Since the middle of 20th century there are a number of studies where the spatial distribution of the active regions on the solar disk as an important component of the “Sun-climate” relationship is considerated. The most useful proxy for such aims is the index A = S N − S S S N + S S (1) where SN and SS are the total sunspot areas in the Northern and Southern hemisphere of the Sun respectively. As it has been pointed out by Loginov(1973) by geometric causes the Sun Northern hemisphere should essentially more geoeffective as the Southern one. As a “geoefficience” the ability of the solar erruptions to force over the Earth magnetosphere and atmosphere , causing geomagnetic storms, aurora , ionospheric disturbances and other geophysical events plus in the stratosphere and troposphere too is there in view off. Consequently, if there a climate forcing by solar erruptive events is assumed, the including of the assymetry index A as an additional factor should be lead to much better model of the total solar effect over the “residual” series (the ∆∆22ΘΘ -values) , as if only the MLA annual number is taken into account. The new model should be of multi-factor type. The best of sunspot arrea north-south assymetry data series for our aims is published in the Pulkovo Observatory Extended Data Archive. It contain the mean annual data of A-index since AD 1821 up to 1994. As in the case of the all other data the 11- year smoothed values are used there. The plot of the both smoothed series of the assimetry index A and temperature residuals ∆∆22ΘΘ (the dotted line) are shown on fig.2. Fig2. The north-south sunspot arrea assymetry index A during the period AD 1821- 1994 (by the bold line) and the temperature “residual” ∆∆22ΘΘ series in Celsius /Kelvin/ degrees (11-year smoothed values). The nummerical values on the Y-axis for the both series are identical. It is clear visible that the strongest negative values of ∆∆22ΘΘ near to AD 1839-1840 and in the end of 1950th are in very good coincidience with the local maximims of A, when the values of the last one are strong positive. The local warming maximum near to AD 1940 correspond to local minimum of A too, but there the values of the last one are slightly > 0. There is only one period berween AD 1880-1910/1911 when the local cooling correpond to a weak local minimum of A. This period is interesting also by the last centurial solar minimum , which has been started at the end of 19th century. But generally there is a well expressed anticorrelation between the north-south sunspot arrea assymetry index and the temperature changes in the Northern hemisphere of the Earth between the deepest phase of the solar Dalton minimum and AD 1980. If the efficiency of the solar erruptive events over Earth climate depends by the north-south solar activity assymetry and the Northern hemisphere is essentially more geoeffective (Loginov, 1973), than the figures 1 and 2 are a good confirmation for the last one. The very deep local mimimum of of ∆∆2Θ Θ near AD 1840 correspond well the both maximums of MLA and A activity (fig,1 and 2). There are not catalogue data for the MLA in the middle of 1950th , but undoubtely the maximum of zurich cycle No 19 in AD 1957 correspond to a very high level of erruptive activiry. As it has been pointed out in Paper II the extrapolated maximums of MLA activity outside the end of Krivsky and Pejml catalogue data series should be near to AD 1910/1911 and 1975 ,while near to AD 1940 should be a minimum. This is well corresponding that the MLA activity during the cycles 18, 19 and 20 should be very high. By the combination with the strong maximum of A (a very expressive domination of Sun Northern hemisphere activity) in 1950th and remaining positive levels in 1960th the temporal climate cooling between AD 1940 and 1975/76 could be satisfactorily explained. This preliminary conclusion should be tested on the base of multi-factor correlation- regressional analysis. This part of the study has been provided on two stages. The Northern hemisphere: AD 1821-1900 On the first stage the period before AD 1900 has been investigated. This separation is taken due to the fact that before this calendar year the MLA annual number Aur_N could be used as a proxy for the solar high energetic events. The smoothed 11-year data series of A, Aur_N and Rh (the Group Sunspot Number) are used as possible factors for the changes of ∆∆22ΘΘ . The first step there has been to determine the coefficients of linear correlation r between the each pair of the parameters ∆∆22ΘΘ , A, Aur_N and Rh. They are correspondly: -0.407 for the pair ∆∆22ΘΘ and AurN, -0.722 for ∆∆22ΘΘ and A , and -0.560 for ∆∆22ΘΘ and Rh. The values of r between the factors Rh and A should also to estimate. The higher by module values are an indicator that they are not enough independent each from other. It is very probable in many of such cases to exclude one of the both factors from the multiple model even if the coefficient of correlation between the factor and the predictant is high. The rule in this case is that in the model remain this factor which is better corellated with the predictant. The coefficients r between the potencial factors are as followed:+0.530 for the pair A and Aur_N; + 0.621 for Aur_N and Rh, and +0.375 for Rh and A . The all obtained values of r are statistically significant with probability >99.9%. It need to point out the very high anticorrelation between the “residual“ temperature data ∆∆22ΘΘ and the sunspot arrea assymetry index A. On other hand there is a relative weak relationship between Rh and A and good correlation between Rh and ∆∆22ΘΘ . The last one is better as the linear or logarythmic relationship between Aur_N and ∆∆22ΘΘ (r= -0.407). . However the correlation between Aur_N and Rh is higher (r=+0.621) as between A and Aur_N (r=+0.53). These results shown that most probably in the multiple regression models the relationship between ∆∆22ΘΘ and MLA could be totally captured and described by the terms describing A and Rh or their interraction  A large number of one-, two- and three factor regressional models , including also different non-limear terms, has been obtained. The multiple coefficient of correlation R and the Snedekor-Fisher’s F-test has been used for the selecting of the best of them. It has been found that the best of the all is: ∆∆22ΘΘ = 0.07625 + 0.499*A –1.224*A2 – 1.943.10-8*Rh4 –0.02334*A*Rh R=0.888; F=4.45 (2) The first important feature of this formula is the absence of any term, comtaining the annual number of middle latitude aurora , i.e Aur_N. The model (2) is a multiple function of two parameters - the group sunspot number Rh and the sunspot area assymetry index A. Obviously the influence of the solar high energetic corpuscules , for which the auroral activity index Aur_N as a proxy has used to this momennt , is better aproximated by the nonlinear terms of the types Rh4 and A*Rh. Both coefficients of these terms are negative. It is provided by the Rh4 term that there is a small climate cooling effect in the range of 0.2K , if the sunspot activity is very high (the smoothed 11-year Rh value is > 80- 100). The strong nonlinearity of this term expresses by our opinion the fact that during the investigated period the intensity of the most powerful erruptive events roughly correspond to the higher levels of sunspot activity. The “interactial” two- factorial A*Rh term is much more interesting. The sign of its climate effect depend by the sign of A: A negative value of the sunspot arrea assymetry correspond to a warming, while in the cases of higher sunspot activity in the Sun Northern hemisphere should be related to a climate cooling effect. If we use typical values for A =0.15 and Rh=50 the mean total amplitude effect over the ∆∆22ΘΘ values is approximately 0.02334 x 0.15 x 50 x 2 = 0.35K. It expresses well the typical maximal deviations of the real smoothed 11-year Northern hemisphere and World Ocean temperatures to the corresponding “sunspot activity –temperature “ models (Paper I). On this base it could argued that the “interactial” term gives in the most of cases the main yield for the ∆∆22ΘΘ magnitude at least in the 19th century. The last conclusion could be confirmed when an estimation of the both “pure” A-terms is made. The nonlinear A2 -term gives always cooling effect, but it is significant only at high by module values of A (>0.2-0.3), while the linear term lead to cooling effect in range of 0.1K if A= - 0.2, or warming if A= +0.2. Thus the both “pure” A- terms are more important only during of the epochs of a very low sunspot activity when Rh tend to zero. 2, where St 2 is the total 2/ S0 The F-parameter used there is defined as F= St 2 +the residual variance S0 2). Consequently for the variance (the factors variance Sf 2 = 4.45 –1 =3.45, i.e the both factors partition in the total variance is model (2) Sf 3.45/4.45 =0.78 .Thus there is an evidence that 78% from the variations of ∆∆22ΘΘ during the epoch AD 1821-1900 are by a certain solar origin and only 22%- by other not taken into account factors or ocasional data errors. The Northern hemisphere: AD 1913-1979 As it has been already noted above, the catalog of Krivsky and Pejml ended at AD 1900 . On other hand it has been pointed out by our multiple regressional analysis that the index of assymetry and non-linear sunspot activity terms in the model are better proxy of the solar corpuscular events over the Northern hemisphere temperatures as the annual numbers of the middle latitude aurora. This is why for the multiple regressional analysis of the “residual” ∆∆22ΘΘ temperature series during the 20th century only the A and Rh indicies as factors has been used. The aim is not only to estimate the yield of the solar corpuscular activity for the climate changes during the first 7-8 decades of the          previous century , but also compare the potental evolution of the relationship with the same one during the 19th century. It has been decide to choice as a start year AD 1913, because of the fact that by many authors has been started the new centurial solar cycle after the deepest phase of the Gleissberg-Gnevishev’s centurial minimum (AD 1898-1923). This is the middle moment of the pair zurich cycles 14-15. On other hand near to this date is a breakpoint for the long time upward trend in the ∆∆22ΘΘ data series, which has been started very soon after the end of the Dalton minima near to AD 1839/40. The comparison of the pair coefficients of linear correlation r show for three significant differences in relation to the period AD 1821-1900: 1. The coefficient of correlation for the pair ∆∆22ΘΘ and Rh is r = - 0.793 vs –0.560 for the 19th century; 2. The corresponding coefficient is r= -0.26 between A and ∆∆22ΘΘ vs –0.722 (19th century), but it remain statistically significant at level >95% ; 3. For the pair A and Rh the coefficient r falls dramatically from –0.375 up to -0.008, i.e. unlike the 19th century a real relationship between the sunspots arrea north-south assymmetry and the sunspot activity is absent during the studied epoch. By our opinion the total independence of the both solar activity indicies each from other is the main cause for the changes of the other above mentioned relationships. It has been found that the best from the all tested multiple regressional model is expressed by the formula: ∆∆22ΘΘ =0.715 – 0.965*A +1 .537*A2 – 0.00852*Rh (3) R = 0.861 ; F = 3.67 There is no “interactial” A*Rh term in this formula, which is easy to explain- the smoothed assymetry index A is positive in the almost all studied interval , except only one smoothed value near to AD 1933. It need to remember, that the yield of such term strongly depend no only by the sunspot actvity level , but also by in what Sun hemisphere this activity is predominated. The factor variance in (3) is equal to 2.67/3.67 = 0.72, i.e. about 72% from the total variance of ∆∆22ΘΘ . The World Ocean “residual“ series (AD 1856-1994) The Pulkovo sunspot assymetry data series is ended at AD 1994. This is why the last AD 1995 has been excluded in the provided multiple regressional analysis for the World Ocean residual temperature data series. It need also to remember that the general statistical relationship “Group Sunspot Number – World Ocean temperature” is much closer as the corresponding one for the Northern hemisphere (r= + 0.877 , see formula (5) in Paper I). By the multiple regressional analysis procedure for the “residual” World Ocean temperature variations ∆∆22θ θ when A and Rh as factors (predictors) are used , it has been found that the best fitting is: ∆∆22θθ = 4.535 – 0.462*A +1.156*A2 – 0.327*Rh +0.08249*Rh2–8.678.10-5Rh3+3.24.10-7Rh4 –0.0132*A*Rh R = 0.814 ; F = 2.79 (4) The model (4) is strong non-linear, but it contains the main features, which has been described above, i.e. general anticorrelation between the temperature “residuals” by 2 one side and A, Rh and A*Rh by the other one. The factor variance Sf 2 . from the total variance St is about 64% 3.2. The cycles in the sunspot arrea assymetry index A series (AD 1821- 1994) One of the steps of the present study is an analysis for an existence of cycles in the A- index data series. The main field of interests there is the possible existence of statistically significant cycles by subcenturial and nearcenturial duration and their comparison with the corresponding temperature “residual” data series spectra. On fig.3 the T-R correlogram (see Paper I) of the A-index series (AD 1821-1994) is shown. The time step ∆∆T is 0.5 years. The starting period is T0 = 2 years, the upper limit is at T=402 years. Fig. 3 The T-R spectra of the north-south sunspot arrea assymetry index A (AD 1821-1994)(smoothed 11-year values) As it is shown there are two main cycles by durations of ~42.5 ( 4 Schwabe- Wolf’s cycles) and 114.5 years. The second one is much more powerful. There is also an adjacent secondary peak at T=145 years, thus it could to assume that there is a doublette structure which mean duration is about 125-130 years. However it should be taken this result with some reserve by the fact that the duration of the founded cycle is comparable with the length of the all data series (174 years), so by this one it could be determined rather as a “trend – hypercycle” as just a “cycle”. On other hand the very high correlation coefficient R=0.72 at T=114.5 years show that most probable the quasi-120 year cycle is an important feature of the north- south sunspot arrea assymetry dynamics. As it has been already shown in Paper II a hypercycle by duration of approximately 117 years has been found for the long-lived solar filaments (Duchlev, 2001). Conseqently, this is a variation of the solar corona dynamics, most probably not only for the filaments, but for other events (like CMEs) too. May be the north-south assymetry of the sunspot activity centers lead to long term tendencies of north-south anisotropy in the corona and the coronal events. This could affect corresponding GCR- flux fluctuations in the Earth atmosphere and on this base – fluctuations in the “cosmogenic” isothopes production rates , including the 10Be too. It is interesting in this course to mark also that a well expressed quasi –130-140 year cycle is visible also in the radiocarbon tree rings data series (Dergachev , 1994). Obviously the same phenomena should be affect the reaching the planets high energetic proton and electron fluxes (the solar cosmic rays, SCR) and this also is taken effect over their atmospheres and climate. The presence of this powerful quasi-120 year oscilation in the A-index data series gives by our opinion a principle explanation why a cycle with the same duration exist both in 10Be and climate data series, including also in the dendrochronological data (Komitov et al.,2003). In fig.3 a weak , but statistically significant oscilation at T=62 year is shown. But as it has been already noted (see Paper I) there are weak oscilations by similar durations in the sunspot data series too (Komitov and Kaftan, 2003). On other hand there is a strong quasi- 60 year climatic cycle, which extremums are in antiphases with the corresponding ones of MLA series (see above), as well as with these of 10Be (Paper II) . These facts are a strong indicator for the solar corpuscular origin of the climatic 60- year cycle. However by our opinion it seems almost unpossible that the weak quasi-60 year cycles in A could be a source of the so powerful oscilations in the aurora, 10Be or climate. Obviously a much stronger solar source of this phenomena should be exist. 3.3 The Group sunspot numbers (Rh) and their subcenturial (quasu-60 yr ) oscilations (AD 1610-1979) The multiple regressional analysis , which has been described above is clearly point out that the MLA annual number Aur_N have not any specific role as a factor proxy for the ∆∆22ΘΘ during the 19th century and ther participation as a factor is totally “captured” by Rh , A and the interactive term of the both last ones. On other hand a weak , but statistically significant 62 year cycle in the dynamics of the assymetry index A also exist . This is why the problem about the origin of the solar source of the subcenturial quasi 60 yr cycle has been rested open. It has been very unexpetable when by using of the T-R periodogram procedure a strong quasi 67 year cycle for the epoch AD 1821-1900 has been detected in the smoothed 11 year Rh series (fig.4). As it is shown the correspondig corelllation coefficient R value is > 0.6 . In the T-R correlogram for the epoch AD 1913-1979 a totally dominant 59 yr cycle (R > 0.9; fig.5) is shown! No other traces of cycles in the subcenturial or quasicenturial range are visible in these both spectra. It is necessary to the studied time intervals are comparable by their length with the so detected subcenturial oscilations. By this one it is more correctly to determine the last ones as “trend – hypercycles”, because they are occur only once in the both studied epochs. These results are very intriguing, because usually for the overall sunspot activity where as a proxy the International Wolf’s number is used, the cycles by longer duration (78, 88-90 , ~100 years) are discussed (Gleissberg, 1944; Vitinski et al., 1976 , Bonev, 1997). On other hand there have brief comments for a 65 yr cycle in the Schove’s series (Schove, 1955) and weak variations in the subcenturial range (50-70 years) in the instrumental sunspot series Rh and Ri (Komitov and Kaftan,2003,2004) . This is why it has been decided to search how stable is this 60 yr cycle in the Group sunspot number series , as well as is there some evolution of the sunspot cycles in the centurial and subcenturial ranges during the period 1610-1979. For this aim a two –dimensional T-R periodogram “moving window” procedure (MWTRPP, see Paper I) has been provided. The moving window lemght has been choiced to be 60 year and the parameters of the single T-R corellogram procedure are T0= 2 years, time step ∆∆T=0.5 years and Tmax= 152 years correspondly. The evolution of the ratio R/SR (the ratio of R to its error) for the cycles in the range [T0,Tmax] is shown due to the map on fig.6. Fig.4. The T-R spectra of the Group sunspot number (11-yr smoothing annual values; AD 1820-1900). Two powerful cycles by duration of 29 and 67 years are shown. Fig.5. The T-R spectra of the Group sunspot number (11-yr smoothing annual values; AD 1913-1979). The most interesting feature on fig.6 is related to the generally high level of presence of the subcenturial 50-70 yr oscilations in the Rh series before AD 1850. In the second half of 19th and the beginning of 20th centuries this type of cyclity has been sharply falling and this seems be much better expressed near to the zurich sunspot cycle No 13, i.e. before the Gleissberg-Gnevishev’s solar centurial minimum. It is also clear visible a tendency for restoring of the quasi subcenturial cyclity after AD 1910. However as it is also shown there is slightly earlier also a tendency for longer quasi centurial 120 year trend -hypercycle . The last one is seems caused by the centurial minimum (AD 1898-1923). It is absent in the most recent “moving window” spectra after AD 1925- 1930 when the data from this minimum are not already included , i.e the last right columns of the map. A good confirmation for this is the strong peak at T= 59 years in fig.5 . In contrary the T-R spectra on fig.4 is related to an transition period from epoch with good expressed 60 yr cycle before AD 1850 and such one when this cycle is totally absent. Thus it could be say that the 50-70 year osilations are much more typical for the Group sunspot number data series during the last 400 years as every of the non-stable quasi-centurial ones. It is even valid for the Maunder minimum epoch too, where the traces of subcenturial oscilations are weak , but even so , more visible as in the second half of 19th century. It should also conclude that the abruption of the 60 year cycle between AD 1850 and 1910 is the main cause by which it is not so visible in the general Rh series. It is necessary note that the second half of the 19th century is also a period of strong decreasing of the MLA activity, negative values of the A-index and a fast climate warming (fig.1 and 2). Fig.6. The T-R spectra evolution of the 11 year smoothed GSN data in the range of periods T between 2 and 152 years. By the horisontal line of T=62 years the typical duration of the subcenturial cycle is signed. The most white arreas on the map to statistically non-significant values of -0. 11 < R/SR < 0 are corresponded. There is no significant change of the results if the last 16 years up to AD 1995 of the Rh- series are included in the MWTRPP . An comparison with the Zurich series (the index Ri) could be very interesting, but it will be an object of a separate study. 4. Discussion A few important conclusions should be derived by the presented in these three papers results and their analysis. There are also a number of questions , which remain still open or are new ones. Most of all, the general influence of the Sun over climate is much more rich and complicate as it is presented on the base only of the TSI variations or even if the additional solar depended mechanism of galactic cosmic rays modulation over the aerosols and clouds production is taken into account. A third and very important Sun- climate forcing channel is related to the powerful erruptive events, which could generate high energetic protons (E ≥≥ 100 MeV). They are able to penetrate very deep in the Earth stratosphere and troposphere and even in the cases when E> 1 GeV to reach also the surface. A solar particles with such energies are labeled very often as solar cosmic rays (SCR). The solar origin of MLA events is out of doubt. As it has been shown in Paper II for these phenomena during 18th –19th century a powerful ~60 yr cycle is typical . The same one is valid for the Northern hemisphere temperature residuals and the “Greenland” 10Be series too. The analysis in Paper II as well as in this paper III are shown that these oscilation in the last two series are not by terrestrial, but rather by solar origin too. Most probably the 10Be and ∆∆22ΘΘ 60 yr cycles are connected to the same solar phenomena as MLA. And there is a question – where on the Sun these phenomena are occurred? Are they coronal events like the coronal mass ejections (CME), or other components of the solar erruptive activity are also taken a significant participation in this channel of the solar forcing over the climate? It need to say that there are not enough clear evidences on this stage about the dominant role of the coronal events (and especially of CMEs) for the generation of quasi- subcenturial climate as well as for the overall dynamics of the temperature “residual “ series ∆∆22ΘΘ and ∆∆22θθ at all. It is rather visible by the results of the multiple regressional analysis both of the Northern hemisphere ((∆∆22ΘΘ ) ) and the “oceanic” ∆∆22θθ residual temperature series that the negative values (cooling) correspond in generally to an increasing of the overall sunspot activity index Rh . It is an indicator that the cooling effect is related to the increasing effect of the erruptive events , which are close connected to the sunspot groups. This concern strongly especially the Northern ((formulas (2) and (3)) , while for the “oceanic” residual hemisphere residuals ∆∆22ΘΘ series the relationship with Rh it is much more complicate (4). The high amplitude and statistical significance of the quasi 60 year oscilations in the GSN data series is an additional evidence that both the climate and 10Be cycles with the same duration should be related to the overall erruptive activity. It also indirectly shown that the Rh index is a very good proxy of the solar corpuscular activity. Unfortunately there are no any updates of the GSN data series after AD 1995. It could say on the base of the results in &3.3 that only by the relative short period at the end of 19th century disturbs for the much better expression of the quasi-60 yr cycle in the GSN data series during the last 300 years after the Maunder minimum. By our opinion the origin of the quasi –60 year cycle in the MLA events is now also clear. It is caused by the corresponding variations in the number of the active centers, which good proxy is the Rh index. No any additional specific sources of solar activity for explanation of the 60 yr cycle of the auroral activity are strongly needed. So it is clear that our preliminary hypothesis for such sources (see Paper I and II) is not without fall necessary. The obtained importance of 60 yr cycle in GSN data series lead on the top the question , what sunspot index is better – the international Wolf’s number Ri or Rh? The problem for a comparison of Ri and Rh as a proxies of the sunspot activity in different aspects of the solar-climatic relationships will be an object of our future paper. By the way there it will be only note that by opinion of many authors the Rh- index relative to the “classic” Ri is much better proxy for the aims of solar-terrestrial physiscs at all , because it described much better the solar erruptive activity. May be the most interesting result from this analysis is the obtained strong relationship between the Northern hemisphere temperature residuals and north –south assymetry index A of the sunspot arreas. The last one is generally in not very strong relationship with the sunspot activity index Rh. By other words the A – index is a second and relatively independent factor , which play very important role both for the ∆∆22ΘΘ and ∆∆22θθ dynamics. The linear correlation coefficent r is equal to - 0.336 for the all period since AD 1821 (the beginning of the Pulkovo archive data) to 1979. It point out for a generall statistically significant reversed relationship. The coefficient r is very high by module ( r = - 0.72) during the 19th century. For the recent part of ∆∆22ΘΘ data series AD 1913-1979 the relationship has been sharp fadded (r = - 0.26), but remain statistically significant over 95%. The “interactive” terms (Rh*A) in formulas 2 and 4 shown that for the climatic effects over climate is very important no only the total level of the solar eruptive activity , but also where on the Sun the active regions are placed. The established fact that the relative increasing of the Sun Northern hemisphere activity lead to a climate cooling effects is in good agreement with the Loginov’s suggestions about 35 years ago (Loginov, 1973), namely that the flare activity centers on the north of the Sun equator are more geoeffective as the southern ones. As it is shown from the relationships (2-4) there are two “pure” north-south assymetry index terms of linear ‘A’ and quadratic ‘A2’ types in the all three formulas. This is an indirect evidence that a climate forcing by solar activity processes, which are not very close connected to sunspot active centers should be also significant. Most probably these terms are connected directly to the coronal phenomena. It could be related no only to the CME events, but also to other coronal phenomena, including the more long time lived structures there and the large scale coronal structure too. It need to include in this course a possibility that the variations of the A-index could no only affect of the geoefficiency of the erruptive coronal processes like CMEs. They could also generate long time anisotropy variations of the interplanetary matter density and by this one lead to corrsponding time and space variations of the falling in the Earth atmosphere GCR –flux. As a result there should be effects over “cosmogenic” isothopes production rates, aerosols and clouds genetartion etc. The index of the north south sunspots assymetry A is an object of studying namely in the field of the solar –terrestrial and especially solar-climatic relationships (Georgieva, 2002). It is inetresting to point out for an interesting group of relationships between the north-south sunspot assymetry, the Earth diurnal rotation rate and the atmospheric circulation changes (Georgieva, 2002). The physical hypothesis for these influences is connected to the “solar wind –geomagnetic field – Earth dynamo” relationship. If this could be true and relationship between the sunspot flare proceses and sunspot assymetry from one side and the Earth tektonic activity from other should be exist. The possible relationship of the solar and Earth voulcanic activity is briefly discussed by the author in his recent overview (Komitov, 2008). The multi-factor statistical models (2,3 and 4) pointed out that between 60 to 80% of the temeperature residuals ∆∆22ΘΘ and ∆∆22θθ are explained by the solar factors. The general anticorrelation both with the assymetry index A and Rh , the very important participation of non-linear terms in these formulas as well as the “interactive” terms of type A*Rh shown that these factors are predominantly connected to the solar corpuscular activity. It need to remember that the general relationships “sunspot activity - temperature changes” ( formulas (4) and(5), Paper I ) express mainly the overall effect of the large time scale electromagnetic flux variations over the climate (the TSI changes). There could be also taken into account the possibility for a significant “hidden” participation in these models of the Forbush-effect and the GCR influence over the climate: The climate cooling effect of the GCR-flux increasing during the sunspot and TSI minimums should be made the overall cooling effect even more expressive. This is why the above mentioned models are explained very well the coincidience between the such significant phenomena like the solar supercenturial solar Maunder minimum (1640-1720) and the deepest phase of the last “Litle ice epoch”, the next one ,also supercenturial Dalton minimum (1795-1835) and the temporal cooling during this time , as well as the Modern supercenturial solar minimum (1933- 1996/2000) and the modern warm climate epoch. However the climate effects of the solar erruptive activity are outside of these models. They are not taken into account in the most apropriated explanations for the climate changes in the modern epoch and especially during the last 35-40 years since the middle of 1970st. The results and the analysis in our study are clearly pointed out that it is a very serious gap in the present domimamt climate changes theories. Only if the solar erruptive activity in combination with its spacial distribution over the Sun surface (the A-index) is taken into account it could be explain successfully the climate dynamics during the relative short “mirror epochs” ,when the sunspot activity and the temperature changes are in anticorrelation (see fig.1 in Paper I). This concern also the last 30-35 years. It will be demonstrated below how the specific combination of the solar erruptive activity and north –south sunspot area assymetry is the most probable factor for the fast warming both in the second half of 19th and in the end of 20th centuries and no additional cause (human activity) is needed for explanation there. The solar activity and the climate changes during the 19th century Undoubtly the most important solar activity event at the beginning of 19th century is the supercenturial Dalton minimum (AD 1795-1830/35) . The essential climate cooling during the this time is related to both the corresponding TSI decreasing and the increasing of the penetrating in Earth atmosphere GCR-flux. In generally this picture of the solar-climatic relationships during the Dalton minimum is correct. However there are some important details. It is clearly shown on fig,1 that there is an initial period at the beginning of the Dalton minimum , when an increasing of the temperature residuals ∆∆22ΘΘ in the Northern hemisphere is observed. The critical moment is near to AD 1805, i.e. near to or slightly after the maximum of zurich sunspot cycle No 5. So there is an delaying of about 10 years after the beginning of the Dalton minimum , when a sharp decreasing in the “residual” temperature series has been started. Consequently it should say that there is an additional climate cooling effect over the long –term downward TSI tendency. The sunspot activity during this time is low, and it is reflected to the low level of MLA activity (Paper II). Even so in the course of our results and analysis it should be assume that this cooling is caused by the solar erruptive factor and its increasing geoefficiency during this time. A very probable situation could be – relative rare, but strong erruptions , originated predominantly on the Northern hemisphere of the Sun. As it has been shown (fig.2) the Pulkovo archive data series is starting at AD 1821 with a positive values of the A-index . The tendency of A is positive during the next two decades and this correspond to even more deep cooling up to AD 1840, when the Dalton minimum has been already ended. However the Rh increasing after AD 1830 is predominantly in the Northern hemisphere and this has been supported the cooling tendency else certain time. A qualitative extrapolation of the smoothed A-index data in the past before AD 1820 shown that most probably it has been positive since 1805-1810, i.e well correspond to the observed decreasing of the temperature residuals since AD 1805. Near to the maximum of zurich cycle No 9 a serious change in the long term solar activity tendencies has been occurred: 1. The “smoothed” assymetry A-index sign has been changed for long time (up to 1910-1912) from positive to negative. 2. There is a clear visible long term decreasing of the 11- year smoothed Group sunspot number data from zurich cycles No 9 to 14. The local peak near to AD 1870 is not affected seriously for this long time tendency (Paper I, fig.1) . Simultaneosly with these two events a long time ∆∆22ΘΘ upward tendency from AD 1840 to 1910 has been occurred. There is only a short temporary stopping between AD 1870-1880. This dynamic of the climate changes is in very good agreement to the presented in this study results, their analysis and the following from them conclusions. The climate changes during the 20th century and the modern “global warming” After the centurial Gleissberg- Gnevishev’s minimum (AD 1898-1923) the solar activity hast been very fast increased. After AD 1934 when sunspot cycle No 17 has been started and especially after 1940 the solar activity has been remained for a few decades on extremally high levels up to the end of sunspot zurich cycle No 22 in AD 1995/96 This epoch (the Modern supercenturial solar maximum ) is characterized no only with the higher for the last 1000 years Sun’s luuminocity , but also with an extremally high erruptive activity which centers has been located predominantly in the Northern hemisphere (fig2.). As it has been shown there, the “smoothed” positive sign of A has been remained up to the middle of 1970st. During the last 20 years after AD 1975 of the Pulkovo archive data series the sunspot assymetry index is predominatly negative. However the sunspot activity and the TSI index has been remained at high long-term level during this time. As a result of all these circumstances , during the middle and the end of 20th century is a supercenturial climate warming tendency maximum. It is connected to the supercenturial maximum of TSI on first place and the supercenturial GCR- flux minimum (less GCR –flux, less aerosols and clouds production ). However, on other hand the very high erruptive activity levels by itteraction with the positive A –index lead to a secondary cooling effect. It is much better expressed between AD 1940 and 1975 when the flare activuty has been very fast increased. Since the 1970st in coincidence with the transition of A-index from positive to negative the secondary cooling has been stopped and the general climate tendency has been changed to warming. The high levels of TSI during this time as well as the downward tendencies in the erruptive activity after the maximum of cycle No 21 (Komitov, 2008; Paper II) are additional factors to forcing of the warming during the last two decades of 20th century. The zurich cycle No 24. What could be expected? There are many indications that with the end of solar cycle No 23 in 2008 a new supercenturial solar Dalton-type minimum is already started (Komitov and Bonev, 2001; Komitov and Kaftan 2003,2004; Shatten and Tobiska, 2003; Ogurtsov, 2005). This is why the next sunspot cycle No 24 should be essentialy weaker by magnitude as the previous few ones. The nearmaximal annual sunspot number Ri in AD 2012 or 2013 is expected to be less than 100 , but there are also predictions for values near to or less than 50 (Cliverd et al., 2006). For an extremal low level of sunspot activity near to AD 2020 –2025 is poinetd out by Hathaway(2006) on the base of “Great Conveiyor Belt” model estimatons . Consequently, a climatic significant decreasing of TSI during the cycle No 24 relative to No 23 as well as an increasing of GCR –flux should be expected as a general tendency. By this one and according the relationships between sunspot activity and climate a cooling effect in range of 0.2-0.3K if an decreasing of the 11 year smoothing Rh values from 65-67 (at the beginning of 1990th ) to 25 between 2010 and 2020 is assumed (see Paper I ,formula (4) and (5)). It need also to remember that there are not actual data for the Group sunspot number index Rh after AD 1995 and any extrapolation should be made only on the base of some similarity with Ri. However there should be added the effect of solar erruptions. Except a prediction for Rh it need to have also a prediction for the index of sunspots area assymetry index A too . If during the sunspot cycle No 24 the active centers are predominantly in the Northern hemisphere, which is by our opinion the most probable scenario (see also “Great Conveyor Belt” model results (Hathaway, 2006)), an additional cooling in order of 0.3 K for the Northern hemisphere over the above signed 0.25-0.3 K should be predicted. Thus a climate conditions could be returned back to almost the same ones as during the Dalton minimum. There are historical evidences that the relationship between the auroral activity and the weather conditions in North Europa has been noticed still by the Vikings in the Middle Ages (Corbin, 2008). On other hand there are many studies during the 1950th - 1970st ,which are focused over the effects of the strong solar corpuscular erruptions for the climate. A good overviews of this studies is given by Rubashov (1963), Vitinskii et al.(1976) and Herman and Goldberg (1978). Our study shown that the understanding of this relationship is of most higher importance for the correct understanding of the climate changes in the modern epoch at all. REFERENCES Beer J., Blinov B., Bonani G, Finkel R.C., Hofmann H., Lelmann B., Oeschger H., Sigg A., Schwander J., Staffelbath T., Stauffer B., Suter M and Glfli W.,1990, Nature v347, 16‘-166. Beer, J., Tobias, S. and Weiss, N., 1998, Solar Phys. 181(1), 237–249 Bonev, B., 1997, Bulg. Geophys. J., , vol. 23, nos. 3/4, pp. 43–47. Clilverd M., Clarke E, Ulich T.,Rishbeth H, Jarvis1 M, 2006, , SPACE WEATHER, VOL. 4, S09005, doi:10.1029/2005SW000207, 2006 de Jager C. and Usoskin I., 2006, J.Atm. Sol-Terr.Phys., v.68, pp2053-2060 Dickinson, R. E., 1975, Bull. Am. Meterol. Soc., v56, pp 1240–1248 Duchlev P., 2001, Sol. Phys., v199, Issue 1, p. 211-215 (2001) Frolich C. et al. ,1997, in The First Results from SOHO,edd. by B.Flrck and Z.Sestka, Solar Phys., 170 Georgieva K., 2002, Phys.Chem.Earth, v.27,Issues 6-8, 433-440 Hathaway D., 2006,NASA Headline News,(05.10.2006) Herman J.R. and Goldberg R., 1978, Sun, Weather and Climate, NASA Sci an Technology Inf. Branch Komitov B. and Kaftan V.,2003, International Journal of Geomagnetism and Aeronomy,v.43,No5,2003,pp 553-561 Komitov B. P and Kaftan V. I. , 2004, in Proceedings IAUS 223 'Multi-Wavelength Investigations of the Solar Activity', eds. A. V. Stepanov, E. E. Benevolenskaya & A. G. Kosovichev, Cambridge University Press, pp.115-116 Komitov B., 2008, The solar activity forcing over climate in the past and presence: Relations to Bulgaria, Alphamarket Press Ltd., St.Zagora, 2008 (in Bulgarian) Krivova N.A., Balmaceda L. and Solanki S., 2007, Astron. Astrophys., 467, 335-346 Krivsky L. and Pejml K.,1988, Solar activity, aurorae and climate in Central Europe in the last 1000 years,Bul.Astron. Inst. Chechosl. Acad. Sci.,No75 Lean, J., J. Beer, and Bradley R., 1995, Geophys. Res. Lett., v.22, No. 23, pp 3195-3198, December 1, 1995. Lean, J. , 2000., Geophysical Research Letters, Vol. 27, No. 16, pp. 2425-2428, Aug. 15 Lean, J. , 2004, Solar Irradiance Reconstruction. IGBP PAGES/World Data Center for Paleoclimatology Data Contribution Series # 2004-035. NOAA/NGDC Paleoclimatology Program, Boulder CO, USA. Lockwood M. and Frolich C.,2007, Proc. R. Soc. A., doi:10.1098/rspa.2007.1880, /Published online/ Loginov,F.N., 1973, Harakter solnechno-zemnih svyazei, Gidrometeoizdat, Leningrad (in Russian) 4-07   $430.3    42703   ,98034    ,70Y 3     Nature 433, 13. Ogurtsov M., 2005, , Sol. Phys., v231, pp 167-176 Pap J.M.,Turmon M.,Floyd L., Frolich C amd Wehrli Ch, 2002, Adv.Space Res., 29,No12, 1923-1932 Parker, D.E., Folland, C.K. and Jackson, M., 1995 , CLIMATIC CHANGE, Vol. 31, pp. 559-600 (1995). Rubashev B., 1963, Problemi solnechnoi aktivnosti, Nauka, Moskwa (in Russian) Solanki S., 2002, Harold Jeffrey’s Lecture Solanki, S. K., Schüssler, M., & Fligge, M. 2002, Aastron. Astrophys, 383, 706 Schove, D. J. 1955, J. Geophys. Res., 60, 127 Svensmark, H., and E. Friis-Christensen, 1997, J. Atmos. Sol. Terr. Phys., 59, 1225– 1232 Tinsley B.A., 2000, Space Sci Rev., 00,1-28 Thompson D, 1997, Proc. Nat. Acad. Sci. USA, v94, pp8370-8377 Usoskin I., Schlussler M., Solanki S and Mursula K.,2005, Proc. 13th Cool Stars Workshop, Hamburg 5-9 July, (ESA SP-560, Jan 2005,F Favata, G.Hussein and B. Battrick eds.) Vitinskii Y,I, Ohl A. and Sazonov A., 1976, Solnce I atmosfera Zemli, Gidrometeoizdat, Leningrad (in Russian) Yu F., 2002, Geophys. Res. Lett., v107, No A7
1912.01637
2
1912
2019-12-08T17:11:32
Collisional disruption of highly porous targets in the strength regime: Effects of mixture
[ "astro-ph.EP" ]
Highly porous small bodies are thought to have been ubiquitous in the early solar system. Therefore, it is essential to understand the collision process of highly porous objects when considering the collisional evolution of primitive small bodies in the solar system. To date, impact disruption experiments have been conducted using high-porosity targets made of ice, pumice, and glass, and numerical simulations of impact fracture of porous bodies have also been conducted. However, a variety of internal structures of high-porosity bodies are possible. Therefore,laboratory experiments and numerical simulations in the wide parameter space are necessary. In this study, high-porosity targets of sintered hollow glass beads and targets made by mixing perlite with hollow beads were used in a collision disruption experiment to investigate the effects of the mixture on collisional destruction of high-porosity bodies. Among the targets prepared under the same sintering conditions, it was found that the targets with more impurities tend to have lower compressive strength and lower resistance against impact disruption. Further, destruction of the mixture targets required more impact energy density than would have been expected from compressive strength. It is likely that the perlite grains in the target matrix inhibit crack growth through the glass framework. The mass fraction of the largest fragment collapsed to a single function of a scaling parameter of energy density in the strength regime ({\Pi}_s) when assuming ratios of tensile strength to compressive strength based on a relationship obtained for ice-silicate mixtures.
astro-ph.EP
astro-ph
Collisional disruption of highly porous targets in the strength regime: Effects of mixture Yuichi Murakami1, Akiko M. Nakamura1, Koki Yokoyama1, Yusuke Seto1, and Sunao Hasegawa2 1: Department of Planetology, Kobe University, Japan 2: Institute of Space and Astronautical Science, Japan Aerospace Exploration Agency, Japan Contact address: 1-1 Rokkodai-cho, Nada-ku, 657-8501 TEL : +81-78-803-5740 FAX: +81-78-803-5791 e-mail: [email protected] Abstract Highly porous small bodies are thought to have been ubiquitous in the early solar system. Therefore, it is essential to understand the collision process of highly porous objects when considering the collisional evolution of primitive small bodies in the solar system. To date, impact disruption experiments have been conducted using high-porosity targets made of ice, pumice, gypsum, and glass, and numerical simulations of impact fracture of porous bodies have also been conducted. However, a variety of internal structures of high-porosity bodies are possible. Therefore, laboratory experiments and numerical simulations in the wide parameter space are necessary. In this study, high-porosity targets of sintered hollow glass beads and targets made by mixing perlite with hollow beads were used in a collision disruption experiment to investigate the effects of the mixture on collisional destruction of high-porosity bodies. Among the targets prepared under the same sintering conditions, it was found that the targets with more impurities tend to have lower compressive strength and lower resistance against impact disruption. Further, destruction of the mixture targets required more impact energy density than would have been expected from compressive strength. It is likely that the perlite grains in the target matrix inhibit crack growth through the glass framework. The mass fraction of the largest fragment collapsed to a single function of a scaling parameter of energy density in the strength regime (𝛱(cid:3046)) when assuming ratios of mixtures. However, the dependence on 𝛱(cid:3046) is much larger than that shown for porous tensile strength to compressive strength based on a relationship obtained for ice-silicate targets with different internal microstructures from the targets in this study. The depth of the deep cavity specific to the high-porosity target was well represented by a dimensionless parameter using the compressive strength of both the pure glass and mixture targets. The empirical relationship of cavity depth was shown to hold for various targets used in previous studies irrespective of the internal microstructure of the targets. 1. Introduction Small bodies in the solar system have porous structures. Asteroids with diameters less than tens of kilometers are known to have considerable porosity. This is especially the case for C-class asteroids, which typically have bulk density lower than 2 g cm -- 3 and porosity of 40% or more (Consolmagno et al., 2008). The bulk density and porosity of the Tagish Lake meteorite were reportedly 1.64 g cm -- 3 and ~40%, respectively (Hildebrand, et al., 2006). Small Saturnian satellites have densities from 0.34 -- 0.69 g cm -- 3, which is indicative of a porous internal structure (Porco et al., 2007). The bulk density of comet 67P/Churyumov-Gerasimenko (67P/C-G) is 0.47 g cm -- 3, which corresponds to a porosity of 70 -- 80% if the solid density of the ice-dust mixture is assumed to be 1.5 -- 2.0 g cm -- 3 (Sierks et al., 2015). Comets 19P/Borrelly, 81P/ Wild2, and 9P/Tempel1 have porosities in a similar range (Consolmagno et al., 2008). The uniaxial compressive strength of the near-surface layer of comet 67P/C-G was estimated at > 2 MPa, which suggests that the near-surface ice -- dust layer was sintered to have a structure with porosity of 30 -- 65 % (Spohn et al., 2015). Sintering is a microscopic process of mass transfer that physically connects adjacent small particles by forming necks between the particles below the melting temperature. The sintering process depends on temperature. For example, the growth of necks between water ice particles of radius 0.1 m takes 0.15 years at 100 K (Sirono, 1999). Sintering of water ice particles can progress within the age of the solar system at the radiation equilibrium temperature in the orbit of Jupiter (Gundlach et al., 2018). The temperature condition for sintering of silicate dust particles might also be satisfied for small objects orbiting very close to the Sun: such temperature is ~1000°C for the lunar simulant basalt (Allen et al., 1992) and 1.5 m amorphous SiO2 particles (Poppe 2003), which is only reached by the bodies within 0.09 AU from the Sun, however, the number of bodies with the perihelion distance smaller than 0.1 AU is very limited in the current era (Granvik et al., 2016). As sintering proceeds, necks grow larger, forming stronger connections between particles (Poppe 2003; Machii and Nakamura, 2011), and the overall structure attains lower bulk porosity. Impact experiments have been conducted to understand the impact cratering and disruption processes of porous small bodies of various densities and strengths using porous targets consisting of particles physically connected to each other. Sintered glass bead targets with porosity of 5 -- 60 % were impacted by a projectile at impact velocity about 5 km s -- 1 (Love et al., 1993). It was shown that specific energy required to destroy targets greatly depends on porosity. Collisional disruption experiments of sintered glass bead targets with various compressive strength but fixed porosity ~40% at impact velocity between 32 m s -- 1 and 2.2 km s -- 1 showed that the specific energy for disruption increases with the compressive strength of targets (Setoh et al., 2010). Collisional disruption experiments of porous ice and snowball targets with porosity of 30 -- 45% and 39 -- 54% at impact velocity between 73 and 308 m s -- 1 and 90 and 155 m s -- 1, respectively, showed that the specific energy required for disruption depends on the internal structure and is higher for porous targets than solid ice targets (Ryan et al., 1999; Giblin et al., 2004). Collisional experiments of sintered ice targets including even higher porosity up to 70% at the impact velocity from 2.4 to 489 m s -- 1 confirmed the tendency; more porous target requires more energy density to be catastrophically disrupted (Shimaki and Arakawa, 2012a). On the other hand, it was shown that disruption threshold decreased with increase of porosity in the case of sintered ice-silicate mixture targets with the mass ratio of ice to silicate, 0.5 and with porosity of 0 -- 39% at the impact velocities of 150 to 670 m s -- 1 (Arakawa et al., 2002; Arakawa and Tomizuka, 2004). The reason of the opposite tendency to pure ice target was shown to be due to significant decrease of strength with increasing porosity of the mixture targets. The number of known low-density, highly porous, small bodies remains limited; however, it is expected that small bodies in the early solar system had high porosity. A theoretical study reported that icy dust grains in the environment of protoplanetary disks accumulate to form planetesimals with radius of 10 km and density of 0.1 g cm -- 3 (Kataoka et al., 2013). A laboratory measurement of the pressure -- density relationship of fine particles was extrapolated to estimate the internal porosity structure of granular small bodies. A spherical body with radius of 10 km consisting of particles having the same compression property as 1.7 m silica beads was shown to have bulk porosity of 82% (Omura and Nakamura, 2018). Numerical simulations of collisional processes have been tested by reproducing laboratory results of non-porous and porous targets for the purpose of effective modeling of collisional processes of small bodies (e.g., Benz and Asphaug, 1994; Jutzi et al., 2009; de Niem, et al., 2018). However, porous small bodies can have a variety of internal structures (Nakamura et al., 2009), so laboratory experiments and numerical simulations covering wide parameter space are useful for understanding the collision process and studying the collisional evolution of small bodies. To examine the outcome of collisional disruption of targets with high porosity, targets formed of hollow-glass beads with bulk porosity of 87% and 94% have been impacted at velocities between 1.8 and 7 km s -- 1 (Okamoto et al., 2015). The specific energy required for disruption was as large as several kJ/kg, which is much larger than that required for basalt targets. Hollow glass beads are useful for making high porosity structures similar to those of perlite and pumice (Nakamura et al., 2009). The void space of perlite and pumice is formed by evaporation and degassing of volatiles. A similar structure is seen as a feature of scoriaceous cosmic spherules that have lost their volatile components due to heating upon entry into the Earth's atmosphere (Rudraswami et al., 2018). On the other hand, dust grains of sub-micron ~ micron size in protoplanetary disks are considered to form coherent aggregates of very porous structure consisting of filamentous skeleton surrounding void spaces (e.g., Poppe, 2003; Wada et al., 2009). The filamentous framework of the dust aggregate may be simulated by a very thin mesh wall surrounding the void. The hollow glass bead used in a previous study (Okamoto et al., 2015) has thin shell with a thickness of 0.95 m. Although the shell of the hollow bead is not a mesh structure, a sintered hollow glass bead target may mimic the mechanical and impact response of primitive highly porous bodies formed by the accumulation of porous dust aggregates, especially thermally evolved icy bodies with enhanced bonding between icy dust grains of ~1 m size. In addition, it has been shown that the structure of silicate dust found in comets varies from solid to very fluffy, including the build-up of sub- structures (Güttler et al., 2019). Accordingly, we conducted impact experiments using the high-porosity targets of hollow glass beads similar to those of the previous study (Okamoto et al., 2015) and targets of porous silicate mixtures to investigate the effects of high porosity and mixture. In this study, experiments were performed only at high impact velocities (> 2 km/s) and compared with previous studies of sintered hollow glass bead targets conducted at similar velocites, however, the low-velocity parameter space needs to be further explored. 2. Experiment Table 1 summarizes the heating conditions and physical properties of the four different types of targets used in this study. Two of the targets were pure glass bead targets and two were mixtures of glass beads and perlite grains. The pure glass bead targets with bulk porosities of 86% (HGB87) and 94% (HGB94) and corresponding bulk densities of 0.36 and 0.15 g cm -- 3, respectively, were prepared in a manner similar to that used in previous experiments (e.g., Okamoto et al., 2013). The beads were hollow soda -- lime -- borosilicate glass microspheres (3M Co.) with an average diameter and shell thickness of 55 m and 0.95 m, respectively. The isostatic crush strength of the bead is 5.2 MPa (3M catalogue). This value of strength corresponds to the static uniaxial compressive strength of less-porous (~10% porosity) pure ice (Arakawa and Tomizuka, 2004; Hiraoka et al., 2008). The beads were heated in molds from room temperature to a peak temperature, 800 and 650 °C, over 30 min, respectively. The mold was cup-shaped. We covered the top of the mold with a lid to ensure relatively uniform heating. The peak temperatures were retained for 6 h. Then, the heater was switched off and the targets were cooled to room temperature over 9 h. The HGB87 target had a roughly cylindrical shape with diameter of 58 mm and height of 56 mm. The HGB94 target had roughly the shape of a truncated cone with top and bottom surfaces with diameters of 78 and 65 mm, respectively. The height of this target was 78 mm. The bulk porosity 𝜙 of the target is defined as follows: 𝜙(cid:3404)1(cid:3398)(cid:3096)(cid:3096)(cid:3116), where 𝜌 is the bulk density of the target and 𝜌(cid:2868) is the true density of the constituent (1) material, i.e., 2.5 g cm -- 3 for the glass. The mixture targets had a mixing ratio of hollow glass beads and perlite grains (<0.6 mm, typically) of 2:1 (mix2:1) and 1:1 (mix1:1) by weight, respectively. Figures 1a and 1b present scanning electron microscopy (SEM) images of the glass beads and perlite grains. The glass beads and perlite grains were placed in a box, shaken for 1 min, poured into molds, and then placed into an oven to be heated to 800 °C. The melting point of perlite is higher than 1093 °C (International Chemical Safety Cards; ICSCs), whereas the softening point of the hollow glass bead material is 600 °C (3M catalogue). The mix2:1 and mix1:1 targets had roughly cylindrical shapes with diameters and heights of 66 mm and 63 mm, and 68 mm and 64 mm, respectively. The bulk porosity of the mixture targets was defined as follows: 𝜙(cid:3404)1(cid:3398) (cid:3096) (cid:3096)(cid:3116)(cid:3117)(cid:3033)(cid:3117)(cid:2878)(cid:3096)(cid:3116)(cid:3118)(cid:3033)(cid:3118), where 𝜌(cid:2868)(cid:2869) and 𝜌(cid:2868)(cid:2870) are the grain densities of the constituent particles and 𝑓(cid:2869) and 𝑓(cid:2870) (2) are the mass fraction of each component, respectively. The perlite grains were themselves porous. We used the true density value of the perlite material (ICSCs), which is 2.2 g cm -- 3. The three targets of HGB87, mix2:1, and mix1:1 were formed with the same heating conditions but with different mixing fraction of perlite grains. The porosity of the target increased with mixing fraction. Figures 1c and 1d present the internal structure of a target and its appearance, respectively. Although the structure is quite inhomogeneous microscopically, the walls of the hollow glass spheres are connected, forming a larger macroscopically continuous structure. The observed thicknesses of necks between glass beads were similar for HGB87, mix2:1, and mix1:1, which is consistent with the fact that they were formed by heating to the same peak temperature. The necks of HGB94 were less thick than the other targets. The microscopic structure of empty void spaces surrounded by thin walls is to some extent similar to the structure of pumice and perlite (Nakamura et al., 2009), although in the case of pumice and perlite the cells are mostly open and the scale of void spaces is larger than in the hollow glass-bead material used in this study. In order to examine vertical variation of the compressive strength of the target, we sliced the target in the horizontal direction to obtain approximately 2 cm thick discs, from which we cut out three cylinders from the central part of the discs with diameter of 1 cm and length of 2 cm and measured the static uniaxial compressive strength. At some level of compression force, the sample started to locally break and collapse. We defined this force level as the threshold and calculated the threshold force per unit area as the compressive strength. The longitudinal and shear wave velocities were determined by measuring the time required to propagate longitudinal and shear waves through samples of three different thicknesses using piezoelectric sensors. Impact disruption experiments were conducted at impact velocities of 2.3 to 7.0 km s -- 1 using a two-stage light-gas gun at the Institute of Space and Astronautical Science (ISAS). In high-velocity impact experiments with porous targets, projectile density has been a challenge: lower-density materials generally have smaller strength and we cannot accelerate a projectile with a density comparable to the target density without destroying the projectile. In this study, we successfully launched a wood projectile with a density lower than that of the nylon projectile typically used in previous studies. Nylon spheres with diameter 3.2 mm and wood (Mempisang) columns with diameter of 3 mm and length of ~2.5 mm were accelerated using a split-type nylon sabot (Kawai et al., 2010). Figure 1e presents an image of the wood projectile. The fiber of the wood was parallel to the axis of the column. The compressive strength of the wood column was about 51 MPa, sufficiently strong to be accelerated to more than 5 km s -- 1. The strength of the column was also measured by pressing from the side of the column, which corresponds to the Brazilian disc test (Diyuan and Louis, 2013). The tensile strength of the column was thus determined to be 2.4 MPa. Table 2 summarizes the physical properties of the projectiles. The targets were either hung by thread or placed on a stand made of plastic or paper. The bottom and backside wall of the experimental chamber were covered with plastic cushioning. Nominally the target was set with its symmetry axis aligned with the projectile trajectory so that it would be impacted on the top surface. However, in the case of the target being hung by thread, the attitude of the target was random and the projectile could strike obliquely onto the top, side, or even bottom surface of the target. To check the point of impact, we used two high-speed video cameras to acquire imagery from directions orthogonal to the projectile trajectory. Shimazu HPV-X and HPV-1 cameras were used in earlier shots, and the Shimazu HPV-X and a Kirana-05M were used in later shots. The cameras were nominally operated at (2 -- 5)(cid:3400)10(cid:2873) fps with exposure durations 0.2 -- 1 and 0.5 s, respectively. The motion of the largest fragments was monitored using another high-speed video camera, the Photron SA1.1, operated at 5400 fps with an exposure duration of 50-185 s. Additionally, because the projectile was accelerated using a sabot, the trajectory could not be adjusted accurately toward the center of mass of the target. The impact parameter 𝑏 was defined as: where ℎ is the distance between the target's center of mass and the projectile trajectory and 𝑅 is the radius of the sphere having the same volume as the target (equivalent sphere 𝑏(cid:3404)(cid:3035)(cid:3019), (3) radius). We defined an apparent impact angle θ, which is the angle between the projectile trajectory and the impacted surface of the target looking from the vertical direction (i.e., looking down from the top), for example, θ=90° indicates an apparent normal impact on the surface. Because the projectile could strike on the side surface of the target higher or lower than the target's center of mass, the apparent impact angle is not an exact one. Table 3 summarizes the impact conditions. 3. Result 3.1 Static strength and wave velocities Figure 2 presents the measurement results of compressive strength,𝑌(cid:3030) . The horizontal axis of Fig. 2a indicates the depth of the cylindrical core specimen normalized by the height of the target. The strength of the parts near the top and the bottom surfaces was generally lower than that of the central part. The variation within a target was most prominent in the most-porous target HGB94, probably because it had the lowest thermal conductivity: the temperature in the central part of the target would have declined more slowly than the temperature near the surface of the target. The variation was within a factor of 2, and we hereafter refer to the average value of strength. Figs. 2b compares the compressive strength and bulk density of the targets in this study with those in previous studies. The compressive strength of HGB87 (1.7 ± 0.4 MPa) agreed to within 1  with a target prepared using the same peak temperature and duration in a previous study, which had a compressive strength of 1.4 ± 0.4 MPa (fluffy87 in Okamoto et al., 2013). The compressive strength of HGB94 (0.09 ± 0.03 MPa) was lower than one with the same porosity, which had a compressive strength of 0.47 ± 0.13 MPa (fluffy94 in Okamoto et al., 2013), because of the lower peak temperature used during the heating process in this study. The compressive strength decreased with the increase of mixing fraction of perlite grains, i.e., HGB87 > mix2:1 > mix1:1. The wave velocity had a tendency similar to that of the compressive strength: the higher the perlite fraction, the slower the wave velocity. In other words, the targets formed using the same peak temperature (HGB87, mix2:1, and mix1:1) had higher compressive strengths and wave velocities with increases in the fraction of glass and bulk density. Table 1 lists the measurement results. 3.2 𝑄(cid:3046)∗ Figure 3 presents an example of a set of fragments collected after impact. In most of the shots, a bulb-shaped cavity formed below the impact point and the target separated into a few to tens of larger pieces. The largest fragments moved with a velocity lower than several m s -- 1 in a direction almost parallel to the projectile trajectory. The center-of- mass velocity of the system was between 0.9 and 2.6 m s -- 1 and the velocities of the largest fragments were within the same order of magnitude. Most of them were not broken when struck and landed on the cushioning. Figure 4 presents the results of the largest fragment mass fraction of the target, (4) 𝑄(cid:3404) (cid:3040)(cid:3023)(cid:3118) (cid:2870)(cid:4666)(cid:3014)(cid:2878)(cid:3040)(cid:4667), 𝑀(cid:2869)/𝑀, where 𝑀(cid:2869) and 𝑀 are the mass of the largest fragment and the initial mass of the target, respectively, versus the specific energy of impact 𝑄 which is defined as where 𝑉 and 𝑚 are the impact velocity and mass of projectile, respectively. The results our data here include data for various impact angle 𝜃 and impact parameter 𝑏 values, whereas the previous data were obtained only for 𝜃(cid:3404)90°. No information about b for of HGB87 agree with the previous results of fluffy87 (Okamoto et al., 2015), although the previous data is available. The projectiles used in the previous study were nylon and titanium spheres: therefore, the data collectively show no clear difference between projectiles of different density ranging from the value of 0.74 g cm -- 3 of wood to the value of 4.5 g cm -- 3 of titanium. Such insensitivity to projectile material was also observed in a previous study of rock disruption (Katsura et al., 2014). As expected from the lower static strength of HGB94 than fluffy94, the HGB94 target was easier to destroy than the fluffy94 target. The data of the mix2:1 and mix1:1 targets were more scattered than those of the HGB87 and HGB94 targets, although the all target types had similar degree of scatter in the static strength (24-32 %) as shown in Table 1, i.e., variations in the bulk properties of the targets by the manufacturing process were similar. The data of mixture targets are plotted between those of HGB87 and HGB94, which is in agreement with the fact that the mix2:1 and mix1:1 targets had static strength between those of the HGB87 and HGB94 targets. The scattering of the data of the mixture targets is probably due to the inhomogeneity in internal structure of these targets. Crack growth in homogeneous targets is reproducible, whereas crack growth in mixture targets has lower reproducibility. Crack growth in a hollow glass bead structure stops when the crack crosses the boundary between the hollow glass bead and a perlite grain. Least-square fits of the following equations were applied to the data: (cid:3014)(cid:3117)(cid:3014)(cid:3404)a(cid:2869)𝑄(cid:2879)(cid:3029)(cid:3117), 𝑄(cid:3404)a(cid:2870)(cid:4666)𝑀(cid:2869)𝑀(cid:4667)(cid:2879)(cid:3029)(cid:3118) (5) (6) using the parameter of Eq. 6 required to make the largest fragment have half the mass of and Table 4 lists the fitted parameters. We derived the shattering specific energy 𝑄(cid:3046)∗ the initial target, i.e., (cid:3014)(cid:3117)(cid:3014)(cid:3404)0.5 . The ambiguity of the estimated 𝑄(cid:3046)∗ was calculated according to error propagation. We confirmed that values of 𝑄(cid:3046)∗ derived from Eq. 5 agree to the data irrespective of 𝜃 and b. Figure 4 presents the fitting curves. To summarize, compressive strength and specific energy 𝑄(cid:3046)∗ were HGB87 > mix2:1 > mix1:1 and HGB87 > mix2:1 (cid:3410) mix1:1, respectively, as shown in Table 1 and Table 4, i.e., more with those from Eq. 6 within the ambiguity shown in Table 4. The fittings were applied the sintering condition of HGB87, mix2:1, and mix1:1 was the same. The resultant fraction of impurity, less static strength and impact resistance. 3.3 Cavity The results revealed that the cavity below the impact point in a highly porous target becomes bulb-shaped when the projectile is broken (Okamoto et al., 2013; 2015). Figure 5 presents an example of a cavity. We measured the depth of cavity 𝑑(cid:3029) and the depth at the maximum diameter of cavity 𝑑(cid:3030), which would correspond to the depth of the center of the spherical source of shock-wave divergence (the center of the isobaric core) (Mizutani et al., 1990), using images of the fragments recovered after shots. Because the symmetry axis of the bulb is not necessarily in the fracture surface, the values shown in Table 5 are reference values. The range of the ratios of these measurements, (cid:3031)(cid:3277)(cid:3031)(cid:3278), was 1.7 ± 0.3. The depth of cavity in this study was as deep as roughly half the target, i.e., (cid:3031)(cid:3277)(cid:3013)(cid:3295)~0.5, where 𝐿(cid:3047) denotes the height of the target. 4. Discussion 4.1 Effects of impact parameter and oblique incidence As mentioned in section 3.2, the impact parameter and impact angle had minor effects on the largest-fragment mass fraction. This tendency is different from what was found in previous impact disruption experiments conducted using spherical rock targets (Fujiwara and Tsukamoto 1980; Nakamura 1993). Note that in the case of a spherical target, the impact parameter has a one-to-one correspondence to the impact angle. Figure 6 presents 𝑄/𝑄(cid:3032)(cid:3033)(cid:3033)(cid:3030)(cid:3047) versus impact parameter and impact angle, where 𝑄(cid:3032)(cid:3033)(cid:3033)(cid:3032)(cid:3030)(cid:3047) is the calculated value required to obtain the measured largest-fragment mass 𝑀(cid:2869) of each shot basalt spheres, an effect of impact geometry appears at 𝑏(cid:3408)~0.5 and 𝜃 (cid:3407)60° : using Eq. (5). Figure 6 also presents previous results for basalt spheres. In the case of however, the results of this study show no clear effect. The degree of destruction was probably insensitive to the impact geometry because of the target shape and the high porosity of the target. The mass of fragments excavated by an impact is known to be dependent on the curvature of the target surface (Fujiwara et al., 1993; Fujiwara et al., 2014; Suzuki et al., 2018) and the distance between the point that corresponding to the center of the isobaric core of pressure propagation and the target free surface (Suzuki et al., 2018). In the present case, the roughly cylindrical shape of the targets and their highly porous internal structure collectively made the distance from the center of the isobaric core to the free surface insensitive to collision geometry, and thus the outcome was insensitive to collision geometry. Additional experiments with spherical targets can reveal the geometric effect of the collision more clearly. 4.2 Cavity depth The cavity depth depends on the target type: deeper depths for more fragile, lower- density targets. A previous study proposed an empirical relationship between depth at maximum cavity width, 𝑑(cid:3030) in this study, and 𝐿(cid:2868)≡ (cid:2869)(cid:2870)(cid:3080) as: Where 𝑑 is the projectile diameter and (cid:2869)(cid:2870)(cid:3080) is the characteristic distance through which (7) (8) (cid:3031)(cid:3278)(cid:3031)(cid:3404)10(cid:2868).(cid:2874)(cid:2868)(cid:4666)(cid:3013)(cid:3116)(cid:3031)(cid:4667)(cid:2868).(cid:2872)(cid:2874), 𝑚(cid:3031)(cid:3023)(cid:3031)(cid:3047)(cid:3404)(cid:3398)(cid:2869)(cid:2870)𝐶(cid:3031)𝜌𝑆𝑉(cid:2870), the projectile is decelerated in the porous target: 𝛼≡(cid:3004)(cid:3279)(cid:3096)(cid:3020)(cid:2870)(cid:3040) , where 𝐶(cid:3031) and 𝑆 are the drag coefficient and projectile cross-sectional area, respectively (9) (Okamoto et al., 2015). When the projectile deforms or breaks, the cross-sectional area and the mass of the projectile change from the initial values. Using X-ray transmission images, the following empirical relationship was derived: (cid:3031)(cid:3013)(cid:3116)(cid:3404)(cid:2871)(cid:2870)𝐶(cid:3031)(cid:4672)(cid:3096)(cid:3083)(cid:4673)(cid:3404)(cid:2871)(cid:2870)(cid:3400)10(cid:2879)(cid:2868).(cid:2868)(cid:2871)(cid:2877)(cid:4666)(cid:3096)(cid:3023)(cid:3118)(cid:3026)(cid:3291)(cid:3295)(cid:4667)(cid:2868).(cid:2870)(cid:2870)(cid:4666)(cid:3096)(cid:3083)(cid:4667), where 𝛿 is the projectile density and 𝑌(cid:3043)(cid:3047) is the tensile strength of projectile. Figure 7 presents (cid:3031)(cid:3278)(cid:3031) versus (cid:3013)(cid:3116)(cid:3031) of this study. We adopted the 'compressive' strength of the wood (10) projectile instead of the 'tensile' strength taking the direction of acceleration into account. The results are not as consistent with Eq. (7) as were seen for fluffy 87 and fluffy 94 in the previous study. It reveals a larger dependence on the property of the target, i.e., density or compressive strength. To take the target strength 𝑌 into account, we adopted the non-dimensional forms according to pi-group scaling (Holsapple, 1993): 𝜋(cid:3031)(cid:3029)/𝜋(cid:2872)(cid:3081)(cid:3279)(cid:3277)∝𝜋(cid:2871)(cid:3080)(cid:3279)(cid:3277), 𝜋(cid:3031)(cid:3030)/𝜋(cid:2872)(cid:3081)(cid:3279)(cid:3278)∝𝜋(cid:2871)(cid:3080)(cid:3279)(cid:3278) where 𝛼(cid:3031)(cid:3029), 𝛼(cid:3031)(cid:3030), 𝛽(cid:3031)(cid:3029), and 𝛽(cid:3031)(cid:3030) are fitting parameters. The non-dimensional parameters are defined as, 𝜋(cid:3031)(cid:3029)(cid:3404)(cid:4666)(cid:3096)(cid:3040)(cid:4667)(cid:2869)/(cid:2871)𝑑(cid:3029), 𝜋(cid:3031)(cid:3030)(cid:3404)(cid:4666)(cid:3096)(cid:3040)(cid:4667)(cid:2869)/(cid:2871)𝑑(cid:3030), 𝜋(cid:2871)(cid:3404) (cid:3026)(cid:3083)(cid:3023)(cid:3118), 𝜋(cid:2872)(cid:3404)(cid:3096)(cid:3083). (12) (11) We assumed 𝛽(cid:3031)(cid:3029)(cid:3404)𝛽(cid:3031)(cid:3030)(cid:3404)0.01 according to the value of 𝛽(cid:3031)(cid:3029) obtained in a previous compressive strength 𝑌(cid:3030) as 𝑌 in this study, whereas the previous study adopted the tensile strength as 𝑌 (Suzuki et al., 2012). Figure 8a presents the results including those study of a sedimentary rock with porosity of ~17% (Suzuki et al., 2012). We adopted the of hollow glass bead targets from a previous study (Okamoto and Nakamura, 2017). Although the previous ones are data of cratering shots, i.e., the targets were not broken as a whole, and the targets in this study were destroyed, the depth data appear similar. Contrary to the result of (cid:3014)(cid:3117)(cid:3014) , the scattering of 𝜋(cid:3031)(cid:3029) of the mixture targets is not apparently 𝑌(cid:3030). We obtained empirical relationships by least squares fittings to the results of this study: larger than those of pure targets. This probably occurred because the cavity depth depends on the compressibility of the target, which is characterized by the compressive strength (cid:3095)(cid:3279)(cid:3277)(cid:3095)(cid:3120)(cid:3116).(cid:3116)(cid:3117)(cid:3404)(cid:4666)1.057(cid:3399)0.030(cid:4667)𝜋(cid:2871)(cid:2879)(cid:2868).(cid:2869)(cid:2876)(cid:2869)(cid:3399)(cid:2868).(cid:2868)(cid:2871)(cid:2873), (cid:3095)(cid:3279)(cid:3278)(cid:3095)(cid:3120)(cid:3116).(cid:3116)(cid:3117)(cid:3404)(cid:4666)0.158(cid:3399)0.006(cid:4667)𝜋(cid:2871)(cid:2879)(cid:2868).(cid:2871)(cid:2868)(cid:2871)(cid:3399)(cid:2868).(cid:2868)(cid:2872)(cid:2873). and (13) (14) The density ratios of projectile to target in this study were between 2 and 7. Ordinary chondrites (3.2 ~ 3.4 g cm-3) (Consolmagno et al., 2008) and comets (for example, 0.47 g cm-3 of 67P/C-G) reach a density ratio of about 7, but collisions between objects with a density ratio of about unity also occur in interplanetary space. To anticipate the outcome of a collision of equally dense objects, we examined whether the empirical relationship shown by Eq. (13) could be extrapolated to the case of collision between projectile and target with smaller density ratio. Figure 8b compares the results of this study with the results of previous studies with projectile to target density ratio less than 2. Pumice and gypsum targets (0.59 and 1.1 g cm -- 3) were impacted by a nylon projectile (1.1 g cm -- 3) with velocities between 3.6 and 7.2 km s -- 1 and a velocity fixed at 3.5 km s -- 1, respectively (Okamoto and Nakamura, 2017; Kadono et al., 2018). Data of a cratering experiment of snow target (36% porosity) conducted with velocities between 27 and 145 m s -- 1 are also shown (Arakawa and Yasui, 2011). We assumed that the dynamically determined strength of snow is comparable to that of the statically measured compressive strength of other materials. All of the previous data of the variety of materials clustered near the empirical relationship of Eq. (13). Shallow craters in the case of snow projectiles rather than ice projectiles may be explained by the lower strength of the snow projectiles. Figure 8b presents the result of crater depth on a basalt target for which the projectile to target density ratio was 0.9 (Table A-1) and an empirical relationship obtained for sedimentary rock (Suzuki et al., 2012), too. The result for basalt falls roughly on the line of Eq. (13). The discrepancy in the empirical relationship for the sedimentary rock target compared with the other results is probably due to the use of the target's tensile strength instead of the target's compressive strength. In summary, the empirical relationship (Eq. 13) holds within factor of 2 not only for the present results but also for the experimental results with projectile to target density ratio between 0.9 -- 2, irrespective of the internal microstructure of the targets: the sintered hollow glass bead targets of this study and pumice target have the porous structure consisting of void spaces surrounded by walls, whereas gypsum and snow targets are coherent aggregates (Nakamura et al., 2009). 4.3 Effect of mixture on degree of destruction As described in section 3.2, Fig. 4 shows that 𝑄(cid:3046)∗ depends on target strength. Yasui and Arakawa (2011) found that the results of impact disruption experiments of gypsum and gypsum -- glass bead mixed targets, which simulated the parent bodies of ordinary chondrites, were well organized using a non-dimensional parameter, that is, the non- dimensional impact stress, 𝑃(cid:3010)(cid:3047), which is defined as (Mizutani et al., 1990): where 𝑃(cid:2868), 𝑉(cid:3043)(cid:3045)(cid:3042)(cid:3037)(cid:3032)(cid:3030)(cid:3047)(cid:3036)(cid:3039)(cid:3032), 𝑉(cid:3047)(cid:3028)(cid:3045)(cid:3034)(cid:3032)(cid:3047), and 𝑌(cid:3047) are initial shock pressure, volume of projectile, 𝑃(cid:3010)(cid:3047)(cid:3404)(cid:3023)(cid:3291)(cid:3293)(cid:3290)(cid:3285)(cid:3280)(cid:3278)(cid:3295)(cid:3284)(cid:3287)(cid:3280) (cid:3023)(cid:3295)(cid:3276)(cid:3293)(cid:3282)(cid:3280)(cid:3295) (cid:3017)(cid:3116)(cid:3026)(cid:3295), (15) volume of target, and tensile strength of target material, respectively. In this study, the tensile strength of target was not measured, so we tried a modified version of impact stress, (16) 𝑃(cid:3010)(cid:3030)(cid:3404)(cid:3023)(cid:3291)(cid:3293)(cid:3290)(cid:3285)(cid:3280)(cid:3278)(cid:3295)(cid:3284)(cid:3287)(cid:3280) (cid:3023)(cid:3295)(cid:3276)(cid:3293)(cid:3282)(cid:3280)(cid:3295) (cid:3017)(cid:3116)(cid:3026)(cid:3278). 𝑃(cid:3010)(cid:3030) (Takagi, et al., 1984; Mizutani et al., 1990) which is defined as: We calculated 𝑃(cid:2868) using Hugoniot parameters and the planar approximation (Melosh, 1989). The shock wave velocity 𝑈(cid:3036) and 𝑃(cid:2868), where i= p and t are projectile and target, respectively, are expressed as: 𝑃(cid:2868)(cid:3404)𝜌𝑈(cid:3047)𝑢(cid:3043)(cid:3047) 𝑜𝑟 𝛿𝑈(cid:3043)𝑢(cid:3043)(cid:3043), 𝑈(cid:3036)(cid:3404)𝐶(cid:3036)(cid:3397)𝑠(cid:3036)𝑢(cid:3043)(cid:3036), where 𝐶(cid:3036) , 𝑠(cid:3036) , 𝑢(cid:3043)(cid:3036) are bulk sound velocity, material constant, and particle velocity, respectively. Table 6 summarizes the Hugoniot parameters 𝐶 and 𝑠 we used. Figure 9a range than in the case in which specific energy 𝑄 was used (Fig. 4). In contrast, the shows that the results of the pure glass targets (HGB87 and 94) collapse into a narrower (17) (18) results of the mixture targets show higher resistance to impact destruction than that expected based on the compressive strength of the target material. The definitions of 𝑃(cid:3010)(cid:3047) and 𝑃(cid:3010)(cid:3030) as expressed in Eqs. (15) and (16), respectively, compared to the target length, i.e., (cid:3031)(cid:3278)(cid:3013)(cid:3295)≪1 . However, in the present study, (cid:3031)(cid:3278)(cid:3013)(cid:3295)~ (cid:2869)(cid:2869).(cid:2875)(cid:3400) 0.5~0.3 and may not be negligible, so we took the effect of the depth of the center of the are approximations when the depth of the center of the isobaric core is negligible isobaric core into account. Figure 9b presents the largest-fragment mass fraction and the modified version of impact stress 𝑃(cid:3010)(cid:3030)′ for the shots in which the projectile impacted the top or bottom surface of the target with impact angle 𝜃(cid:3408)80°; We used values of 𝑑(cid:3030) calculated using the empirical relationship expressed in Eq. (14). 𝑃(cid:3010)(cid:3030)(cid:4593) (cid:3404)(cid:3023)(cid:3291)(cid:3293)(cid:3290)(cid:3285)(cid:3280)(cid:3278)(cid:3295)(cid:3284)(cid:3287)(cid:3280) (cid:4666)(cid:3013)(cid:3295)(cid:2879)(cid:3031)(cid:3278)(cid:4667)(cid:3119) (cid:3017)(cid:3116)(cid:3026)(cid:3278). (19) Because of the low compressive strength of the target, the depth of the center of the isobaric core of the mixture target is large: thus, the difference between the pure glass target and the mixture target does not become small by this modification. The apparent high resistance of the mixture targets shown in Figs. 9a and 9b may be interpreted as the fracture growth in the network of glass beads in the mixture target being blocked by perlite grains or by pores introduced by the mixing with perlite grains, making it difficult for fracture to propagate. Not only the structural discontinuities introduced by perlite grains, but also the relative weakness of perlite grains over hollow glass beads could have acted as an obstacle for the propagation of fracture. In other words, more energy may have dissipated by the pulverization of perlite grains. Hiraoka et al. (2008) found that the ratio of tensile strength to compressive strength of an ice -- silicate mixture with porosity of ~10% increased with increase of silicate fraction, 𝑓 (100(cid:3400) 𝑓%) as: (cid:3026)(cid:3295)(cid:3026)(cid:3278)(cid:3404)(cid:2868).(cid:2874)(cid:2876)(cid:3032)(cid:3118).(cid:3122)(cid:3281) (cid:2874).(cid:2873)(cid:3032)(cid:3116).(cid:3120)(cid:3281)(cid:3404)0.10𝑒(cid:2870).(cid:2870)(cid:3033). (20) The tendency of the silicate mixture to increase the ratio of tensile strength to compressive strength of the mixture target is consistent with the results shown in Figs. 9a and 9b. A non-dimensional scaling parameter 𝛱(cid:3046)_(cid:3042)(cid:3045)(cid:3036)(cid:3034)(cid:3036)(cid:3041)(cid:3028)(cid:3039), which was introduced based on the coupling parameter concept (Holsapple and Housen, 1986; Housen and Holsapple, 1990), was shown to be effective for compiling laboratory results of targets with various porosities conducted at a wide range of velocities (Okamoto et al., 2015). By using this parameter, the largest-fragment mass fraction data of gypsum and gypsum -- glass bead mixed targets were fitted by a single relationship (Yasui and Arakawa, 2011). The dependence of the strength of target material on size scale and strain rate, respectively, expression of 𝛱(cid:3046)_(cid:3042)(cid:3045)(cid:3036)(cid:3034)(cid:3036)(cid:3041)(cid:3028)(cid:3039) is (Eqs. 40 of Housen and Holsapple, 1990): 𝛱(cid:3046)_(cid:3042)(cid:3045)(cid:3036)(cid:3034)(cid:3036)(cid:3041)(cid:3028)(cid:3039)(cid:3404)𝑄𝑠(cid:4666)(cid:3096)(cid:3026)(cid:4667)(cid:3119)(cid:3339)(cid:3118)(cid:3127)(cid:3338)𝑅(cid:2879)(cid:2871)(cid:3091)(cid:4666)(cid:3090)(cid:2878)(cid:3099)(cid:4667)/(cid:4666)(cid:3099)(cid:2879)(cid:2870)(cid:4667)𝑉(cid:2871)(cid:3091)(cid:2879)(cid:2870)(cid:4666)(cid:3096)(cid:3083)(cid:4667)(cid:2869)(cid:2879)(cid:2871)(cid:3092), where 𝑅 and 𝑌 denote target radius and strength, respectively, 𝜆 and 𝜏 describe and 𝜇 and 𝜈 are constants used in the coupling parameter. When the strength of the target material does not depend on size scale but only on the strain rate (λ(cid:3404)0(cid:4667) and if we assume 𝜈(cid:3404)(cid:2869)(cid:2871) then we obtain: The value of 𝜇 is theoretically limited between 1/3 and 2/3. Similarly to the case of the previous study (Okamoto et al., 2015), we assumed a linear relationship between 𝜇 and 𝛱(cid:3046)_(cid:3042)(cid:3045)(cid:3036)(cid:3034)(cid:3036)(cid:3041)(cid:3028)(cid:3039)(cid:4666)𝜆(cid:3404)0,𝜈(cid:3404)(cid:2869)(cid:2871)(cid:4667)(cid:3404)𝑅(cid:2879)(cid:2871)(cid:3091)(cid:3099)/(cid:4666)(cid:3099)(cid:2879)(cid:2870)(cid:4667)𝑄𝑠(cid:4666)(cid:3096)(cid:3026)(cid:4667)(cid:3119)(cid:3339)(cid:3118)𝑉(cid:2871)(cid:3091)(cid:2879)(cid:2870). (21) (22) mixture target materials and 0.42 -- 0.43 for pumice (Table A-2), which are similar to those porosity, 𝜇(cid:3404)(cid:2870)(cid:2871)(cid:3398)(cid:3109)(cid:2871) . Thus, the value of 𝜇 becomes 0.35 -- 0.38 for glass beads and of porous targets such as perlite -- sand mixture with 60% porosity (𝜇(cid:3404)0.35), sand -- fly ash mixture with 45% porosity (𝜇(cid:3404)0.4) (Housen and Holsapple, 2011), and gypsum with 65 -- 69% porosity (𝜇(cid:3404)0.398) (Nakamura et al., 2015). Figure 10a shows (cid:3014)(cid:3117)(cid:3014) versus 𝑄(cid:3046)(cid:4666)𝜌𝑌𝑐(cid:4667)3𝜇2𝑉3𝜇(cid:3398)2 for glass bead and mixture targets as well as pumice data. We substituted 𝑌(cid:3030) for 𝑌 of Eq. (22). The data of the glass beads (HGB87 and HGB94) and the pumice are fitted by a single relationship: (cid:3014)(cid:3117)(cid:3014)(cid:3404)(cid:4666)2.5(cid:3399)0.7(cid:4667)(cid:3400)10(cid:2879)(cid:2872)(cid:4668)𝑄𝑠(cid:4666)(cid:3096)(cid:3026)(cid:3278)(cid:4667)(cid:3119)(cid:3339)(cid:3118)𝑉(cid:2871)(cid:3091)(cid:2879)(cid:2870)(cid:4669)(cid:2879)(cid:2869).(cid:2873)(cid:2870)(cid:3399)(cid:2868).(cid:2868)(cid:2875). (23) The agreement between the hollow glass bead and pumice results may be due to the similar internal structure of the two substances. The discrepancy of the data of the mixture targets from Eq. (23) is probably due to the effect of the perlite grains on crack growth in the mixture targets. We used Eq. (20) to anticipate the tensile strength of the targets used in this study, although the ratio of tensile strength to compressive strength was obtained for different compositions (ice and serpentinite powders in the previous study) and for different porosity (~10%). For pumice, we used the measured value (1 MPa, Jutzi et al., 2009). In Fig. 10b, all the data are collapsed into a single relationship: (cid:3014)(cid:3117)(cid:3014)(cid:3404)(cid:4666)1.16(cid:3399)0.04(cid:4667)(cid:3400)10(cid:2879)(cid:2870)𝛱(cid:3046)(cid:2879)(cid:2869).(cid:2874)(cid:2876)(cid:3399)(cid:2868).(cid:2868)(cid:2875). (24) where, 𝛱(cid:3046)(cid:3404)𝑄𝑠(cid:4666)(cid:3096)(cid:3026)(cid:3295)(cid:4667)(cid:3119)(cid:3339)(cid:3118)𝑉(cid:2871)(cid:3091)(cid:2879)(cid:2870). (25) The fitted line is shown in the figure. The power index of 𝛱(cid:3046) in Eq. (24) is (cid:3398)1.7 and mixed targets ((cid:3398)0.96) (Yasui and Arakawa, 2011). The cavity depth is expressed by the shows much larger dependence than that found for the gypsum and gypsum -- glass bead power law relationship of the dimensionless forms (Eq. 13) irrespective of the internal microstructure of the target, while the largest fragment mass fraction has different dependence on 𝛱(cid:3046),(cid:3047) depending on the internal microstructure of the target. 5. Summary There were likely a variety of porous small bodies in the solar system, especially in the early phase of the solar system. Laboratory and numerical studies on the collisional processes of porous bodies of various internal structure increase our understanding of these collisional process. In this study, we conducted impact disruption experiments of highly porous targets made of pure glass and glass -- silicate mixture. The targets have a porosity in the range of 86 ~ 94%, and may be regarded as a mimic of primitive highly porous bodies consisting of dust aggregates with enhanced bonding due to thermal processing. The impact velocities ranged from 2.3 -- 7.1 km s−1 in this study, however, to understand the collisional evolution of small bodies since the early solar system, it is also necessary to study the low-velocity parameter space. To reduce the density ratio of projectile to target, we used nylon and wood projectiles. The results revealed that the depth of the deep cavity specific to the high-porosity target was well represented by the dimensionless parameter of PI-group scaling for cratering in the strength regime using the compressive strength of both the pure glass targets and the mixture targets. It was shown that the same relationship holds widely for other targets with different internal microstructures. The largest fragment mass fraction was insensitive to collision geometry, i.e., impact parameter and impact angle, probably due to the deep center of the isobaric core. The targets with more impurities tend to have lower compressive strength and lower resistance against impact disruption. Further, the mixture targets required more impact energy density than would have been expected from the static compressive strength. This was probably because the impurity inhibited the growth of cracks in the framework structure made of glass. The largest-fragment mass fraction of the pure glass targets and the mixture targets, as well as the results of pumice targets, collapsed to a single function of a non-dimensional scaling parameter of energy density in the strength regime (𝛱(cid:3046)) (e.g., Housen and Holsapple, 1990) by assuming ratios of tensile strength to compressive strength based on a relationship obtained for an ice -- silicate mixture (Hiraoka et al., 2008). The largest fragment mass fraction obtained in this study showed a greater dependence on 𝛱(cid:3046) than previously obtained for porous targets with different internal microstructures. The tensile strength of a mixture depends on various factors, including porosity, fraction of impurity, composition of impurity, temperature, and grain size (e.g., Arakawa and Tomizuka, 2004; Hiraoka et al., 2008; Litwin et al., 2012). Although the results of this study suggested that the degree of collisional destruction of the targets would depend on tensile strength rather than compressive strength, the tensile strength of the targets were not directly measured in this study. Further studies on the tensile strength of high- porosity structures of primitive small bodies in the solar system are needed, including whether the tensile strength of the targets used in this study has the impurity fraction dependence assumed here. Acknowledgement We are grateful to the constructive comments of two anonymous reviewers. The series of experiments was supported by the Hypervelocity Impact Facility (the Space Plasma Laboratory) at ISAS, raJAXA, Japan. Y. M. is grateful to M. Hyodo for allowing him to acquire SEM images of the hollow-glass spheres and perlite grains. Appendix Table A-1 lists the conditions of the impact experiment with the basalt (Kinosaki, Japan) target. The target was a rectangular parallelepiped with sides ranging from 12.1 cm to18.5 cm. The density of the target was 2.7 g cm -- 3. We did not measure the strength of the target material and we assumed the compressive strength of 220 MPa according to a previous study (Takagi et al., 1984). Crater depth was measured using a laser profiler and is shown in the table (e.g., Suzuki et al., 2012). Table A-2 lists the conditions of the impact experiment with the pumice target. A portion of the result was presented in previous studies (Nakamura et al., 2009; Jutzi et al., 2009): however, a few new experiments were conducted after these studies. References Allen, C. C., Hines, J. A., Altemir, D. A., Mckay, D. S., 1992. Sintering of lunar simulant basalt. Proc. Lunar and Planet. Sci. Conf. XXIII, pp. 19-20. Arakawa, M., Leliwa-Kopystynski, J., Maeno, N., 2002. Impact experiments on porous icy-silicate cylindrical blocks and the implication for disruption and accumulation of small icy bodies. Icarus 158, 516-531. Arakawa, M., Tomizuka, D., 2004. Ice silicate fractionation among icy bodies due to the difference of impact strength between ice and ice silicate mixture. Icarus 170, 193 -- 201. Arakawa, M., Yasui, M., 2011. Impact crater formed on sintered snow surface simulating porous icy bodies. Icarus 216, 1-9. Benz, W., Asphaug. E., 1994. Impact simulations with fracture. I - Method and tests. Icarus 107, 98-116. Chronological Scientific Tables, 2019. National Astronomical Observatory. Maruzen Co. Ltd. (in Japanese). Consolmagno, G., Britt, D., Macke, R., 2008. The significance of meteorite density and porosity. Chemie der Erde - Geochemistry 68, 1 -- 29. De Niem, D., Kührt, E., Hviid, S., Davidsson, B., 2018. Low velocity collisions of porous planetesimals in the early solar system. Icarus 301, 196-218. Diyuan L., Louis N. Y. W., 2013. The Brazilian Disc Test for Rock Mechanics Applications: Review and New Insights. Rock Mech Rock Eng 46: 269-287. Fujii, Y., Nakamura, A. M., 2009. Compaction and fragmentation of porous gypsum targets from low-velocity impacts. Icarus 201, 795-801. Fujiwara, A., Tsukamoto, A., 1980. Experimental study on the velocity of fragments in collisional breakup. Icarus 44, 142-153. Fujiwara, A., Kadono, T., Nakamura, A.M., 1993. Cratering experiments into curved surfaces and their implication for craters on small satellites. Icarus 105, 345 -- 350. Fujiwara, A., Onose, N., Setoh, M., Nakamura, A.M., Hiraoka, K., Hasegawa, S., Okudaira, K., 2014. Experimental study of impact-cratering damage on brittle cylindrical column model as a fundamental component of space architecture. Adv. Space Res. 54, 1479 -- 1486. Giblin, I., Davis, D. R., Ryan, E.V., 2004. On the collisional disruption of porous icy targets simulating Kuiper belt objects. Icarus 171, 487 -- 505. Granvik, M., Morbidelli, A., Jedicke, R., Bolin, B., Bottke, W. F., Beshore, E., Vokrouhlický, D., Delbò, M., Michel, P., 2016. Super-catastrophic disruption of asteroids at small perihelion distances. Nature 530, 303-306. Güttler, C., Mannel, T., Rotundi, A., Merouane, S., Fulle, M., Bockelée-Morvan, D., Lasue, J., Levasseur-Regourd, A. C., Blum, J., Naletto, G., Sierks, H., Hilchenbach, M., Tubiana, C., Capaccioni, F., Paquette, J. A., Flandes, A., Moreno, F., Agarwal1, J., Bodewits, D., Bertini, I., Tozzi, G. P., Hornung, K., Langevin, Y., Krüger, H., Longobardo, A., Della Corte, V., Tóth, I., Filacchione, G., Ivanovski, S. L., Mottola, S., Rinaldi, G., 2019. Synthesis of the morphological description of cometary dust at comet 67P/Churyumov-Gerasimenko. Astron. Astrophys., in press. Gundlach, B., Ratte, J., Blum, J., Oesert, J., Gorb, S. N., 2018. Sintering and sublimation of micrometre-sized water-ice particles: the formation of surface crusts on icy Solar System bodies. MNRAS 479, 5272 -- 5287. Hildebrand, A. R., McCausland, P. J. A., Brown, P. G., Longstaffe, F. J., Russell, S. D. J., Tagliaferri, E., Wacker, J. F., Mazur, M. J., 2006. The fall and recovery of the Tagish Lake meteorite. Meteoritics & Planet. Sci. 41, 407-431. Hiraoka, K., Arakawa, M., Setoh, M., Nakamura, A. M., 2008. Measurements of target compressive and tensile strength for application to impact cratering on ice-silicate mixtures. J. Geophys. Res. 113, E02013 (7 pages). Holsapple, K., Housen, K., 1986. Scaling laws for the catastrophic collisions of asteroids. Società Astronomica Italiana, Memorie 57, 65-85. Holsapple K. A., 1993. The scaling of impact process inn planetary science. Ann. Rev. Earth and Planet. Sci. 21, 333-373. Housen, K. R., Holsapple, K.A., 1990. On the fragmentation of asteroids and planetary satellites. Icarus 84, 226 -- 253. Housen, K. R., Holsapple, K. A., 2011. Ejecta from impact craters. Icarus 211, 856 -- 875. Jutzi, M., Michel, P., Hiraoka, K., Nakamura, A. M., Benz, W., 2009. Numerical simulations of impacts involving porous bodies: II. Comparison with laboratory experiments. Icarus 201, 802-813. Kadono, T., Suzuki, A. I., Araki, S.-I., Asada, T., Suetsugu, R., Hasegawa, S., 2018. Investigation of impact craters on flat surface of cylindrical targets based on experiments and numerical simulations. Planet. Space Sci. 163, 77-82. Kataoka, A., Tanaka, H., Okuzumi, S., Wada, K., 2013. Fluffy dust forms icy planetesimals by static compression. Astron. Astrophys. 557, L4. 4pp. Katsura, T., Nakamura, A. M., Takabe, A., Okamoto, T., Sangen, K., Hasegawa, S., Liu, X., Mashimo, T., 2014. Laboratory experiments on the impact disruption of iron meteorites at temperature of near-Earth space. Icarus 241, 1-12. Kawai, N., Tsurui, K., Hasegawa, S., Sato, E., 2010. Single microparticle launching method using two-stage light-gas gun for simulating hypervelocity impacts of micrometeoroids and space debris. Rev. Sci. Instrum. 81, 115105. Litwin, K. L., Zygielbaum, B. R., Polito, P. J., Sklar, L. S., Collins, G. C., 2012. Influence of temperature, composition, and grain size on the tensile failure of water ice: Implications for erosion on Titan. J. Geophys. Res. 117, E08013 (14 pages). Love, S. G., Hörz, F., Brownlee, D. E., 1993. Target porosity effects in impact cratering and collisional disruption. Icarus 105, 216 -- 224. Machii, N., Nakamura. A. M., 2011. Experimental study on static and impact strength of sintered agglomerates, Icarus 211, 885-893. Marsh, S.P. (Ed.), 1980. Last Shock Hugoniot Data. University of California Press. Melosh, H. J., 1989. Impact Cratering: A Geologic Process. Oxford University Press, New York. Mizutani, H., Takagi, Y., Kawakami, S.-I., 1990. New scaling laws on impact fragmentation. Icarus 87, 307-326. Nakamura, A., Fujiwara, A., 1991. Velocity distribution of fragments formed in asimulated collisional disruption. Icarus 92, 132 -- 146. Nakamura, A., 1993. Laboratory studies on the velocity of fragments from impact disruptions. ISAS Report, 651. Nakamura, A. M., Hiraoka, K., Yamashita, Y., Machii, N., 2009. Collisional disruption experiments of porous targets. Planet. Space Sci. 57, 111-118. Nakamura, A. M., Yamane, F., Okamoto, T., Takasawa, S., 2015. Size dependence of the disruption threshold: laboratory examination of millimeter -- centimeter porous targets. Planet. Space Sci. 107, 45 -- 52. Okamoto, T., Nakamura, A.M., Hasegawa, S., Kurosawa, K., Ikezaki, K., Tsuchiyama, A., 2013. Impact experiments of exotic dust grain capture by highly porous primitive bodies. Icarus 224, 209 -- 217. Okamoto, T., Nakamura, A.M., Hasegawa, S. 2015. Impact experiments on highly porous targets: cavity morphology and disruption thresholds in the strength regime. Planet. Space Sci. 107, 36 -- 44. Okamoto, T., Nakamura, A. M., 2017. Scaling of impact-generated cavity-size for highly porous targets and its application to cometary surfaces. Icarus 292, 234-244. Omura, T., Nakamura, A. M., 2018. Estimating the Porosity Structure of Granular Bodies Using the Lane -- Emden EquationApplied to Laboratory Measurements of the Pressure -- Density Relationof Fluffy Granular Samples. Astrophys. J. 860, 123 (7pp). Poppe, T., 2003. Sintering of highly porous silica-particle samples: analogues of early Solar-System aggregates. Icarus 164, 139-148. Porco, C. C., Thomas, P. C., Weiss, J. W., Richardson, D. C., 2007. Saturn's small inner satellites: Clues to their origins. Science 318, 1602-1607. Rudraswami, N. G., Fernandes, D., Naik, A. K., Shyam Prasad, M., Taylor, S., 2018. Fine- grained volatile components ubiquitous in solar nebula: Corroboration from scoriaceous cosmic spherules. Meteoritics & Planet. Sci. 53, 1207-1222. Ryan, E.V., Davis, D.R., Giblin, I., 1999. A laboratory impact study of simulated Edgeworth -- Kuiper belt objects. Icarus 142, 56 -- 62. Setoh, M., Nakamura, A. M., Michel, P., Hirata, N., Yamashita, Y., Hasegawa, S., Onose, N., Okudaira, K., 2010. High- and low-velocity impact experiments on porous sintered glass bead targets of different compressive strength: Outcome sensitivity and scaling. Icarus 205, 702-711. Shimaki, Y., Arakawa, M., 2012a. Experimental study on collisional disruption of highly porous icy bodies. Icarus 218, 737 -- 750. Shimaki, Y., Arakawa, M., 2012b. Low-velocity collisions between centimeter-sized snowballs: Porosity dependence of coefficient of restitution for ice aggregates analogues in the Solar System. Icarus 221, 310-319. Sierks, H. , Barbieri, C. , Lamy, P. L., Rodrigo, R., Koschny, D., Rickman, H., Keller, H.U., Agarwal, J., A'Hearn, M. F., Angrilli, F., Auger, A.-T., Barucci, M. A., Bertaux, J.-L., Bertini, I., Besse, S., Bodewits, D., Capanna, C., Cremonese, G., Da Deppo, V., Davidsson, B., Debei, S., De Cecco, M., Ferri, F., Fornasier, S., Fulle, M., Gaskell, R., Giacomini, L., Groussin, O., Gutierrez-Marques, P., Gutiérrez, P. J., Güttler, C., Hoekzema, N., Hviid, S. F., Ip, W.-H., Jorda, L., Knollenberg, J., Kovacs, G., Kramm, J.R., Kührt, E., Küppers, M., La Forgia, F., Lara, L. M., Lazzarin, M., Leyrat, C., Lopez Moreno, J. J., Magrin, S., Marchi, S., Marzari, F., Massironi, M., Michalik, H., Moissl, R., Mottola, S., Naletto, G., Oklay, N., Pajola, M., Pertile, M., Preusker, F., Sabau, L., Scholten, F., Snodgrass, C., Thomas, N., Tubiana, C., Vinent, J.-B., Wenzel, K.-P., Zaccariotto, M., Pätzold, M. 2015. On the nucleus structure and activity of comet 67P/Churyumov-Gerasimenko. Science 347, aaa1044 (6 pages). Sirono, S., 1999. Effects by sintering on the energy dissipation efficiency in collisions of grain aggregates. Astron. Astrophys. 347, 720-723. Spohn, T., Knollenberg, J., Ball, A. J., Banaszkiewicz, M., Benkhoff, J., Grott, M., Grygorczuk, J., Hüttig, C., Hagermann, A., Kargl, G., Kaufmann, E., Kömle, N., Kührt, E., Kossacki, K. J., Marczewski, W., Pelivan, I., Schrödter, R., Seiferlin, K. 2015. Thermal and mechanical properties of the near-surface layers of comet 67P/Churyumov-Gerasimenko. Science 349, aab0464 (4 pages). Suzuki, A., Hakura, S., Hamura, T., Hattori, M., Hayama, R., Ikeda, T., Kusuno, H., Kuwahara, H., Muto, Y., Nagaki, K., Niimi, R., Ogata, Y., Okamoto, T., Sasamori, T., Sekigawa, C., Yoshihara, T., Hasegawa, S., Kurosawa, K., Kadono, T., Nakamura, A. M., Sugita, S., Arakawa, M., 2012. Laboratory experiments on crater scaling-law for sedimentary rocks in the strength regime. J. Geophys. Res. 117, E08012 (7 pages). Suzuki, A. I., Okamoto, C., Kurosawa, K., Kadono, T., Hasegawa, S., Hirai, T. (2018) Increase in cratering efficiency with target curvature in strength-controlled craters. Icarus 301, 1-8. Takagi, Y., Mizutani, H., Kawakami, S. 1984. Impact fragmentation experiments of basalts and pyrophyllites. Icarus 59, 462 -- 477. Wada, K., Tanaka, H., Suyama, T., Kimura, H., Yamamoto, T., 2009. Collisional growth conditions for dust aggregates. Asprophys. J. 702, 1490-1501. Yasui M., Arakawa, M., 2011. Impact experiments of porous gypsum -- glass bead mixtures simulating parent bodies of ordinary chondrites: Implications for re-accumulation processes related to rubble-pile formation. Icarus 214, 754 -- 765. Table 1 Sintering conditions and physical properties of targets. 100(cid:3400)𝜙 Porosity (%) 85.8 94.1 89.3 90.6 Compressiv e strength, 𝑌(cid:3030) (MPa) Longitudinal wave speed, Vp (km s-1) Shear wave speed,Vs (km s-1) 1.70±0.43 1.49±0.04 0.088±0.028 0.71±0.01 1.07±0.05 0.43±0.03 0.25±0.06 0.70±0.04 0.325±0.002 0.092±0.022 0.57±0.02 0.33±0.01 Type Peak temp. (°C) HGB87 HGB94 800 650 Mix2:1 800 Height 𝐿(cid:3047) (mm) 56 78 63 64 Bulk density, ρ (g cm-3) 0.36±0.01 0.15±0.00 Diameter1) (mm) 58 78 (t) 65 (b) 66 0.26±0.01 800 Mix1:1 1) t: top surface, b: bottom surface. 68 0.23±0.01 Table 2 Properties of projectiles. Type Shape Nylon Wood Size (mm) 3.175 (diam.) Column 3 (diam.)(cid:3400)2.5 (length) Sphere 1) Chronological Scientific Tables (2019). Density 𝛿 (g cm-3) 1.1 0.74 Strength (MPa) 62 -- 831) 2.4 (tensile) 51 (compressive) 𝑀 Ty pe1 (g) ) 53.994 w 54.820 w 54.406 w 51.938 w 55.682 w 50.860 w n 53.219 n 52.094 n 55.870 n 55.413 50.959 n n 51.712 n 51.958 53.694 n 55.467 n 51.194 w 47.047 w 47.574 w 45.824 w 47.387 w n 46.223 n 53.082 n 55.956 n 56.579 n 54.643 n 54.918 n 56.070 n 55.527 n 55.400 52.649 w 49.540 w Shot # Type m11 HGB87 m10 HGB87 m13 HGB87 m17 HGB87 m15 HGB87 m14 HGB87 m2 HGB87 m12 HGB87 m25 HGB87 m24 HGB87 m8 HGB87 HGB87 m6 HGB87 m5 m3 HGB87 m4 HGB87 m26 HGB94 m30 HGB94 m33 HGB94 m32 HGB94 m29 HGB94 m27 HGB94 m40 mix2:1 m48 mix2:1 m50 mix2:1 m39 mix2:1 m46 mix2:1 m41 mix2:1 m53 mix2:1 m34 mix2:1 m42 mix1:1 m54 mix1:1 𝜃4) (°) 90 63 35 56 86 28 (90) 35 53 42 (90) (90) (90) (90) (90) (90) (90) (90) (90) (90) (90) 21 64 42 59 87 87 (90) 65 75 (90) 𝑏 0.30 0.54 0.52 0.04 0.90 1.0 0.32 >0.04 >0.56 0.25 0.25 0.28 0.31 0.27 0.50 0.06 - 0.47 0.71 0.38 0.49 0.42 0.32 0.28 0.54 0.34 0.49 0.27 0.42 0.41 0.49 𝑀(cid:2869)/𝑀 0.9978 0.9931 0.9902 0.9351 0.9387 0.9303 0.7815 0.7021 0.8124 0.5013 0.3618 0.5405 0.3394 0.3165 0.4021 0.9713 0.9764 0.9949 0.6370 0.3219 0.08434 0.6990 0.9552 0.8655 0.7430 0.4395 0.6520 0.5498 0.3450 0.9862 0.8236 Target support 2) f3) thread thread thread thread thread thread stand1 thread thread thread stand2 stand2 stand2 stand1 stand2 stand2 stand2 stand2 stand2 stand2 stand2 thread thread thread thread thread thread stand2 thread thread stand2 t s s b s s t s s s t t t t t t t t t t t t s s s t t t s s t Table 3 Impact conditions and the largest-fragment mass fraction. Target Projectile 𝑚 (g) 0.013 0.016 0.013 0.013 0.017 0.014 0.019 0.019 0.019 0.019 0.019 0.019 0.019 0.019 0.019 0.012 0.010 0.011 0.012 0.011 0.019 0.019 0.019 0.019 0.019 0.019 0.019 0.019 0.019 0.011 0.012 (km s-1) 𝑉 2.60 2.60 3.32 4.78 5.17 5.27 5.07 5.09 5.23 6.20 6.30 6.31 7.02 7.02 7.09 2.30 2.58 3.03 3.54 5.46 5.24 3.13 3.50 4.00 4.15 4.20 4.28 5.00 5.11 4.79 4.99 5.05 3.76 3.87 4.07 4.30 4.39 5.09 6.10 0.011 0.019 0.019 0.019 0.019 0.019 0.02 0.019 stand2 thread stand2 thread stand2 thread thread thread 49.805 w 54.056 n n 52.688 n 52.180 n 54.479 n 54.600 n 55.370 n 52.461 m51 mix1:1 m47 mix1:1 m55 mix1:1 m49 mix1:1 m52 mix1:1 m45 mix1:1 m43 mix1:1 m445) mix1:1 1) w: wood cylinder, n: nylon sphere 2) Method of supporting target. "thread": target was hung by thread. "stand1": target was put on a plastic stand. "stand2": target was put on a paper stand. 3) The face of target hit by the projectile. "t": top. "s": side. "b": bottom. 4) In case of use of a stand for target support, we did not measure the impact angle, however it is expected almost normal impact (impact angle ~ 90°). 5) Images of one of cameras was not available. t s t t t b s b (90) 62 (90) 66 (90) 86 90 - 0.52 0.16 0.12 0.26 0.29 0.27 0.61 >0.19 0.8651 0.4809 0.5280 0.4065 0.4594 0.2547 0.4586 0.2784 Table 4 Fitted parameters of Eqs. (5) and (6), and the disruption threshold 𝑄(cid:3046)∗. 𝑏(cid:2869) 𝑏(cid:2870) HGB87 2.31(cid:4666)(cid:3399)0.06(cid:4667)(cid:3400)10(cid:2872) 1.22(cid:3399)0.11 3.85(cid:4666)(cid:3399)0.11(cid:4667)(cid:3400)10(cid:2871) 0.77(cid:3399)0.07 HGB94 1.69(cid:4666)(cid:3399)0.11(cid:4667)(cid:3400)10(cid:2872) 1.39(cid:3399)0.26 1.12(cid:4666)(cid:3399)0.11(cid:4667)(cid:3400)10(cid:2871) 0.67(cid:3399)0.13 Mix2:1 2.57(cid:4666)(cid:3399)0.22(cid:4667)(cid:3400)10(cid:2870) 0.75(cid:3399)0.31 2.19(cid:4666)(cid:3399)0.16(cid:4667)(cid:3400)10(cid:2871) 0.66(cid:3399)0.27 0.88(cid:3399)0.41 2.55(cid:4666)(cid:3399)0.25(cid:4667)(cid:3400)10(cid:2871) 0.42(cid:3399)0.19 6.6(cid:4666)(cid:3399)0.7(cid:4667)(cid:3400)10(cid:2870) 𝑎(cid:2870) 𝑎(cid:2869) Target Mix1:1 𝑄(cid:3046)∗ 6.54(cid:4666)(cid:3399)0.33(cid:4667)(cid:3400)10(cid:2871) 1.78(cid:4666)(cid:3399)0.20(cid:4667)(cid:3400)10(cid:2871) 3.44(cid:4666)(cid:3399)0.67(cid:4667)(cid:3400)10(cid:2871) 3.42(cid:4666)(cid:3399)0.54(cid:4667)(cid:3400)10(cid:2871) 𝑑(cid:3030) (mm) 11.4 14.2 14.6 12.7 9.5 10.5 27.0 21.8 19.2 22.3 20.6 24.8 30.9 18.3 24.4 26.3 23.0 31.6 𝑑(cid:3029) (mm) 28.1 25.3 24.3 19.8 21.3 22.1 43.6 37.0 31.6 35.2 39.2 38.2 44.2 33.9 36.0 41.1 35.0 42.0 Target type HGB87 HGB87 HGB87 HGB87 HGB87 HGB87 HGB94 mix2:1 mix2:1 mix2:1 mix2:1 mix2:1 mix1:1 mix1:1 mix1:1 mix1:1 mix1:1 mix1:1 Projectile type n n n n n n w n n n n n n n w n w n 𝑑(cid:3029)/𝐿(cid:3047) 0.502 0.452 0.434 0.354 0.380 0.395 0.559 0.587 0.502 0.559 0.622 0.606 0.691 0.530 0.563 0.642 0.547 0.656 Table 5 Cavity depth. Shot # m2 m3 m4 m5 m6 m8 m32 m39 m40 m41 m46 m53 m43 m45 m51 m52 m54 m55 Table 6 Parameters used in this study. Material HGB87 HGB94 Mix2:1 Mix1:1 Nylon Wood Density (g cm-3) C (km s-1) 0.361 0.151 0.261 0.231 1.11 0.741 0.832 0.512 0.592 0.432 2.64 0.845 s 1.43 1.43 1.43 1.43 1.74 1.55 1: Values for the material of this study. 2: Calculated according to 𝐶(cid:3404)(cid:3495)𝑉(cid:3043)(cid:2870)(cid:3398)(cid:2872)(cid:2871)𝑉(cid:3046)(cid:2870). 3: Assumed value. 4: Marsh (1980). 5: Values of birch wood (Marsh, 1980). Table A-1 Impact conditions and the crater depth of basalt target. 𝑉 (km s-1) 5.23 5.13 5.04 5.03 Depth (cm) 1.14 1.02 0.73 0.78 Projectile 𝑚 (g) 0.043 0.043 0.043 0.043 Shot # 201809-1 201809-2 201811-1 201811-2 Type1 g g g g 1 g: glass spheres with diameter 3.2 mm. Table A-2 Impact conditions and the largest-fragment mass fraction of pumice target. Size (cm) Target 6.0×6.0×6.0 6.0×6.0×6.0 4.0×4.0×4.05 4.0×4.0×3.95 4.05×4.05× 4.03 𝑀 (g) 144 147.8 39.35 41.65 38.74 Shot # 060418-3 060418-4 060822-1 060822-2 060822-4 060824-5 4.05×4.05× 40.19 4.05 060824-6 060825-4 4.0×4.0×4.0 4.03×4.03× 40.09 38.7 4.03 70427 100602-1 100602-3 4.0×4.0×4.0 4.0×4.0×4.0 4.0×4.0×4.0 37.28 42.28 43.53 Projectile 𝑚 (g) 0.213 0.213 0.213 0.213 0.044 𝑉 (km s-1) 2.19 2.58 1.51 2.13 3.27 𝑀(cid:2869)/𝑀 0.123 0.135 0.085 0.037 0.156 0.044 1.64 0.983 0.044 0.213 0.018 0.219 0.219 4.47 3.28 0.125 0.028 3.94 7.10 7.04 0.693 0.007 0.002 type1 n7 n7 n7 n7 g g g n7 n n7 n7 1 n7: nylon sphere with diameter 7 mm, g: glass sphere with diameter 3.2 mm, n: Nylon sphere with diameter 3.2 mm. Fig. 1 Images of target and projectile. (a) SEM image of hollow glass beads. (b) SEM image of perlite grains. (c) SEM image of internal structure of HGB87. (d) Target (mix2:1). (e) Wood projectile. Fig. 2 (a) Compressive strength of the targets used in this study versus depth where the cylindrical specimens were drilled out. (b) Comparison of compressive strength and bulk density of the targets used in this study and other porous materials used in previous studies. Tagish Lake data is shown as a reference. 1 and 6: Okamoto and Nakamura (2017), 2: Hildebrand et al. (2006) for density and Tsuchiyama (private communication) for the lower limit of compressive strength (the original data was of crush strength), 3 and 4: Arakawa and Tomizuka (2004), 5: Shimaki and Arakawa (2012b). Fig.3 Fragments collected after impact (shot #m53). Bulb shaped cavity was formed in the largest fragment (top left corner). Fig. 4 Largest fragment mass fraction versus specific impact energy. Dash-dot, dashed, solid, and dotted lines show the least square fits to the data by Eq. (6). Filled marks show data used in the fitting, while open marks show those not included in the fitting. 1: Data of previous study of the targets with similar porosity to HGB94 (fluffy94 with axial ratio of the target of ~0.9) and the targets with similar porosity and strength to HGB87 (fluffy 87 with axial ratio of the target of ~0.8) (Okamoto et al., 2015). Fig. 5 (a) Spherical cavity formed in shot #m39. Projectile hit at the right side from lower than the horizontal direction. (b) Definition of cavity depth 𝑑(cid:3029) and the depth of the maximum diameter of cavity 𝑑(cid:3030). Fig. 6 Effect of oblique incidence. (a) Ratio of the actual specific energy to the one angle. Data indicated by filled marks in Fig. 4 are plotted here. Previous results of spherical basalt targets (1: Fujiwara and Tsukamoto, 1980; 2: Nakamura and Fujiwara required to obtain the largest fragment 𝑄(cid:3032)(cid:3033)(cid:3033)(cid:3032)(cid:3030)(cid:3047) versus impact parameter and (b) impact 1991 and Nakamura 1993) are also shown. Empirical relationships (cid:3014)(cid:3117)(cid:3014)(cid:3404) 1.96(cid:4666)(cid:3399)0.14(cid:4667)(cid:3400)10(cid:2871)𝑄(cid:2879)(cid:2869).(cid:2870)(cid:2871)(cid:4666)(cid:3399)(cid:2868).(cid:2868)(cid:2873)(cid:4667) for basalt1 and (cid:3014)(cid:3117)(cid:3014)(cid:3404)1.40(cid:4666)(cid:3399)0.11(cid:4667)(cid:3400)10(cid:2871)𝑄(cid:2879)(cid:2869).(cid:2868)(cid:2876)(cid:4666)(cid:3399)(cid:2868).(cid:2872)(cid:2868)(cid:4667) calculate 𝑄(cid:3032)(cid:3033)(cid:3033)(cid:3032)(cid:3030)(cid:3047). for basalt2 were derived based on the results of vertical shots of sphere targets of Fujiwara and Tsukamoto (1980) and Nakamura (1993), respectively, and used to Fig. 7 Normalized depth at the maximum diameter of cavity 𝑑(cid:3030)/𝑑 versus normalized characteristic length. Solid line is the empirical relationship (Eq. 7) obtained in previous study (Okamoto et al., 2015). Fig. 8 (a) Normalized depth of cavity 𝜋(cid:3031)(cid:3029) (filled marks) and normalized the depth at the maximum diameter of cavity 𝜋(cid:3031)(cid:3030) (open marks) versus 𝜋(cid:2871). Previous data (fluffy94, fluffy87, weak_fluffy93) (Okamoto and Nakamura, 2017) are also plotted. Solid line and dashed line are fitted curves to the data of this study. (b) Normalized depth of cavity for porous targets. 1: Okamoto and Nakamura (2017), 2: Okamoto and Nakamura (2017) and Kadono et al. (2018). Compressive strength of 15.6 MPa obtained in a previous study (Fujii and Nakamura, 2009) was assumed, although the target density is slightly different. 3: Arakawa and Yasui (2011). Strength based on dynamic measurement is adopted. Solid line is the same as in (a) and dotted line is the empirical relationship previously obtained for sedimentary rock, in which the tensile strength instead of compressive strength of target was used in 𝜋(cid:2871) (Suzuki et al., 2012). Experiment of basalt target is described in Appendix. Fig. 9 Largest fragment mass fraction versus (a) non-dimensional impact stress, 𝑃(cid:3010)(cid:3030), and (b) corrected version of non-dimensional impact stress where the depth of the center of the isobaric core is take into account. Fig. 10 Largest fragment mass fraction versus PI-group scaling parameters in which (a) compressive strength 𝑌(cid:3030), and (b) tensile strength 𝑌(cid:3047) is taken as the strength, respectively. The values of 𝜇 was assumed 𝜇(cid:3404)(cid:2870)(cid:2871)(cid:3398)(cid:2869)(cid:2871)𝜙.
1610.02848
1
1610
2016-10-10T11:07:46
Exploring Biases of Atmospheric Retrievals in Simulated JWST Transmission Spectra of Hot Jupiters
[ "astro-ph.EP" ]
With a scheduled launch in October 2018, the James Webb Space Telescope (JWST) is expected to revolutionise the field of atmospheric characterization of exoplanets. The broad wavelength coverage and high sensitivity of its instruments will allow us to extract far more information from exoplanet spectra than what has been possible with current observations. In this paper, we investigate whether current retrieval methods will still be valid in the era of JWST, exploring common approximations used when retrieving transmission spectra of hot Jupiters. To assess biases, we use 1D photochemical models to simulate typical hot Jupiter cloud-free atmospheres and generate synthetic observations for a range of carbon-to-oxygen ratios. Then, we retrieve these spectra using TauREx, a Bayesian retrieval tool, using two methodologies: one assuming an isothermal atmosphere, and one assuming a parametrized temperature profile. Both methods assume constant-with-altitude abundances. We found that the isothermal approximation biases the retrieved parameters considerably, overestimating the abundances by about one order of magnitude. The retrieved abundances using the parametrized profile are usually within one sigma of the true state, and we found the retrieved uncertainties to be generally larger compared to the isothermal approximation. Interestingly, we found that using the parametrized temperature profile we could place tight constraints on the temperature structure. This opens the possibility to characterize the temperature profile of the terminator region of hot Jupiters. Lastly, we found that assuming a constant-with-altitude mixing ratio profile is a good approximation for most of the atmospheres under study.
astro-ph.EP
astro-ph
Draft version November 6, 2018 Preprint typeset using LATEX style AASTeX6 v. 1.0 EXPLORING BIASES OF ATMOSPHERIC RETRIEVALS IN SIMULATED JWST TRANSMISSION SPECTRA OF HOT JUPITERS M. Rocchetto1 , I. P. Waldmann1, O. Venot2, P.-O. Lagage3,4, G. Tinetti1 1 Department of Physics & Astronomy, University College London, Gower Street, WC1E6BT London, United Kingdom 2 Instituut voor Sterrenkunde, Katholieke Universiteit Leuven, Celestijnenlaan 200D, 3001 Leuven, Belgium 3 Irfu, CEA, Universit´e Paris-Saclay, F-9119 Gif-sur Yvette, France and 4 AIM, Universit´e Paris Diderot, F-91191 Gif-sur-Yvette, France ABSTRACT With a scheduled launch in October 2018, the James Webb Space Telescope (JWST ) is expected to revolutionise the field of atmospheric characterization of exoplanets. The broad wavelength coverage and high sensitivity of its instruments will allow us to extract far more information from exoplanet spectra than what has been possible with current observations. In this paper, we investigate whether current retrieval methods will still be valid in the era of JWST, exploring common approximations used when retrieving transmission spectra of hot Jupiters. To assess biases, we use 1D photochemical models to simulate typical hot Jupiter cloud-free atmospheres and generate synthetic observations for a range of carbon-to-oxygen ratios. Then, we retrieve these spectra using TauREx, a Bayesian retrieval tool, using two methodologies: one assuming an isothermal atmosphere, and one assuming a parametrized temperature profile. Both methods assume constant-with-altitude abundances. We found that the isothermal approximation biases the retrieved parameters considerably, overestimating the abundances by about one order of magnitude. The retrieved abundances using the parametrized profile are usually within one sigma of the true state, and we found the retrieved uncertainties to be generally larger compared to the isothermal approximation. Interestingly, we found that using the parametrized temperature profile we could place tight constraints on the temperature structure. This opens the possibility to characterize the temperature profile of the terminator region of hot Jupiters. Lastly, we found that assuming a constant-with-altitude mixing ratio profile is a good approximation for most of the atmospheres under study. Keywords: methods: data analysis methods: statistical radiative transfer techniques: spectroscopic 1. INTRODUCTION The discovery of over three thousands exoplanets has unveiled a large and diverse population. We see planets in a range of sizes, temperatures, and orbits, far exceed- ing the diversity found in our own Solar System. To- day, emphasis in the field of exoplanets is shifting from the discovery to the characterization of these exoplane- tary bodies, as understanding their nature will in turn provide important clues on the planet's formation and evolution history. The study of exoplanetary atmospheres represents one [email protected] of the most immediate and direct ways to characterize exoplanets. To date, the atmospheres of several tens of giant planets, sub-Neptunes and super-Earths have been studied and characterized with the Hubble Space Tele- scope (HST ) (Charbonneau et al. 2002; Tinetti et al. 2010; Swain et al. 2013; Kreidberg et al. 2014; Sing et al. 2016; Tsiaras et al. 2016) Spitzer space telescope (Tinetti et al. 2007; Knutson et al. 2007; Grillmair et al. 2007; Charbonneau et al. 2008; Beaulieu et al. 2010; Stevenson et al. 2010; Knutson et al. 2011; Deming et al. 2011; Todorov et al. 2013) and other ground based facili- ties (Redfield et al. 2008; Snellen et al. 2008; Swain et al. 2010; Waldmann et al. 2012; Bean et al. 2013; Danielski et al. 2014; Zellem et al. 2014; Brogi et al. 2014). With the imminent launch of the JWST, it has be- 2 come fundamental to assess whether the current meth- ods used to interpret these spectra will still be valid when higher quality datasets will be available. In this work we aim to answer this question in part, explor- ing the biases induced by common assumptions used in atmospheric retrievals. One of the major limitations of current observations is the limited wavelength coverage. The best quality datasets, which led to the confident detection of water vapor in several hot Jupiters and warm Neptunes, have been mainly obtained with the Wide Field Camera 3 onboard HST, covering the spectral range 1.1 -- 1.7 µm. Nevertheless, it is at longer wavelengths that most roto- vibrational transitions of molecular species occur. While the Spitzer Space Telescope has given some insight into the longer wavelength regime to several tens of close-in hot Jupiters, the data have relatively large uncertainties, and they are mostly only photometric measurements. Significant advances in the field of atmospheric char- acterization can therefore only happen if high quality observations extending to the longer wavelength regime are obtained. In this scenario, the James Webb Space Telescope (JWST ) will undoubtedly revolutionise the field of exo- planetary atmospheres, addressing two major problems affecting current observations: wavelength coverage and instrument sensitivity. With a scheduled launch for 2018 October, the large spectral coverage (0.7 -- 28 µm) cov- ered by its multiple instruments, combined with high sensitivity and high degree of instrumental characteri- zation and calibration, will ensure a significant advance in atmospheric characterization (Beichman et al. 2014; Cowan et al. 2015; Batalha et al. 2015; Greene et al. 2016; Barstow et al. 2015; Barstow & Irwin 2016). Atmospheric spectra of transiting exoplanets in a broad spectral range will enable us to constrain the abundances of different molecular species, the temper- ature structure of the atmosphere, and the presence or absence of clouds and hazes. In the case of warm H/He dominated atmospheres one of the key elemental ratio that we aim to constrain is the carbon-to-oxygen ratio (C/O). Such measurements will enable us to distinguish between different formation and migration scenarios, so far poorly constrained ( Oberg et al. 2011; Madhusudhan et al. 2011b; Ali-Dib et al. 2014; Thiabaud et al. 2015). While transmission and emission spectra do not pro- vide direct constraints on the elemental abundances, the measurement of the absolute abundances of O-bearing and C-bearing molecules will provide some constraints on the C/O ratio. In particular, the excess carbon and oxygen not locked in CO will form either oxygen-bearing molecules such as H2O in atmospheres with C/O < 1, or, in atmospheres with C/O > 1, carbon-rich species such as HCN, C2H2 and CH4 (Madhusudhan 2012; Moses et al. 2013a,b; Venot et al. 2015). Determining the atmospheric abundances of these gases in hot Jupiters with high accuracy is therefore paramount and JWST will give us direct access to absorption features of these molecules both in emission and transmission. Determining the absolute abundances of atmospheric gases from atmospheric spectra requires the use of re- trieval methods. Atmospheric retrieval techniques are now commonly used to infer the properties of exo- planetary atmospheres, including molecular abundances and temperature profiles (Madhusudhan & Seager 2009; Madhusudhan et al. 2011a; Benneke & Seager 2012, 2013; Lee et al. 2011; Line et al. 2012, 2013, 2014; Ir- win et al. 2008; de Wit & Seager 2013; Waldmann et al. 2015a,b). These tools enable us to fully map the likeli- hood space of atmospheric models, and to place upper limits and constraints on the abundances of molecules and temperature profiles. The lack of high signal-to-noise and broad wave- length coverage observations have however led current retrievals and forward models to make several assump- tions and approximations to reduce the parameter space. The forward model included in most retrieval methods is a 1D radiative transfer model (Brown 2001; Seager 2011; Tinetti et al. 2012; Liou 2002; Hollis et al. 2013), imple- menting opacity cross sections for the major molecular absorbers, Rayleigh scattering and collision induced ab- sorption. Transmission spectra are usually retrieved as- suming constant-with-altitude temperature and molec- ular abundances. This might be a fair approximation when probing narrow wavelength ranges, but can lead to significant biases when larger wavelength ranges are probed. One of the pressing questions we are facing to- day is whether these assumptions will still be valid in the era of JWST. In this paper, we aim to address these issues. We study the biases and degeneracies of atmospheric re- trievals of high quality, broad wavelength range trans- mission spectra of hot Jupiters, such as those that will be obtained with instruments onboard JWST. We apply and compare different retrieval approaches to synthetic observations for a range of hot atmospheres with dif- ferent C/O computed using photochemical models, and study the biases of common assumptions used in today's retrievals. This study aims at answering the following questions: a) Are our retrieval approaches and forward mod- els appropriate for the high signal-to-noise, and broader wavelength range spectra expected from future facilities such as JWST ? b) Can we confidently retrieve absolute molecular abundances and infer the C/O ratio? In Section 2 we describe the chemical and radiative transfer models used to generate the synthetic trans- mission spectra. We also present the JWST synthetic observations, and describe the two retrieval approaches used to interpret these synthetic observations. In Sec- tion 3 we describe qualitatively the simulated transmis- sion spectra and present the results of the retrievals. In Section 4 we discuss our results, and in Section 5 we summarize the main conclusions of this study. 2. METHOD 2.1. Chemical models The 1D atmospheric chemical models were generated using the photochemical model developed for hot at- mospheres (Venot et al. 2012, and references therein). These models have been used to study exoplanets (Venot et al. 2014; Ag´undez et al. 2014; Venot et al. 2015; Venot & Ag´undez 2015; Tsiaras et al. 2016) as well as Solar System giant planets (Cavali´e et al. 2014; Mousis et al. 2014). The chemical scheme has been developed with combustion specialists and validated in a wide range of pressures (0.001 -- 100 bar) and temperatures (300 -- 2500 K), making this model one of the currently most reliable chemical schemes (Battin-Leclerc et al. 2006; Bounaceur et al. 2007; Anderlohr et al. 2010; Wang et al. 2010). Venot et al. (2015) showed that the use of more com- plete chemical models, including species with up to six carbon atoms, has little effect on the synthetic spec- tra. We therefore used the simpler, and computation- ally faster, scheme which includes species with up to four carbon atoms and is able to model the kinetic behaviour of species with up to two carbon atoms. This scheme includes 105 neutral species and 960 reactions (and their reverse reactions). We used a constant diffusion coeffi- cient, Kzz = 108 cm2s−1 due to the uncertainties on the vertical mixing acting in exoplanet atmospheres. A sim- ilar value has been often used in the literature (Lewis et al. 2010; Moses et al. 2011; Line et al. 2011; Venot et al. 2013). We note however that this value might be too high, as pointed out by Parmentier et al. (2013). We used a temperature-pressure (TP) profile with a high-altitude temperature of 1500 K. The vertical pro- file is the same as the one used in Venot et al. (2015). It was computed using the analytical model one of Par- mentier & Guillot (2014), using coefficients from Par- mentier et al. (2015) and the opacities from Valencia et al. (2013). The profile, shown in Figure 1, was ob- tained by setting the irradiation temperature to 2300 K and the internal temperature Tint = 100 K. We assumed a planet with Rp = 1.162 J and Mp = 1.138 MJ . We computed chemical models for an atmosphere of solar metallicity with C/O of 0.5, 0.7, 0.9, 1.0, 1.1, 1.3, and 1.5. 3 Figure 1. Temperature-pressure profile used for the atmo- spheres under study. 2.2. Synthetic high resolution transmission spectra High resolution (R ≈ 10000) synthetic transmission spectra were computed using the forward models in- cluded in TauREx Waldmann et al. (2015b). This for- ward model is based on a 1D radiative transfer model that calculates the optical path through the planetary It results in a transmission spectrum of atmosphere. (Rp/R∗)2, as a function of wave- transit depth, i.e. length. The temperature profile used is the same as the one used for the computation of the photochemical mod- els (Figure 1). We include a precise computation of the pressure-altitude profile, and take into account the effect of gravity, temperature and mean molecular weight in the computation of the scale height in each of the 100 at- mospheric layers included in the model. We compute the pressure grid from 10−4 to 10 bar, and define the 10 bar pressure radius to be Rp = 1.162 RJ . The mass is set to Mp = 1.138 MJ . Amongst the 105 molecules considered in the photochemical model we only consider the follow- ing seven molecules in the computation of the opacity in the synthetic spectra: C2H2, CH4, CO, CO2, H2O, HCN and NH3. We found that amongst the complete set of 105 molecules contained in the chemical model, these are the most abundant ones in all cases and will there- fore dominate the spectral modulation. The wavelength dependent cross sections for these absorbing molecules were computed using line lists from ExoMol (Barber et al. 2006; Harris et al. 2006; Yurchenko et al. 2011; Tennyson & Yurchenko 2012; Yurchenko et al. 2013; Barber et al. 2014), HITRAN (Rothman et al. 2013) and HITEMP (Rothman et al. 2010). Note that the mean molecular weight of each atmospheric layer is coupled to the mixing ratio of all 105 molecules. We included additional opacity from Rayleigh scattering of H2 and from collision induced absorption of He and H2-H2 and H2-He pairs (Richard et al. 2012). 2.3. JWST spectra 1400180022002600Temperature(K)10−610−510−410−310−210−1100101Pressure(bar) 4 Table 1. JWST instrument modes Instrument Mode Wavelength range (µm) NIRISS SOSS/GR700XD 1.0 -- 2.5 µm NIRCam LW grism/F322W2 2.5 -- 3.9 µm NIRCam LW grism/F444W 3.9 -- 5.0 µm MIRI slitless/LRS prism 5.0 -- 10.0 µm Figure 2. Simulated JWST observation for C/O = 0.5. The spectrum was obtained combining four separate synthetic observations obtained with NIRISS, NIRCam and MIRI to cover the 1 -- 10 µm spectral range. This spectrum would therefore require observing a total of four transits. We simulated spectra for the Near-InfraRed Imager and Slitless Spectrograph (NIRISS) in Single-Object Slitless Spectroscopy (SOSS) mode using the GR700XD optics (Doyon et al. 2012). We applied a lower wave- length cutoff at 1 µm to avoid saturation and a long wavelength cutoff at 2.5 µm to avoid spectral contam- ination (Greene et al. 2016). We then used the Near Infrared Camera (NIRCam) using the long wavelength (LW) channel and the F322W2 and F444W filters, cov- ering the 2.5 -- 3.9 and 3.9 -- 5.0 µm spectral ranges re- spectively (Greene et al. 2007). An alternative could be the use of the Near Infra Red SPECtrometer (NIR- SPEC) in its high resolution mode with the 2 instrumen- tal configurations: F170LP/G235H (1.7 -- 3.1 µm] and F290LP/G395H (2.9 -- 5.2 µm) (Ferruit et al. 2014). Fi- nally, we use the Mid Infrared Instrument (MIRI) to cover the 5.0 -- 10.0 µm wavelength range. We use MIRI in slitless mode, using the Low Resolution Spectrom- eter (LRS) and we apply a long-wavelength cutoff of 10 µm due to the degrading S/N at longer wavelengths (Kendrew et al. 2015). Each observation covering the full wavelength range 1 -- 10 µm will therefore require four separate observations. We have considered a one hour effective integration time during the transit and the same amount of time on the star alone. For each mode, the same amount of time was used. Table 2 sum- marizes the instrument modes considered in this study. The noise in the spectra was calculated taking into ac- count the star photon noise, the zodiacal and telescope background noise (integrated over the entire band pass of the spectrometer for the slitless mode), the detector dark current and noise. We assumed a star similar to HD189733. The star spectrum used was generated us- ing the PHOENIX atmosphere star code (Husser et al. 2013). For NIRISS and NIRCAM, we have binned the spectra to a constant spectral resolution of R = 100. For such a bright star we realized that we are in fact very close to the limitation from systematics of the JWST. Such systematics are difficult to assess but we can rea- sonably assume that they will be lower than HST. Given the latest performances achieved with HST (e.g. Tsiaras et al. 2016), we can anticipate that the systematics for NIRISS and NIRCAM will be better than about 20 ppm. For MIRI, Greene et al. (2016) adopted a value of 50 ppm, and Beichman et al. (2014) took a value of 30 ppm; we have adopted an intermediate value of 40 ppm. An example of a final spectrum is shown in Figure 2. 2.4. Atmospheric Retrieval The analysis and interpretation of the simulated ob- served spectra was carried out using TauREx (Wald- mann et al. 2015b). Recently, TauREx has been used to model the spectra obtained with the Wide Field Camera 3 onboard the Hubble Space Telescope for HD209458b and 55 Cnc e (Tsiaras et al. 2015, 2016) Two retrieval approaches were used as part of the cur- rent study. Both approaches did not assume any prior knowledge on the chemistry, i.e. the absolute abundance of all gases taken into account is fitted independently. The only difference between the two approaches is in the parametrization of the temperature profile: • In the first case we assumed an isothermal TP pro- file. We will refer to this method as "tp-iso". This approach is the most commonly used when fitting transmission spectra (Line et al. 2012; Ben- neke & Seager 2012; Irwin et al. 2008), and in- cludes a parametrization of the atmosphere assum- ing constant-with-altitude mixing ratio and tem- perature profiles. Crucially, it does not assume any prior on the chemistry of the atmosphere. The free parameters of the retrieval were the ab- solute abundance of each atmospheric constituent taken into account, the isothermal temperature, the cloud parameters and the 10 bar pressure ra- dius. The mean molecular weight is coupled to the fitted composition, and we assumed the bulk atmosphere to be formed by a mixture of hydro- gen and helium, whose ratio is fixed to solar value (85% H2 and 15% He). We assumed uniform pri- ors in log space for the absolute abundances, rang- 12345678910Wavelength(µm)0.02340.02360.02380.02400.02420.02440.0246(Rp/R∗)2 Table 2. Free parameters of the two retrieval approaches used in this study. The tp-iso approach refers to the retrieval using an isothermal TP profile, while the tp-param refers to the retrieval using a parametrized TP profile. 5 Approach tp-iso (10 free parameters) tp-param (14 free parameters) 1300 . . . 2600 1.05. . . 1.28 Prior Parameter log H2O, log CO, log CO2, −12 . . . 1 log CH4 log NH3, log HCN, log C2H2 Tiso [K] Rp [RJup] log(Ptop [Pa]) log H2O, log CO, log CO2, −12 . . . 1 log CH4 log NH3, log HCN, log C2H2 Tirr [K] log κIR log κν1, log κν2 α 0 . . . 6 Description Molecular abundances Isothermal temperature Planetary radius at 10 bar Cloud top pressure Molecular abundances Rp [RJup] log(Ptop [Pa]) ing from 10−12 to 1. We assumed uniform priors for the temperature (1300 -- 2500) K and for the 10 bar radius (1.05 -- 1.28 RJup). The prior width of the 10 bar radius was determined by assuming a relative uncertainty on Rp of 20% (Rp = 1.162 RJ ). Lastly, we fitted the cloud top pressure with a uniform prior in log space (10−5 -- 10 bar). This parametrization resulted in 10 free variables. • In the second case, we assumed a more com- plex TP profile described by five separate parame- ters. We will refer to this method as "tp-param". Since the temperature profile of the atmospheres under study is highly non-isothermal for pressures greater than 1 mbar (see Figure 1), fitting an isothermal profile might lead to biases. We there- fore investigated the effectiveness of fitting a more complex profile using this second method. We used the parametrization of Guillot (2010) modi- fied by Line et al. (2013) and Parmentier & Guillot (2014). There are five parameters that define the temperature profile: the planet internal heat flux (Tint), the stellar irradiation flux (Tirr), the opac- ities in the optical and infrared (κν1, κν2), and a weighting factor between optical opacities (α). For a full description of this model we refer the reader to Section 3.1 in Line et al. (2013). These five pa- rameters replace the single parameter used for the isothermal profile in the first method. This model only differs from the first one for the type of TP profile used. This parametrization resulted in 14 free variables. The parametrized profile described above is commonly 1300 . . . 2600 Stellar flux at the top of the atmosphere −4 . . . 1 −4 . . . 1 0 . . . 1 Optical opacity sources Mean infrared opacity Weighting factor for κν1and κν2 Planetary radius at 10 bar 1.05. . . 1.28 0 . . . 6 Cloud top pressure used in the retrieval of emission spectra, where the spec- tral features are more sensitive to temperature gradi- ents than in transmission. It has received little atten- tion in the retrieval of transmission spectra, as it is as- sumed that transmission spectra are much less sensi- tive to temperature gradients, and therefore isothermal profiles, thought to represent the "average" atmospheric temperature, have always been used. Previous studies have addressed the potential bias of the isothermal as- sumption (Barstow et al. 2013), and found that some in- formation on the temperature profile could be retrieved in transmission only in the highest signal-to-noise and broad wavelength coverage cases. We used these two approaches to interpret the syn- thetic JWST observations in a range of C/O. In all cases we used the MultiNest sampling algorithm (Feroz & Hobson 2008) to finely sample the parameter space and obtain the posterior distributions of the model pa- rameters. We chose this method instead of a more classi- cal MCMC, as MultiNest can better map the likelihood of highly degenerate parameter spaces. Table 2 sum- marises the free parameters and the corresponding prior widths used in the two retrieval methods. 3. RESULTS 3.1. Chemical models and transmission spectra Figure 3 shows the vertical abundance profiles of seven molecules for all C/O ratios considered in this study, and Figure 4 shows the synthetic transmission spectra and contributions of the major opacity sources for the same C/O values. It can be clearly seen that the chemistry and the resulting spectra change significantly between 6 Figure 3. Vertical abundance profiles for different molecules for a range of C/O. The different coloured lines show the molar fraction profiles at different C/O, as shown by the legend. Figure 4. Synthetic transmission spectra (black lines) and contributions of the major opacity sources (colored lines, see legend) for the atmospheres whose chemistry is shown in Figure 3, for different C/O values. The opacity sources include the seven molecules considered in this study, and the collision induced absorption (CIA) from H2 -- H2 and H2 -- He pairs. Note that for each plot we only show the major opacity contributors to the spectrum, and we hide the molecules that do not significantly contribute to the transmission spectrum features. 10−710−510−310−1101Pressure(bar)C2H2CO2HCN10−1110−910−710−510−3MolarfractionNH310−1110−910−710−510−3Molarfraction10−710−510−310−1101Pressure(bar)CH410−1110−910−710−510−3MolarfractionCO10−1110−910−710−510−3MolarfractionH2OC/O=0.5C/O=0.7C/O=0.9C/O=1.0C/O=1.1C/O=1.3C/O=1.50.02250.02300.02350.02400.0245(Rp/R∗)2C/O=0.5C/O=0.7C/O=0.90.02250.02300.02350.02400.0245(Rp/R∗)2C/O=1.012345678910Wavelength(µm)C/O=1.112345678910Wavelength(µm)C/O=1.312345678910Wavelength(µm)0.02250.02300.02350.02400.0245C/O=1.5HCNCOCIAC2H2H2OCO2NH3CH4 C/O < 1, C/O = 1 and C/O > 1. Firstly, we note that while the transmission spectra of an oxygen rich atmosphere are dominated almost en- tirely by H2O, with additional features from CO at 4.6 µm and from CO2 at 4.3 µm, a carbon rich atmosphere is dominated by HCN and CH4, with additional features from CO at 4.6 µm and C2H2 at 1.7, 3.0 and 7.5 µm. At the C/O = 1.0 threshold the transmission spectrum is dominated by H2O and HCN, and exhibits strong fea- tures of CO at 2.3 and 4.6 µm. Weak features from CH4 are also seen at 3.4 and 7.6 µm. Tight constraints on the abundances of all these molecules is therefore paramount to constrain the chemistry and C/O of these atmospheres. Between C/O = 0.5 and 0.9 we see a gradual decrease in the molar fractions of H2O and CO2, and a slight increase in the CH4, HCN and C2H2 abundance, while CO remains relatively constant. The resulting trans- mission spectra in this C/O range show the progressive decrease in the absorption of H2O (which remains the dominant absorber across this C/O range) and the re- sulting emergence of CO, while all the other molecules remain hidden. It is only at C/O > 1.0 that HCN is suf- ficiently abundant to be clearly seen in the transmission spectrum. We note that at this threshold we see the minimum average absorption from active gases across most of the spectrum, so that in some regions we can also see the emergence of the collision induced absorp- tion from H2 -- H2 and H2 -- He pairs. At C/O = 1.1 the H2O and CO2 content drastically drops, while the abundances of CH4, HCN and C2H2 increase significantly. The corresponding transmission spectra show features of CH4, HCN, CO, and C2H2. At progressively higher C/O ratios we see the increase in abundance of CH4, HCN, and C2H2, and the progressive decrease of CO abundance. However, we note that the resulting spectra are very similar to each other. The only differences in the spectra are the weakening of CO at 4.6 µm and the strengthening of C2H2 at 3 and 7.5 µm. Finally, we note that C2H2 might actually have addi- tional and much stronger features than those seen here. This is because the line list used for this molecule comes from HITRAN and has been computed experimentally at Earth-like temperatures. It is therefore sub-optimal to use this line list for such high temperatures (> 1500 K). As an appropriate hot line list would include many more transitions resulting from the population of higher vibrational levels, additional spectral features (i.e. "hot bands") are expected, together with the strengthening of the features that can already be seen at lower tem- 7 peratures. Such a list is under development at ExoMol1 (private communication). 3.2. Retrieval of temperature profiles Figure 5 shows the retrieved temperature profiles us- ing the two approaches for all C/O values. It can be clearly seen that in most cases the retrieved TP profile is within 1 sigma of the input profile using the tp-param method, while using the tp-iso method the input pro- file is almost entirely outside the 1 sigma retrieved error bars. For C/O > 1 (first three plots), it can be seen how the tp-param method fits both the upper atmosphere temperature and the lower-altitude part of the atmo- sphere. We found that the upper atmospheric temper- ature could be well fitted within about 1 to 3 sigma using the parametrized TP profile in all cases. The high- altitude temperature was found to be T = 1502 ± 66 K, T = 1425 ± 27 and T = 1433 ± 50 K for C/O = 0.5, 0.7 and 0.9 respectively. Using the tp-iso method the the retrieved temperatures for the same C/O values were T = 1572 ± 14 K, T = 1610 ± 17 K and T = 1716 ± 24 K, respectively. In all cases, the input profile has a high altitude temperature of 1500 K. From these plots we can also appreciate that the non- isothermal part of the profile could be fitted within one sigma for C/O = 0.5, 0.7 and 0.9. Interestingly, we also note that for C/O < 1 the constraint of the low-altitude temperature (P > 10−3 bar) improves for higher C/O, while the fit of the high-altitude part of the profile (P > 10−3 bar) improves for lower C/O. The last three plots in Figure 5 show the retrieved temperature profiles using both approaches for C/O > 1. We can see that the tp-iso approach retrieves a tem- perature of ≈ 2000 K, with an uncertainty of ≈ 20 K in all cases. Using the parametrized approach we could fit the high-altitude temperature within about 1 sigma for C/O = 1.1 and 1.5, and within 3.4 sigma for C/O = 1.3. We also note that while the low-altitude part of the TP profile for C/O = 1.1 and 1.3 is well constrained within about 1 sigma, for C/O = 1.5 the fit is poor for pressures higher than 0.1 bar. For C/O = 1 we note that the TP profile is poorly retrieved, with the tp-param method giving slightly better results. In both cases however the input pro- file cannot be retrieved within several sigma: the re- trieved upper atmosphere temperature is 6 and 18 sigma away from the true state using the tp-param and tp- iso methods respectively. Additionally, the lower atmo- sphere temperature (P < 0.1 bar) is not retrieved in both cases. 1 http://www.exomol.com 8 Figure 5. Retrieved temperature profiles for the approach with isothermal profile (pink) and parametrized profile (blue) for different C/O . The red line shows the input profile. The shaded areas show the 1 sigma confidence level. 3.3. Retrieval of atmospheric abundances The atmospheric retrieval results for the atmospheric abundances of H2O, CO, CO2, CH4, HCN, C2H2 and NH3 are shown in Table 1 in the Appendix and in Fig- ures 6 and 7. In these plots, the input mixing ratios for each molecule at each C/O are also shown with solid lines as a function of pressure. We note again that the retrieved abundances are constant-with-altitude, so that a single parameter is retrieved for each molecule using both the tp-iso (left plots) and tp-param (right plots) retrieval methods. Moreover, we found that NH3 is never well retrieved, hence we do not show its retrieved values in these figures. This is not surprising given that NH3 is never seen in the simulated transmission spectra (Figure 4). In general, we found that the tp-iso method retrieves higher abundances by about one order of magnitude and significantly underestimates the error bars, causing strong biases, while the tp-param method retrieves the correct true state within one sigma for all atmospheres with C/O greater and less than 1, but not for C/O = 1. This is also because the retrieved error bars are signifi- cantly larger. Looking at the transmission spectra for C/O < 1 in Figure 4 it can be seen that H2O has multiple features across the entire wavelength range and is therefore the dominant molecule. Indeed, we found that, for these C/O values, the retrieved abundance of H2O has the smallest uncertainties, but only the approach using the parametrized TP profile gives unbiased results. Inter- estingly, the retrieval method using the isothermal ap- proximation was found to bias the results significantly. For example, for C/O = 0.7 the true abundance for H2O altitude. The retrieved abundance using the isothermal at 0.1 bar is 2.5 × 10−4 and is relatively constant with approximation was found to be 3 − 4 × 10−3, and 16 sigma away from the true value. On the contrary, the retrieved abundance using the parametrized TP profile is 1.8− 6.3× 10−4 and well within 1 sigma from the true value. For C/O = 0.5 and 0.9 we see similar results: us- ing the tp-param method the true state is within 1 to 2 sigma of the retrieved values, but if we use the tp-iso method, the same retrieved values are 15 and 20 sigma away respectively from the true state. The two other molecules that contribute to the spec- trum, CO and CO2, were found to be highly degenerate, but could be retrieved within 1 to 2 sigma using the tp- param method. Using the tp-iso method, abundances were however overestimated. For CO2 we found that us- ing the parametrized TP profile the true state is within 2 sigma of the retrieved state for C/O = 0.5, and within 1.5 and 1.1 sigma for C/O = 0.7 and 0.9 respectively. Using the isothermal profile, we obtain retrieved values that are significantly overestimated, and are 11.0, 10.2 and 9.6 sigma away from the true state for C/O = 0.5, 0.7 and 0.9 respectively. In these hot oxygen-rich atmo- spheres the retrieved abundances of CO and CO2 must however be interpreted with caution, as both molecules have the only detectable feature in the same wavelength range (≈ 4.0 − 5.5 µm). From Figure 8, showing the posterior distribution of CO and CO2 for C/O = 0.7 using the parametrized TP-profile, it can be appreci- ated that the retrieved absolute abundances for these two molecules are highly degenerate. For C/O < 1 no other molecules could be retrieved, and only upper lim- its could be obtained. The transmission spectra of these atmospheres with C/O > 1 show that the dominant molecules are CH4, HCN, C2H2 and CO, while all other molecules remain hidden below these stronger absorbers (Figure 4). Only these dominant absorbers could be retrieved, while for all other molecules only upper limits could be placed. We found that also for these carbon-rich atmospheres the tp-param retrieval method gives considerably bet- 1.01.52.02.53.0T(×1000K)10−610−510−410−310−210−1100101Pressure(bar)C/O=0.51.01.52.02.53.0T(×1000K)C/O=0.71.01.52.02.53.0T(×1000K)C/O=0.91.01.52.02.53.0T(×1000K)C/O=1.01.01.52.02.53.0T(×1000K)C/O=1.11.01.52.02.53.0T(×1000K)C/O=1.31.01.52.02.53.0T(×1000K)C/O=1.5 9 Figure 6. Retrieved H2O (top), CO (middle) and CO2 (bottom) abundance for C/O = 0.5 -- 1.5 using the approach with isothermal profile (left) and parametrized TP profile (right). The solid lines show the input mixing ratio profiles for different C/O, with different colors corresponding to different C/O, as shown by the legend. The retrieved absolute mixing ratios for the different C/O are shown with error bars. Note that we retrieve constant-with-altitude mixing ratio profiles. Note also that the vertical position of the retrieved values are arbitrary. ter results. Figure 7 shows the retrieved CH4, HCN and C2H2 abundances. For C/O > 1 we can see that the in- put abundance profiles change significantly as a func- tion of pressure, especially for CH4. In the case of CH4 we found that the tp-iso method significantly overesti- mates the abundances. For all C/O > 1 the retrieved abundances are higher than the true abundances at all pressures in the atmosphere. More reasonable results are obtained with the tp-param method, where the re- trieved abundances are always between the maximum and minimum true abundance. For the same carbon-rich atmospheres, the retrieved abundances of HCN and C2H2 using the tp-param ap- proach are all within 1 to 2 sigma of the input abun- dance, while the values obtained with the tp-iso are always overestimated by about one order of magnitude, and are 8 to 11 sigma away from the true state. Lastly, we note that the retrieved abundances of CO were within 1 sigma of the true state using the tp-param method, while using the tp-iso method the same values are an order of magnitude higher than the true state, and have underestimated error bars. This case with C/O = 1 is the most peculiar as many molecules are visible in the spectrum, and their abun- dance varies significantly as a function of altitude. In the case of H2O, CO2 and CH4 the true abundance profile changes by about one order of magnitude at the typ- ical pressures probed by transmission spectra (10−1 -- 10−4) bar (see Figure 3). For H2O, small differences are seen between the tp-param and tp-iso methods. The retrieved abundances are 1.4 − 1.6 × 10−6 in the first case, and 1.6 − 1.9 × 10−6 in the second case, while the input profile varies between 5 × 10−7 and 4 × 10−6 for 10−1210−1110−1010−910−810−710−610−510−410−310−2H2Omixingratio10−410−310−210−1100Pressure(bar)TP-ISO10−1210−1110−1010−910−810−710−610−510−410−310−2H2OmixingratioTP-PARAMC/O=0.5C/O=0.7C/O=0.9C/O=1.0C/O=1.1C/O=1.3C/O=1.510−410−310−2COmixingratio10−410−310−210−1100Pressure(bar)TP-ISO10−410−310−2COmixingratioTP-PARAMC/O=0.5C/O=0.7C/O=0.9C/O=1.0C/O=1.1C/O=1.3C/O=1.510−1210−1110−1010−910−810−710−610−5CO2mixingratio10−410−310−210−1100Pressure(bar)TP-ISO10−1210−1110−1010−910−810−710−610−5CO2mixingratioTP-PARAMC/O=0.5C/O=0.7C/O=0.9C/O=1.0C/O=1.1C/O=1.3C/O=1.5 10 Figure 7. Retrieved CH4 (top), HCN (middle) and C2H2 (bottom) abundance for C/O = 0.5 -- 1.5. Caption as in Figure 6. pressures between 1 and 10−4 bar. For carbon monox- ide, in both cases the retrieved abundances are over- estimated by about one order of magnitude, with val- ues 6 to 7.5 sigma away from the true state. This is somewhat surprising, considering that the input profile is constant with altitude. For CO2 the retrieved abun- dance is within 1 sigma using the tp-param method, and within 2 sigma using the tp-iso approach. Finally, for CH4 and HCN the retrieved abundances are very similar using both methods, and are found to be within the maximum and minimum abundances of the input profiles, which both vary significantly as a function of altitude. 4. DISCUSSION 4.1. The impact of common approximations The results presented in the previous section high- light how common assumptions used in current retrieval methods for exoplanets can potentially lead to wrong conclusions. Strong biases are seen for all C/O ratios, where we see that the isothermal approximation causes in general an overestimation of the absolute abundances by one or- der of magnitude, and significantly underestimates error bars. The strongest biases are seen for H2O, CO and CO2 in the C/O < 1 atmospheres, and for HCN, CH4, C2H2 for the C/O > 1 atmospheres. This is not surprising, given that these are the strongest absorbers for these C/O ranges, and therefore those with the smallest retrieved uncertainties. For all these atmospheres, excluding C/O = 1, the re- trieval method assuming a parametrized TP profile was found to describe the more complex temperature struc- ture of the atmosphere, leading to retrieved values in general agreement with the true state within 1 sigma on average. This finding opens even new prospects for the use of this technique to characterise exoplanetary at- mospheres, showing how high signal-to-noise and broad wavelength coverage transmission spectra can lead to sig- 10−1010−910−810−710−610−510−4CH4mixingratio10−410−310−210−1100Pressure(bar)TP-ISO10−1010−910−810−710−610−510−4CH4mixingratioTP-PARAMC/O=0.5C/O=0.7C/O=0.9C/O=1.0C/O=1.1C/O=1.3C/O=1.510−1010−910−810−710−610−510−4HCNmixingratio10−410−310−210−1100Pressure(bar)TP-ISO10−1010−910−810−710−610−510−4HCNmixingratioTP-PARAMC/O=0.5C/O=0.7C/O=0.9C/O=1.0C/O=1.1C/O=1.3C/O=1.510−1210−1110−1010−910−810−710−610−510−410−310−2C2H2mixingratio10−410−310−210−1100Pressure(bar)TP-ISO10−1210−1110−1010−910−810−710−610−510−410−310−2C2H2mixingratioTP-PARAMC/O=0.5C/O=0.7C/O=0.9C/O=1.0C/O=1.1C/O=1.3C/O=1.5 11 files are constant-with-altitude, which is clearly wrong for most molecules. In this case, the different features of the same molecules seen at different wavelengths (e.g. H2O and CO) probe different regions of the atmosphere, where the abundances can vary significantly. Trying to fit these features using the same abundances throughout the entire atmosphere clearly leads to strong biases. We did not explore here the possibility to fit a more complex abundance profile for the molecules, but future work in this direction will be required. The retrieved abundances obtained with the tp- param method will enable placing some limits on the C/O values of the observed atmospheres. Firstly, it will be clearly possible to differentiate between C/O greater or less than unity, and C/O = 1, as the spectra sig- natures change dramatically at this threshold. Tighter constraints on C/O can be obtained by linking the re- trieved absolute abundances with atmospheric chemical models. However, our results indicate that it will be difficult. For C/O < 1, the strongest tracer for C/O is water. Increasingly lower H2O abundances are ex- pected at increasing C/O, but the differences seen here are rather small, and comparable with the retrieved un- certainties (see Figure 6). Similarly, for C/O > 1, the strongest tracers are HCN and C2H2 (and, to a lesser extent, CH4, which has however a non uniform abun- dance profile). However, even in this case the difference in absolute abundance is quite small, and comparable with the error bars of the retrieved values. This is not totally surprising, given that the simulated transmission spectra show very little variation between similar C/O in both the oxygen- and carbon-rich regimes. Higher signal-to-noise observations might further decrease these uncertainties, and therefore improve the inferred C/O, but we note that we are already very close to the system- atic uncertainties. Other techniques might prove more effective to constrain the C/O ratio, such as emission spectra through secondary eclipse measurements and/or using chemically consistent retrieval approaches (see e.g. Greene et al. 2016). 4.2. Understanding the biases In order to understand why, and in which scenarios, a non-isothermal profile and constant-with-altitude abun- dance profiles might lead to strong biases, it is instruc- tive to look at the spectral transmittance as a function of pressure for the atmospheres under study. Figure 9 shows the spectral transmittance integrated over the path parallel to the line of sight as a function of pres- sure, together with the temperature and scale height profiles. It can be seen that different spectral regions probe different pressure ranges, and therefore different temperatures and scale heights. Firstly, we note that the scale height does not increase exponentially with al- Figure 8. Posterior distributions of CO and CO2 for C/O = 0.7 for the retrieval approach with a parametrized TP profile. Dashed lines in the histogram plots show the 1 sigma con- fidence intervals. The true state (absolute input abundance at 0.1 bar) is shown with a blue square box and straight blue lines. Note that the mixing ratios of these two molecules is approximately constant-with-altitude in this case, as seen in Figure 3. nificant constraints on the temperature profiles of the terminator region of hot Jupiter atmospheres. In general, the retrieval of constant-with-altitude mix- ing ratio profiles seems sufficient to describe the more complex real profiles when the tp-param approach is used, and is therefore a fair approximation in most cases. This is especially true for the C/O < 1 atmospheres, where the true profiles of the most abundant molecules are constant, but it is also true for the C/O > 1 at- mospheres, where one of the most abundance molecules, CH4, has a profile that varies significantly with altitude. The retrieved abundance of this molecule falls within the minimum and maximum true abundance, indicat- ing that the features seen in the transmission spectra at 3.4 and 7.6 µm probe similar pressure regions in the atmosphere. Retrieved parameters are more strongly affected for the C/O = 1 case, where the biases introduced by as- suming a constant-with-altitude abundance profile dom- inate. Small differences in the retrieved values are seen using the tp-param and tp-iso methods, and the re- trieved results are in both cases several sigma away from the true state. Interestingly, the TP profile retrieved using the tp-param method is also several sigma away from the input profile. This indicates that the biases are driven by the assumption that the abundance pro- CO2 = −7.14+0.30−0.247.57.06.56.05.5CO23.63.22.82.42.0CO3.63.22.82.42.0COCO = −3.25+0.37−0.32 12 Figure 9. The first plot on the left shows the temperature profile (blue line) and scale height profile (dashed orange line) as a function of pressure. The other plots show the spectral transmittance as a function of pressure for the models with C/O of 0.5, 1.0 and 1.5. Note that the pressure axis is the same as the first plot. The transmittance is integrated over the path parallel to the line of sight. The transmittance plots allow us to see the pressures (and therefore the temperature and scale height) probed at different wavelengths for different C/O regimes. titude between 10−3 and 1 bar, as one would expect in a purely isothermal atmosphere. On the contrary, the strong temperature gradient seen at these pressures causes the scale height to stay relatively constant at ≈ 200 km. For the atmosphere with C/O = 0.5 we see that most of the absorption occurs between 10−4 and 10−1 bar, while for the C/O = 1.1 case the transmission spec- trum probes higher-pressure regions, from 1 bar to 10−3 bar. At these pressures the temperature varies from 1500 K to about 2500 K. We also note that the peak of the absorption features probe the higher-altitude and lower-temperature part of the atmospheres, while the troughs probe the regions of the atmosphere that are almost 1000 K hotter. An isothermal approximation will clearly lead to sev- eral problems. Firstly, as we noted before, the scale height of an isothermal atmosphere will increase expo- nentially, while in this case it is roughly constant with pressure up to 1 mbar. Spectral features that probe different pressures, such as the strong water features seen for C/O < 1, will therefore vary considerably if the scale height is constant with pressure or not. A second, equally important effect, is caused by the very differ- ent temperatures probed. Molecular opacity cross sec- tions vary considerably between the temperature regions probed here (1500 K to 2500 K), and therefore assum- ing a single temperature will obviously lead to further biases. Additionally, Figure 9 helps to explain why for the retrievals of the atmospheres with C/O < 1 we found that the fit of the low-altitude temperature improves for higher C/O, while that for the high-altitude part of the TP profile improves for lower C/O. As the C/O value in- creases from 0.5 to 0.9 we see that the water abundances decreases from about 4 × 10−4 to 1 × 10−4. The effect in the transmission spectrum is a vertical shift towards lower absorption, which also translates into a vertical shift in the transmissivity plot. This means that as the water content drops, we probe increasingly higher pres- sure regions of the atmospheres, meaning that we in- creasingly lose information from the upper-altitude part of the atmosphere. This easily explains why the un- certainty on the retrieved upper-altitude temperature of these atmospheres progressively increases, while the constraint of the temperature in the bottom layers im- proves for higher C/O. So far we have only considered cloud free, broad wave- length range observations. This is the case where com- 140020002600Temperature(K)10−610−510−410−310−210−1100101Pressure(bar)100400700Scaleheight10−610−510−410−310−210−1100101Pressure(bar)TemperatureScaleheight (km)00.51TransmittanceWavelength (µm)12345678910C/O = 0.512345678910Wavelength (µm)C/O = 1.0Wavelength (µm)1234567891010−610−510−410−310−210−1100101Pressure(bar)00.51TransmittanceC/O = 1.5 mon approximations are most likely to break down. Shorter wavelength ranges will for example tend to probe specific regions of the TP profiles. For instance, an atmosphere with C/O = 1.1 observed between 1 and 3 µm will only probe pressures between 1 and 0.1 bar, where the temperature is roughly constant at ≈ 2400 K. In this scenario, we expect the isothermal approxima- tion to be sufficiently good. However, this is not always the case. If the atmosphere with C/O = 0.5 is observed between 2.5 and 4 µm, we will see a strong water fea- ture with a peak absorption coming from a region with a temperature of about 1500 K, and with wings probing increasingly higher temperatures. Clearly, even in this case an isothermal approximation would give biased re- sults, and our study indicates that the retrieved uncer- tainty of the abundance will be likely underestimated. The presence of uniform clouds will increase the de- generacy of model parameters, somewhat hiding the un- derlying biases, as the effect of a cloud deck is that of making the atmosphere opaque. A cloud deck extend- ing to 10 mbar would for example make the atmosphere opaque to incoming radiation for pressures higher than 10 mbar. This also means that it will be impossible to probe the temperature and mixing ratio profiles in this pressure regime. In the case under study, the TP profile for pressure lower than 10 mbar is relatively isothermal, and in the presence of clouds, an isothermal approxi- mation would therefore be appropriate. Note however that cloud models commonly used in current retrievals were found to cause significant degeneracies. Line & Parmentier (2016) investigated the biases of retrieving a uniform cloud cover in the presence of patchy clouds and found significant degeneracies in the retrieved mean molecular weight. We also note that similar biases are expected for cooler planets. Figure 10 shows the spectral transmittance as a function of pressure for an atmosphere with a cooler TP profile, with high altitude temperature of 1000 K. The spectrum was computed from a chemical model with C/O = 0.5, and assuming the same hot Jupiter used in this work. It can be seen that the spectrum probes the range of pressures (10−3 to 1 bar) where the TP profile changes more significantly. We therefore expect that the use of an isothermal profile to retrieve this spectrum will lead to similar biases to those found for the hotter planet case. Lastly, we note that this study focused on two spe- cific common assumptions in current retrieval methods, constant mixing ratio and temperature profiles, and the biases that these approximations can lead to. However, other strong assumptions are likely to bias our retrievals. One of the most important one is to neglect 3D dy- namical effects. The simulated observations have in fact been generated using 1D chemical models, and assume a 13 uniform chemistry and atmospheric temperature at the terminator region. Further studies that compare trans- mission spectra obtained with general circulation models and retrieved with the simpler 1D models are needed to address the biases of this assumption. A recent study in this direction is presented in Feng et al. (2016). 5. CONCLUSIONS In this paper we investigated the biases caused by two common assumptions in the forward models used by current retrieval methods of transmission spectra of hot Jupiter atmospheres: the use of an isothermal pro- file and constant-with-altitude abundances. We investi- gated whether these assumptions will still be valid for high signal-to-noise, broad wavelength coverage spectra such as those expected by JWST. In order to do this, we simulated high quality observations using a chemical scheme developed by Venot et al. (2012), which include detailed temperature and abundance profiles, and we re- trieved them using two simpler forward models: the first one assumes an isothermal profile (tp-iso), while the second one assumes a parametrized temperature pro- file (tp-param). In both cases, constant-with-altitude abundances were retrieved. We found that: • The non-uniform temperature profile could be well retrieved within about 1 sigma for all cases but C/O = 1 using the tp-param method. This is an important result, opening the possibility to obtain detailed temperature structure information about the terminator region of a hot Jupiter. • The retrieval approach that assumes an isother- mal profile led to strong biases. We found that, on average, the retrieved abundances using this method are overestimated by about one order of magnitude and the error bars are underestimated. The tp-param approach leads to much improved constraints, with retrieved abundances within 1 -- 2 sigma of the input values in most cases. This is also because the retrieved uncertainties are gener- ally larger. • The retrieval assumption that abundance profiles are constant-with-altitude was found to be a good approximation for C/O < 1 and C/O > 1 atmo- spheres, but not for C/O = 1. In this latter case, most of the abundance profiles have strong varia- tions, and a uniform abundance profile is a poor approximation that leads to significant biases. Fu- ture work will therefore be needed to address the feasibility of fitting more complex abundance pro- files. • Although we found that differentiating between C/O < 1, C/O = 1 and C/O > 1 was straight 14 Figure 10. Temperature pressure profile and spectral transmittance for a planet with a cooler TP profile. Caption as in Figure 9. forward, we also found that tighter constraints are more difficult to obtain as the differences between the transmission spectra are relatively small. Higher signal-to-noise observations might lead to better constraints, but other biases, due to systematic uncertainties for example, might be- come more dominant. Emission spectra observa- tions, possibly combined with transmission spec- tra, might give better constraints than transmis- sion spectra alone. These results show that when broad wavelength ranges and high signal-to-noise observations are used, the forward models used in our retrieval approaches need to allow for larger flexibility. One very simple solution is to adopt a parametrization of the temperature-pressure profile, as the one used here, but other techniques, such as the two-stage approach used in Waldmann et al. (2015a), could be considered in the future. We thank the referee for providing useful comments. This work was supported by STFC (ST/K502406/1) and the ERC projects ExoLights (617119) and Exo- Mol (267219). O.V. acknowledges support from the KU Leuven IDO project IDO/10/2013 and from the FWO Postdoctoral Fellowship programme. POL ac- knowledges support from the LabEx P2IO, the French ANR contract 05-BLAN-NT09-573739. REFERENCES Ag´undez, M., Venot, O., Selsis, F., & Iro, N. 2014, ApJ, 781, 68 Ali-Dib, M., Mousis, O., Petit, J.-M., & Lunine, J. I. 2014, ApJ, -- . 2013, ApJ, 778, 153 Bounaceur, R., Glaude, P. A., Fournet, R., et al. 2007, 785, 125 International Journal of Vehicle Design Anderlohr, J. M., Pires da Cruz, A., Bounaceur, R., & Brogi, M., de Kok, R. J., Birkby, J. L., Schwarz, H., & Snellen, I. Battin-Leclerc, F. 2010, Combustion Science and Technology, 182, 39 Barber, R. J., Strange, J. K., Hill, C., et al. 2014, MNRAS, 437, 1828 A. G. 2014, A&A, 565, A124 Brown, T. M. 2001, ApJL, 553, 1006 Cavali´e, T., Moreno, R., Lellouch, E., et al. 2014, A&A, 562, A33 Charbonneau, D., Brown, T. M., Noyes, R. W., & Gilliland, Barber, R. J., Tennyson, J., Harris, G. J., & Tolchenov, R. N. R. L. 2002, ApJL, 568, 377 2006, Monthly Notices of the Royal Astronomical Society, 368, 1087 Charbonneau, D., Knutson, H. A., Barman, T. S., et al. 2008, ApJ, 686, 1341 Barstow, J. K., Aigrain, S., Irwin, P. G. J., et al. 2013, Monthly Cowan, N. B., Greene, T., Angerhausen, D., et al. 2015, Notices of the Royal Astronomical Society, 430, 1188 Publications of the Astronomical Society of Pacific, 127, 311 Barstow, J. K., Aigrain, S., Irwin, P. G. J., Kendrew, S., & Fletcher, L. N. 2015, Monthly Notices of the Royal Astronomical Society, 448, 2546 Barstow, J. K., & Irwin, P. G. J. 2016, arXiv, arXiv:1605.07352 Batalha, N., Kalirai, J., Lunine, J., Clampin, M., & Lindler, D. 2015, arXiv, arXiv:1507.02655 Battin-Leclerc, F., Bounaceur, R., Glaude, P.-A., & Belmekki, N. 2006, arXiv, arXiv:physics Bean, J. L., Desert, J.-M., Seifahrt, A., et al. 2013, ApJ, 771, 108 Beaulieu, J. P., Kipping, D., Batista, V., et al. 2010, Monthly Notices of the Royal Astronomical Society, 409, 963 Beichman, C., Benneke, B., Knutson, H. A., et al. 2014, Publications of the Astronomical Society of Pacific, 126, 1134 Benneke, B., & Seager, S. 2012, ApJL, 753, 100 Danielski, C., Deroo, P., Waldmann, I., et al. 2014, ApJ, 785, 35 de Wit, J., & Seager, S. 2013, Science, 342, 1473 Deming, D., Knutson, H. A., Agol, E., et al. 2011, ApJ, 726, 95 Doyon, R., Hutchings, J. B., Beaulieu, M., et al. 2012, in SPIE Astronomical Telescopes + Instrumentation, ed. M. C. Clampin, G. G. Fazio, H. A. MacEwen, & J. M. Oschmann (SPIE), 84422R -- 13 Feng, Y. K., Line, M. R., Fortney, J., et al. 2016, arXiv, arXiv:1607.03230 Feroz, F., & Hobson, M. P. 2008, MNRAS, 384, 449 Ferruit, P., Birkmann, S., Boker, T., et al. 2014, in Proceedings of the SPIE, ed. J. M. Oschmann, M. Clampin, G. G. Fazio, & H. A. MacEwen, European Space Research and Technology Ctr. (Netherlands) (SPIE), 91430A 500100015002000Temperature(K)10−610−510−410−310−210−1100101Pressure(bar)Temperature12345678910Wavelength(µm)10−610−510−410−310−210−1100101Pressure(bar)00.51Transmittance 15 Greene, T., Beichman, C., Eisenstein, D., et al. 2007, Techniques Redfield, S., Endl, M., Cochran, W. D., & Koesterke, L. 2008, and Instrumentation for Detection of Exoplanets III. Edited by Coulter, 6693, 66930G Greene, T. P., Line, M. R., Montero, C., et al. 2016, ApJ, 817, 17 Grillmair, C. J., Charbonneau, D., Burrows, A., et al. 2007, ApJ, 658, L115 Guillot, T. 2010, A&A, 520, A27 Harris, G. J., Tennyson, J., Kaminsky, B. M., Pavlenko, Y. V., & Jones, H. R. A. 2006, Monthly Notices of the Royal Astronomical Society, 367, 400 Hollis, M. D. J., Tessenyi, M., & Tinetti, G. 2013, Computer Physics Communications, 184, 2351 Husser, T. O., Wende-von Berg, S., Dreizler, S., et al. 2013, A&A, 553, 6 Irwin, P. G. J., Teanby, N. A., De Kok, R., et al. 2008, Journal of Quantitative Spectroscopy & Radiative Transfer, 109, 1136 Kendrew, S., Scheithauer, S., Bouchet, P., et al. 2015, Publications of the Astronomical Society of Pacific, 127, 623 Knutson, H. A., Charbonneau, D., Allen, L. E., et al. 2007, ApJL, 447, 183 The Astrophysical Journal Letters, 673, L87 Richard, C., Gordon, I. E., Rothman, L. S., et al. 2012, Journal of Quantitative Spectroscopy & Radiative Transfer, 113, 1276 Rothman, L. S., Gordon, I. E., Barber, R. J., et al. 2010, Journal of Quantitative Spectroscopy & Radiative Transfer, 111, 2139 Rothman, L. S., Gordon, I. E., Babikov, Y., et al. 2013, Journal of Quantitative Spectroscopy & Radiative Transfer, 130, 4 Seager, S. 2011, Contemporary Physics, 52, 602 Sing, D. K., Fortney, J., Nikolov, N., et al. 2016, Nature, 529, 59 Snellen, I. A. G., Albrecht, S., de Mooij, E. J. W., & Le Poole, R. S. 2008, A&A, 487, 357 Stevenson, K. B., Harrington, J., Nymeyer, S., et al. 2010, Nature, 464, 1161 Swain, M., Deroo, P., Griffith, C. A., et al. 2010, Nature, 463, 637 Swain, M., Deroo, P., Tinetti, G., et al. 2013, Icarus, 225, 432 Tennyson, J., & Yurchenko, S. N. 2012, Monthly Notices of the Royal Astronomical Society, 425, 21 Thiabaud, A., Marboeuf, U., Alibert, Y., Leya, I., & Mezger, K. 2015, A&A, 580, A30 Tinetti, G., Deroo, P., Swain, M., et al. 2010, The Astrophysical Knutson, H. A., Madhusudhan, N., Cowan, N. B., et al. 2011, Journal Letters, 712, L139 ApJ, 735, 27 Kreidberg, L., Bean, J. L., Desert, J.-M., et al. 2014, Nature, 505, 69 Tinetti, G., Tennyson, J., Griffith, C. A., & Waldmann, I. 2012, Philosophical Transactions of the Royal Society A: Mathematical, 370, 2749 Lee, J. M., Fletcher, L. N., & Irwin, P. G. J. 2011, MNRAS, 420, Tinetti, G., Vidal-Madjar, A., Liang, M.-C., et al. 2007, Nature, 170 448, 169 Lewis, N. K., Showman, A. P., Fortney, J., et al. 2010, ApJ, 720, Todorov, K. O., Deming, D., Knutson, H. A., et al. 2013, ApJ, 344 Line, M. R., Fortney, J., Marley, M. S., & Sorahana, S. 2014, ApJ, 793, 33 Line, M. R., & Parmentier, V. 2016, ApJ, 820, 78 Line, M. R., Vasisht, G., Chen, P., Angerhausen, D., & Yung, Y. L. 2011, ApJ, 738, 32 Line, M. R., Zhang, X., Vasisht, G., et al. 2012, ApJ, 749, 93 Line, M. R., Wolf, A. S., Zhang, X., et al. 2013, ApJ, 775, 137 Liou, K. N. 2002, An introduction to atmospheric radiation, Academic Press Madhusudhan, N. 2012, ApJ, 758, 36 Madhusudhan, N., Burrows, A., & Currie, T. 2011a, ApJ, 737, 34 Madhusudhan, N., & Seager, S. 2009, ApJ, 707, 24 Madhusudhan, N., Harrington, J., Stevenson, K. B., et al. 2011b, Nature, 469, 64 770, 102 Tsiaras, A., Waldmann, I., Rocchetto, M., et al. 2015, arXiv, arXiv:1511.07796 Tsiaras, A., Rocchetto, M., Waldmann, I., et al. 2016, ApJ, 820, 99 Valencia, D., Guillot, T., Parmentier, V., & Freedman, R. S. 2013, ApJ, 775, 10 Venot, O., & Ag´undez, M. 2015, Experimental Astronomy, 40, 469 Venot, O., Ag´undez, M., Selsis, F., Tessenyi, M., & Iro, N. 2014, A&A, 562, 51 Venot, O., H´ebrard, E., Ag´undez, M., Decin, L., & Bounaceur, R. 2015, A&A, 577, A33 Venot, O., H´ebrard, E., Ag´undez, M., et al. 2012, A&A, 546, A43 Venot, O., Fray, N., B´enilan, Y., et al. 2013, A&A, 551, A131 Waldmann, I., Rocchetto, M., Tinetti, G., et al. 2015a, ApJ, 813, Moses, J. I., Madhusudhan, N., Visscher, C., & Freedman, R. S. 13 2013a, ApJ, 763, 25 Moses, J. I., Visscher, C., Fortney, J., et al. 2011, ApJL, 737, 15 Moses, J. I., Line, M. R., Visscher, C., et al. 2013b, ApJ, 777, 34 Mousis, O., Fletcher, L. N., Lebreton, J. P., et al. 2014, Planetary and Space Science, 104, 29 Oberg, K. I., Murray-Clay, R., & Bergin, E. A. 2011, The Astrophysical Journal Letters, 743, L16 Parmentier, V., & Guillot, T. 2014, A&A, 562, A133 Parmentier, V., Guillot, T., Fortney, J., & Marley, M. S. 2015, A&A, 574, A35 Parmentier, V., Showman, A. P., & Lian, Y. 2013, A&A, 558, A91 Waldmann, I., Tinetti, G., Drossart, P., et al. 2012, ApJL, 744, 35 Waldmann, I., Tinetti, G., Rocchetto, M., et al. 2015b, ApJ, 802, 107 Wang, H., Warner, S. J., Oehlschlaeger, M. A., et al. 2010, Combustion and Flame, 157, 1976 Yurchenko, S. N., Barber, R. J., & Tennyson, J. 2011, MNRAS, 413, 1828 Yurchenko, S. N., Tennyson, J., Barber, R. J., & Thiel, W. 2013, Journal of Molecular Spectroscopy, 291, 69 Zellem, R. T., Griffith, C. A., Deroo, P., Swain, M., & Waldmann, I. 2014, ApJ, 796, 48 APPENDIX 16 Table 1. Retrieved absolute abundances with 1 sigma uncertainty for the seven molecules and seven C/O values considered in this study. For each retrieved parameter, we show in parenthesis how many sigma away the retrieved value is from the true state. Parameter C/O Input value Retrieved value H2O CO CO2 CH4 HCN C2H2 (at 0.1 bar) 4.08 × 10−4 2.52 × 10−4 8.63 × 10−5 1.20 × 10−6 4.56 × 10−8 2.44 × 10−8 1.85 × 10−8 4.13 × 10−4 5.70 × 10−4 7.34 × 10−4 8.22 × 10−4 8.24 × 10−4 8.24 × 10−4 8.24 × 10−4 3.60 × 10−8 3.07 × 10−8 1.35 × 10−8 2.10 × 10−10 8.01 × 10−12 4.28 × 10−12 3.25 × 10−12 6.43 × 10−11 1.44 × 10−10 5.39 × 10−10 4.36 × 10−8 1.16 × 10−6 2.20 × 10−6 2.92 × 10−6 1.32 × 10−9 2.94 × 10−9 1.10 × 10−8 8.84 × 10−7 2.09 × 10−5 3.58 × 10−5 4.42 × 10−5 8.54 × 10−14 4.26 × 10−13 5.96 × 10−12 3.88 × 10−8 2.72 × 10−5 9.66 × 10−5 tp-iso 4.66 × 10−3 − 6.53 × 10−3 (15.4) 2.98 × 10−3 − 4.12 × 10−3 (16.2) 6.61 × 10−4 − 8.18 × 10−4 (20.1) 1.63 × 10−6 − 1.94 × 10−6 (4.6) < 1.04 × 10−8 (1.4) < 5.13 × 10−9 (1.5) < 5.74 × 10−9 (1.3) 8.74 × 10−3 − 1.10 × 10−2 (28.0) 9.46 × 10−3 − 1.17 × 10−2 (27.0) 1.15 × 10−2 − 1.44 × 10−2 (25.4) 2.24 × 10−3 − 3.38 × 10−3 (5.9) 7.39 × 10−3 − 9.62 × 10−3 (17.6) 8.35 × 10−3 − 1.06 × 10−2 (20.7) 8.17 × 10−3 − 1.03 × 10−2 (20.9) 9.46 × 10−7 − 1.82 × 10−6 (11.0) 5.95 × 10−7 − 1.14 × 10−6 (10.2) 1.57 × 10−7 − 2.77 × 10−7 (9.6) < 2.78 × 10−9 (1.1) < 3.32 × 10−10 (0.6) < 7.23 × 10−10 (0.9) < 4.05 × 10−10 (1.0) < 1.85 × 10−7 (0.6) < 2.86 × 10−7 (0.5) < 2.03 × 10−8 (0.1) 8.48 × 10−8 − 1.13 × 10−7 (5.6) 9.23 × 10−6 − 1.15 × 10−5 (20.0) 2.03 × 10−5 − 2.55 × 10−5 (20.4) 1.80 × 10−5 − 2.22 × 10−5 (18.6) < 2.15 × 10−5 (0.4) < 2.94 × 10−7 (0.1) < 5.54 × 10−7 (0.3) 6.60 × 10−7 − 8.20 × 10−7 (1.7) 5.53 × 10−5 − 6.74 × 10−5 (10.8) 1.20 × 10−4 − 1.49 × 10−4 (12.1) 1.01 × 10−4 − 1.23 × 10−4 (9.2) < 4.18 × 10−7 (2.0) < 5.26 × 10−7 (1.6) < 2.97 × 10−7 (1.1) < 4.08 × 10−9 (1.6) 1.66 × 10−4 − 2.88 × 10−4 (7.5) 7.19 × 10−4 − 1.17 × 10−3 (9.2) tp-param 3.73 × 10−4 − 1.98 × 10−2 (1.0) 1.82 × 10−4 − 5.37 × 10−4 (0.4) 4.84 × 10−5 − 9.75 × 10−5 (0.3) 1.43 × 10−6 − 1.62 × 10−6 (4.0) < 5.44 × 10−12 (6.2) < 2.58 × 10−8 (1.1) < 1.49 × 10−8 (0.8) 3.64 × 10−4 − 5.33 × 10−2 (0.9) 2.69 × 10−4 − 1.17 × 10−3 (0.0) 2.90 × 10−4 − 7.79 × 10−4 (0.4) 2.44 × 10−3 − 3.40 × 10−3 (7.5) 1.01 × 10−3 − 1.01 × 10−3 (799.1) 6.33 × 10−4 − 1.67 × 10−3 (0.5) 8.91 × 10−4 − 1.26 × 10−3 (1.4) 8.63 × 10−8 − 4.13 × 10−6 (1.5) 4.13 × 10−8 − 1.27 × 10−7 (1.5) 1.41 × 10−8 − 2.87 × 10−8 (1.1) < 3.24 × 10−9 (0.3) < 7.64 × 10−10 (2.5) < 8.56 × 10−12 (0.4) < 4.51 × 10−7 (1.3) < 1.30 × 10−7 (0.7) < 1.97 × 10−8 (0.3) < 7.50 × 10−10 (0.5) 4.61 × 10−8 − 6.35 × 10−8 (1.4) 3.06 × 10−6 − 3.06 × 10−6 (43189.6) 4.85 × 10−6 − 8.79 × 10−6 (3.6) 5.66 × 10−6 − 6.57 × 10−6 (10.0) < 1.15 × 10−7 (0.0) < 2.68 × 10−8 (0.4) < 3.03 × 10−8 (0.9) 5.18 × 10−7 − 6.14 × 10−7 (5.3) 2.18 × 10−5 − 2.18 × 10−5 (1355.8) 3.00 × 10−5 − 5.79 × 10−5 (0.5) 3.47 × 10−5 − 4.26 × 10−5 (0.3) < 9.43 × 10−8 (2.2) < 1.03 × 10−7 (1.8) < 5.24 × 10−7 (1.3) < 4.77 × 10−9 (1.6) 2.51 × 10−5 − 2.51 × 10−5 (0.1) 5.76 × 10−5 − 4.44 × 10−4 (0.5) 0.5 0.7 0.9 1.0 1.1 1.3 1.5 0.5 0.7 0.9 1.0 1.1 1.3 1.5 0.5 0.7 0.9 1.0 1.1 1.3 1.5 0.5 0.7 0.9 1.0 1.1 1.3 1.5 0.5 0.7 0.9 1.0 1.1 1.3 1.5 0.5 0.7 0.9 1.0 1.1 1.3 17 Table 1. Retrieved absolute abundances (continued). Parameter C/O Input value Retrieved value NH3 (at 0.1 bar) 1.68 × 10−4 9.73 × 10−9 9.73 × 10−9 9.65 × 10−9 9.62 × 10−9 8.70 × 10−9 7.97 × 10−9 7.51 × 10−9 tp-iso 9.42 × 10−4 − 1.47 × 10−3 (8.7) < 1.04 × 10−7 (0.5) < 4.43 × 10−7 (0.3) < 3.53 × 10−6 (10.9) < 3.76 × 10−7 (1.4) < 1.80 × 10−6 (36.4) < 2.02 × 10−6 (29.5) < 2.15 × 10−6 (35.7) 1.5 0.5 0.7 0.9 1.0 1.1 1.3 1.5 tp-param 1.64 × 10−4 − 2.86 × 10−4 (0.9) 9.74 × 10−12 − 3.33 × 10−8 (0.7) < 1.11 × 10−8 (0.9) < 9.63 × 10−8 (0.5) < 9.12 × 10−8 (3.2) < 2.89 × 10−7 (4651.3) < 5.93 × 10−7 (5.6) < 1.76 × 10−7 (1452.3)
1211.5899
1
1211
2012-11-26T09:38:44
Origin of anomalous Xe-H in presolar diamonds: Indications of a "cold" r-process
[ "astro-ph.EP", "astro-ph.SR" ]
We report on a concerted effort aimed at understanding the nucleosynthesis origin of Xe-H in presolar nanodiamonds. Previously explored possible explanations have included a secondary neutron-burst process occurring in the He-shell of a type II supernova (SN), as well as a rapid separation, between unstable precursor isobars of a primary r-process, and stable Xe isotopes. Here we present results from the investigation of a rapid neutron-capture scenario in core-collapse SNe with different non-standard r-process variants. Our calculations are performed in the framework of the high-entropy-wind (HEW) scenario using updated nuclear-physics input. We explore the consequences of varying the wind expansion velocity (Vexp) for selected electron fractions (Ye) with their correlated entropy ranges (S), and neutron-freezeout temperatures (T9(freeze)) and timescales (tau-r(freeze). We draw several conclusions: For Xe-H a "cold" r-process with a fast freezeout seems to be the favored scenario. Furthermore, eliminating the low-S range (i.e. the "weak" r-process component) and maintaining a pure "main" or even "strong" r-process leads to an optimum overall agreement with the measured iXe/136Xe abundance ratios. Our results can provide valuable additional insight into overall astrophysical conditions of producing the r-process part of the total SS heavy elements in explosive nucleosynthesis scenarios.
astro-ph.EP
astro-ph
Origin of Anomalous Xe-H in Presolar Diamonds: Indications of a "cold" r-process Karl-Ludwig Kratz Max-Planck-Institut für Chemie Hahn-Meitner-Weg 1, D-55128 Mainz, Germany and: Department of Physics, University of Notre Dame, USA E-mail: [email protected] Khalil Farouqi Max-Planck-Institut für Chemie Hahn-Meitner-Weg 1, D-55128 Mainz, Germany E-mail: [email protected] Ulrich Ott1 Max-Planck-Institut für Chemie Hahn-Meitner-Weg 1, D-55128 Mainz, Germany; and: University of Western Hungary, Szombathely, Hungary E-mail: [email protected] We report on a concerted effort aimed at understanding the nucleosynthesis origin of Xe-H in presolar nanodiamonds. Previously explored possible explanations have included a secondary neutron-burst process occurring in the He-shell of a type II supernova (SN), as well as a rapid separation, between unstable precursor isobars of a primary r-process, and stable Xe isotopes. Here we present results from the investigation of a rapid neutron-capture scenario in core- collapse SNe with different non-standard r-process variants. Our calculations are performed in the framework of the high-entropy-wind (HEW) scenario using updated nuclear-physics input. We explore the consequences of varying the wind expansion velocity (Vexp) for selected electron fractions (Ye) with their correlated entropy ranges (S), and neutron-freezeout temperatures (T9(freeze)) and timescales (r(freeze). We draw several conclusions: For Xe-H a "cold" r- process with a fast freezeout seems to be the favored scenario. Furthermore, eliminating the low-S range (i.e. the "weak" r-process component) and maintaining a pure "main" or even "strong" r-process leads to an optimum overall agreement with the measured iXe/136Xe abundance ratios. Our results can provide valuable additional insight into overall astrophysical conditions of producing the r-process part of the total SS heavy elements in explosive nucleosynthesis scenarios. XII International Symposium on Nuclei in the Cosmos August 5-12, 2012 Cairns, Australia  Copyright owned by the author(s) under the terms of the Creative Commons Attribution-NonCommercial-ShareAlike Licence. http://pos.sissa.it Xe-H in presolar diamonds Ulrich Ott 1. Introduction Apart from the historical "bulk" solar system (SS) isotopic abundances (N⊙) [1, 2] and the elemental abundances more recently measured for metal-deficient halo stars (for a review, see e.g. [3]), meteoritic grains of stardust, which survived from times before the SS formed (see, such as SiC and oxide grains. A major problem is the small size (average ≃2.6 nm), which does e.g. [4-6] represent a third group of "observables" crucial to our understanding of the various nucleosynthesis processes in stars. Within the meteoritic grains, attempts to explain the origin of the nanodiamonds has not progressed as much as the understanding of other types of stardust, not permit classical single-grain isotopic measurements. Therefore, "bulk" samples (i.e. millions of tiny nanodiamond grains) have to be analyzed. Diagnostic isotopic features are present in several trace elements and suggest a connection to SNe. These include Xe-HL [7] (with enhancements in p- and r-isotopes relative to SS), Kr-H [7] and Pt-H [8, 9] (where the heavy isotopes are enhanced), and Te-H [10] (with a clear enhancement of the r-only isotopes). So far, nucleosynthesis processes suggested to account for the H (heavy, r-process) enhancements include "neutron burst" scenarios of secondary nature [11-13], as well as a regular primary r- process augmented by an "early" separation between final, stable isotopes and radioactive isobars formed during the -decay back to stability from the initial precursors in the assumed r- process [14]. Since both earlier scenarios are not fully convincing in terms of modern nucleosynthesis conditions, we have initiated a concerted effort to look for isotopic features that may be diagnostic for a standard r-process in high-entropy-wind (HEW) ejecta of core-collapse SNe (for details of our HEW-model, see e.g. [15]. 2. r-Process Calculations and Xe-H Production mechanisms for elements beyond Fe by slow (s-process) and rapid (r-process) captures of neutrons, have been known for a long time [2]. However, the search for a robust r- process production site has proven difficult. All proposed scenarios face problems with astrophysical conditions as well as with the necessary nuclear-physics input for isotopes far from β-stability. Among the various models, the SN neutrino-driven or high-entropy wind (HEW) model is one of the best studied mechanisms. However, also for this attractive scenario until recently even in the most sophisticated hydrodynamical models the neutrino-driven wind is proton-rich (electron fraction Ye>0.5) during its entire life, thus precluding r-process nucleosynthesis (see, e.g. [16]). Instead of assuming uncommon conditions to obtain at least low-entropy neutron-rich ejecta (Ye<0.5) and / or artificially increasing the wind entropy by factors of several (see, e.g. [17]), we have steadily optimized our SN HEW parameter studies with simplified dynamics using the computational method(s) and basic nuclear-physics input as outlined in [15]. In fact, this seems to be justified by new investigations on charged-current neutrino interaction rates, 2 Xe-H in presolar diamonds Ulrich Ott which may significantly alter the earlier proton-rich Ye towards more neutron-rich conditions (see, e.g. [18]). Nr,⊙≃N⊙-Ns,⊙. Such comparisons of isotopic abundance patterns require best possible Commonly, r-process nucleosynthesis calculations are performed to reproduce the SS r- process pattern (see, e.g., [1, 15, 19]), historically defined as the abundance differences between the total SS abundances [2] and the assumed s-process contribution for SS metallicity [20] -- boulevard (see, e.g, [19, 21-24]), in particular for nuclei involved in the creation of the A ≃ 130 accuracy of both astrophysical and nuclear-physics parameter input. Given the progress during the past three decades on modeling and measuring nuclear properties near and in the r-process Nr,⊙ peak, we have performed an exploratory study of possible modifications of the "standard" r-process (for SS abundances) that can reproduce the observed peculiar isotopic pattern of Xe- H in presolar nanodiamonds [7, 9, 14]. In these stardust samples, the Xe-H abundance pattern is given as ratios relative to the assumed "r-only" isotope 136Xe, and have (somewhat model- different from the Nr,⊙ abundance ratios of 3.37, 2.55, 2.24, 1.18 [20]. dependent) values of 129,131,132,134Xe/136 Xe = 0.207, 0.178, 0.167, 0.699 [9, 14], which are clearly As for the nuclear-physics input used in our calculations, all theoretical properties are Compared to [15], two additional local improvements in the A ≃ 130 peak region [24] were consistently based on the deformed, quenched mass model ETFSI-Q [22]. Using its Qβ and Sn values as well as the predicted ground-state shape, β-decay half-lives (T1/2) and β-delayed neutron branching ratios (Pxn) were calculated within the QRPA approach as outlined in [21]. added. First is the inclusion of the T1/2 of known and theoretical πp1/2 isomers in addition to the πg9/2 ground-state decays in the N = 82 odd-proton nuclei from ¹³¹In to ¹²³Tc. The second "bulk" Nr,⊙ isotopic abundances in the A ≃ 130 peak obtained by these updates is shown, for corresponds to new QRPA(GT+ff) calculations for the mass region of exotic 83 < N < 87 r- process isotopes between Mo and Cd, which are based on optimized model parameters to reproduce experimental level structures in 130In and ¹³¹In. The improvement in reproducing the example, in Fig. 5 of [24]. It is worth emphasizing at this point that a satisfactory agreement of measured abundance patterns deviating from the Nr,⊙ distribution. between observations and calculations for bulk SS abundances is essential for the interpretation neutrons to seed nuclei is given as Yn/Yseed ≃ kSN  Vexp  (S/Ye)³) with an electron fraction of All results shown in the following were obtained using our parameterized SN HEW model, as outlined in [15]. We have started our study using the correlated "standard" astrophysics parameters (according to our "r-process strength function", where the ratio of free freezeout temperature for the free neutrons of T₉(freeze) ≃ 0.82, which is reached at a time τr(freeze) ≃ 138 ms after the r-process seed formation. This r-process variant reaches the Ye=0.45, an expansion velocity of the ejecta of Vexp = 7500 km/s and an entropy range of maximum abundance at the top of the A ≃ 130 r-abundance peak at S≃195 for Yn/Yseed ≃ 35 20<S<280 kB/baryon. Under these conditions, we have a "hybrid" r-process type with a as representative, is 2.77, which is close to the Nr,⊙ ratio of 3.37. This result is not sensitive to with the peak abundance at A = 128. The corresponding 129Xe/136Xe abundance ratio, taken here the assumed Ye value (i.e. the neutron-richness of the wind ejecta). We found that, compared to 3 Xe-H in presolar diamonds Ulrich Ott the above abundance ratio for Ye=0.45, this value only increases by about 12 % by going down to Ye=0.40 or up to Ye=0.48, respectively. Given this, for the further calculations we have kept the electron fraction constant at Ye=0.45 and studied the influence of the expansion velocity in the range 1000<Vexp<30,000 freezeout time of τr(freeze) ≃ 560 ms and a freezeout temperature of T₉(freeze) ≃ 1.3. The top km/s, together with the other parameters correlated according to the above defined "strength of the second r-process peak lies at A = 130, for a high S ≃370 kB/baryon. In contrast, for a high function", while still keeping the full entropy ranges. For a low expansion velocity of the ejecta Vexp of 30,000 km/s we obtain a fast, "cold" r-process variant with τr(freeze) ≃ 55 ms, of Vexp = 1000 km/s we obtain a so-called "hot" r-process variant with a relatively long neutron T₉(freeze) ≃ 0.4. The top of the second r-process peak is at A = 127 for a low S of ≃120 kB/baryon. As can be seen from Fig. 1, which shows the results for a number of selected expansion velocities of the wind ejecta (2000<Vexp<18,000 km/s), none of these parameter combinations with the respective full entropy ranges is able to reproduce in a satisfactory manner all iXe/136Xe ratios measured in the presolar nanodiamond samples. However, it is apparent that the "best" overall agreement with the observations will very likely be obtained under "cold" r-process conditions. Therefore, in the last step of our parameter study we focused on the parameter combinations of such "non-standard" 10,000<Vexp<30,000 km/s r-process variants, and studied the effect of choosing different entropy ranges. 6 5 4 3 2 1 0 132 132 mass mass Figure 1: HEW predictions of iXe/136Xe Figure 2: HEW predictions of iXe/136Xe- abundance ratios for neutron-rich r-process abundance ratios for constant Ye = 0.45 and ejecta with an electron fraction Ye = 0.45 as different expansion velocities Vexp [km/s] a function of expansion velocity Vexp (as with the respective "best-fit" entropy ranges full entropy ranges. The Nr⊙ residuals and For comparison, again Nr⊙ residuals and given in the legend in km/s units) for the Delta-S [kB/baryon], as given in the legend. Xe-H data are included for comparison. Xe-H data included. Xe-H nuclide 129Xe is formed most effectively at S≃185 kB/baryon, at a slightly lower entropy As we pointed out earlier already [9, 25], for given Ye and Vexp, the different Xe isotopes are produced in different entropy ranges. For our standard conditions used here, the "lightest" than required to form the Nr,⊙ A=130 top of the peak. The heavier isotopes follow at higher S S≃220 kB/baryon, where already the REE pygmy-peak region and part of the third Nr,⊙ peak are values "beyond" the peak-top conditions. Finally, 136Xe has its most effective production at 4 r-solar vexp = 2000 vexp = 6000 vexp = 10,000 vexp = 14,000 vexp = 18,000 Xe-H r-solar 3000, 276-358 7500, 208-260 14000, 180-216 30000, 142-174 Xe-H e X 6 3 1 / e X i e X 6 3 1 / e X i 3 2 1 0 134 136 130 130 134 136 Xe-H in presolar diamonds Ulrich Ott formed (see, e.g. Fig. 8 in [15]). This suggests that the "low" iXe/136Xe ratios observed in the nanodiamonds can be reproduced when limiting the HEW ejecta to a "main" or even "strong" r- process component with “high" entropies; i.e. in principle by cutting out from the entropy parameter space the lower-S "weak" r-process component. Thus, for our "standard hybrid" Ye=0.45, Vexp = 7500 km/s conditions, a cut of the low-S "weak" r-process component (S<185 kB/baryon) leads to a considerable reduction of the 129Xe with an entropy cut at S≃208 kB/baryon, with abundance ratios 129,131,132,134Xe/136Xe = 0.370, abundance but has only a small effect on 136Xe. This results in a decrease of the original, full-S range abundance ratio of 129Xe/136Xe = 2.77 by a factor 3 to 0.923. Under these HEW conditions, the "best" possible fit to the Xe-H pattern is obtained for a "strong" r-process variant 0.347, 0.561, 0.608, respectively. Although 129Xe, ¹³¹Xe and ¹³²Xe are still too high by factors 1.8, 1.9 and 3.4, respectively, 134Xe is met within 15 % to the observed ratio. As expected, even better agreement with the Xe-H pattern is obtained for "cold" r-process variants in the range 10,000<Vexp<20,000 km/s, with corresponding entropy cuts of S>206 to S>162 kB/baryon, respectively. A selection of "best fit" examples is shown in Fig. 2, which shows that a "strong cold" r-process can indeed reproduce the Xe-H abundance ratios of 129,131,134Xe/136Xe within 20 to 30 %. Only ¹³²Xe/136Xe, with a "best" ratio of 0.436, remains too extrapolations, the region beyond A ≃ 133 is rather uncertain, not so much in terms of T1/2 high (by a factor ~2.6). In this special case, the remaining deviation may be due to a non- optimum nuclear-physics input. Whereas for the lighter stable Xe isotopes most of the nuclear parameters are either experimentally available or are reasonably well understood by short-range values, but rather in the individual branching ratios of β-delayed neutrons. Their measurements nanodiamonds. With Xe-H lying in the A ≃ 130 Nr,⊙ peak and Pt-H in the A ≃ 195 peak, these clearly will be a challenge for future experiments. Interestingly, preliminary results for platinum indicate that the HEW conditions found to be favorable for Xe-H can also account for the tentatively reported Pt-H in presolar cosmochemical samples with isotopic abundance patterns can provide constraints on the astrophysical conditions for the production of a full r-process, which cannot be deduced with this sensitivity by the elemental abundance patterns of metal-poor halo stars. References [1] E.M. Burbidge et al. ("B²FH"), Synthesis of the elements in stars, Rev. Mod. Phys. 29 (1956) 547. [2] K. Lodders, H. Palme, H.-P. Gail, Solar System elemental abundances in 2009, Landolt- Börnstein, New Series VI/4B-34 (2009) 1. [3] C. Sneden, J.J. Cowan, R. Gallino, Neutron capture elements in the early Galaxy, ARA&A 46 (2008) 241. [4] D.D. Clayton, L.R. Nittler, Astrophysics with presolar stardust, ARA&A 42 (2004) 39. [5] K. Lodders, S. Amari, Presolar grains from meteorites: Remnants from the early times of the solar system, ChEG 65 (2005) 93. 5 Xe-H in presolar diamonds Ulrich Ott [6] E. Zinner, Presolar grains, Update, in: Treatise on Geochemistry, Oxford: Elsevier Ltd., 1.02 (online update only), 1. [7] G.R. Huss, R.S. Lewis, Noble gases in presolar diamonds I: Three distinct components and their implications for diamond origins, Meteoritics 29 (1994) 791. [8] U. Ott et al., Platinum-H in presolar nanodiamond, Meteoritics 45 (2010) A159. [9] U. Ott et al., New attempts to understand nanodiamond stardust, PASA 29 (2012) 90. [10] S. Richter, U. Ott, F. Begemann, Tellurium in pre-solar diamonds as an indicator for rapid separation of supernova ejecta, Nature 391 (1998) 261. [11] D.D. Clayton, Origin of heay xenon in meteoritic diamonds, Ap. J. 340 (1989) 613. [12] B.S. Meyer, D.D. Clayton, L.-S. The, Molybdenum and zirconium isotopes from a supernova neutron burst, Ap. J. 540 (2000) L49. [13] T. Rauscher et al., Nucleosynthesis in massive stars with improved nuclear and stellar physics, Ap. J. 576 (2002) 323. [14] U. Ott, Interstellar diamond xenon and timescales of supernova ejecta, Ap. J. 463 (1996) 344. [15] K. Farouqi et al., Charged-particle and neutron-capture processes in the high-entropy wind of core-collapse supernovae, Ap. J. 712 (2010) 1359 (and references therein). [16] L.F. Roberts, S.E. Woosley, R.D. Hoffman, Integrated nucleosynthesis in neutrino-driven winds, Ap. J. 722 (2010) 954. [17] A. Arcones, G. Martinez-Pinedo, Dynamical r-process studies within the neutrino-driven wind scenario and its sensitivity to the nuclear physics input, Phys. Rev. C83 (2011) 045809. [18] L.F. Roberts, S. Reddy, Medium modification of the charged current neutrino opacity and its implications, arXiv: 1205.4066 (2012). [19] } K.-L. Kratz et al., Explorations of the r-processes: Comparisons between calculations and observations of low-metallicity stars, Ap.J. 662 (2007) 39 (and references therein). [20] S. Bisterzo et al., The s-process in low-metallicity stars - II. Interpretation of high-resolution spectroscopic observations with asymptotic giant branch models, MNRAS 418 (2011) 284. [21] P. Möller et al., Nuclear properties for astrophysical and radioactive-ion-beam applications, ADNDT 66 (1997) 131; and New calculations of gross β-decay properties for astrophysical applications: Speeding-up the classical r-process, Phys. Rev. C67 (2003) 055802. [22] J.M. Pearson et al., Nuclear mass formula with Bogolyubov-enhanced shell-quenching: application to r-process, Phys. Lett. B387 (1996) 455. [23] K.-L. Kratz et al., Constraints on r-process conditions from beta-decay properties far off stability and r-abundances, J. Phys. G14 (1988) 331; and Nuclear structure studies at ISOLDE and their impact on the astrophysical r-process, Hyperfine Interact. 129 (2000) 185; and r-process isotopes in the 132Sn region, EPJA 25 (2005) 633. [24] O. Arndt et al., Decay of the r-process nuclides 137,138,139Sb, and the A=130 solar r-process abundance peak, Phys. Rev. C84 (2011) 061307. [25] U. Ott et al., On ways for making Xenon-HL, MAPS 44 (2009), A162. 6
1103.5499
1
1103
2011-03-28T22:02:19
Debris disk size distributions: steady state collisional evolution with P-R drag and other loss processes
[ "astro-ph.EP" ]
We present a new scheme for determining the shape of the size distribution, and its evolution, for collisional cascades of planetesimals undergoing destructive collisions and loss processes like Poynting-Robertson drag. The scheme treats the steady state portion of the cascade by equating mass loss and gain in each size bin; the smallest particles are expected to reach steady state on their collision timescale, while larger particles retain their primordial distribution. For collision-dominated disks, steady state means that mass loss rates in logarithmic size bins are independent of size. This prescription reproduces the expected two phase size distribution, with ripples above the blow-out size, and above the transition to gravity-dominated planetesimal strength. The scheme also reproduces the expected evolution of disk mass, and of dust mass, but is computationally much faster than evolving distributions forward in time. For low-mass disks, P-R drag causes a turnover at small sizes to a size distribution that is set by the redistribution function (the mass distribution of fragments produced in collisions). Thus information about the redistribution function may be recovered by measuring the size distribution of particles undergoing loss by P-R drag, such as that traced by particles accreted onto Earth. Although cross-sectional area drops with 1/age^2 in the PR-dominated regime, dust mass falls as 1/age^2.8, underlining the importance of understanding which particle sizes contribute to an observation when considering how disk detectability evolves. Other loss processes are readily incorporated; we also discuss generalised power law loss rates, dynamical depletion, realistic radiation forces and stellar wind drag.
astro-ph.EP
astro-ph
Noname manuscript No. (will be inserted by the editor) DEBRIS DISK SIZE DISTRIBUTIONS: STEADY STATE COLLISIONAL EVOLUTION WITH P-R DRAG AND OTHER LOSS PROCESSES M. C. WYATT1, C. J. CLARKE1 and M. BOOTH1,2,3 1 Institute of Astronomy, University of Cambridge, Madingley Road, Cambridge CB3 0HA, UK 2 Herzberg Institute of Astrophysics, 5071 West Saanich Road, Victoria, B. C., V9E 2E7, E-mail: [email protected] 3 University of Victoria, Finnerty Road, Victoria, B. C., V8W 3P6, Canada Canada Submitted: 26 January 2011, Revised: 1 March 2011 Abstract We present a new scheme for determining the shape of the size distribution, and its evolution, for collisional cascades of planetesimals undergoing destructive collisions and loss processes like Poynting-Robertson drag. The scheme treats the steady state portion of the cascade by equating mass loss and gain in each size bin; the smallest particles are ex- pected to reach steady state on their collision timescale, while larger particles retain their primordial distribution. For collision-dominated disks, steady state means that mass loss rates in logarithmic size bins are independent of size. This prescription reproduces the ex- pected two phase size distribution, with ripples above the blow-out size, and above the tran- sition to gravity-dominated planetesimal strength. The scheme also reproduces the expected evolution of disk mass, and of dust mass, but is computationally much faster than evolving distributions forward in time. For low-mass disks, P-R drag causes a turnover at small sizes to a size distribution that is set by the redistribution function (the mass distribution of frag- ments produced in collisions). Thus information about the redistribution function may be recovered by measuring the size distribution of particles undergoing loss by P-R drag, such as that traced by particles accreted onto Earth. Although cross-sectional area drops with age t −2.8, underlining the importance of understanding which particle sizes contribute to an observation when considering how disk detectability evolves. Other loss processes are readily incorporated; we also discuss gener- alised power law loss rates, dynamical depletion, realistic radiation forces and stellar wind drag. t −2 in the PR-dominated regime, dust mass falls (cid:181) Keywords circumstellar matter · planetary systems · debris disks · collisions 1 Introduction Many nearby main sequence stars are seen to be surrounded by m m-sized dust known as debris disks (see review in Wyatt 2008). The dust is short-lived indicating that larger plan- etesimals are present that feed the observed dust (Backman & Paresce 1993). It is thought 1 1 0 2 r a M 8 2 . ] P E h p - o r t s a [ 1 v 9 9 4 5 . 3 0 1 1 : v i X r a (cid:181) 2 M.C.Wyatt et al. that the large planetesimals are destroyed in mutual collisions, creating fragments that also collide to produce yet smaller fragments, and so on until the dust is small enough to be re- moved from the system by radiation pressure; i.e., that there is a continuous size distribution extending from small to large objects known as a collisional cascade. In this respect extrasolar debris disks are similar to the asteroid and Kuiper belts in the Solar System. The size distribution of large objects in the Solar System's belts is well char- acterised observationally (Gladman et al. 2009; Fuentes et al. 2010). There remains debate as to the origin of this distribution at the largest sizes (Durda et al. 1998; Bottke et al. 2005; Morbidelli et al. 2009), but it is recognised that these populations have undergone, and con- tinue to undergo, collisional evolution which means that their size distributions should also extend down to dust sizes. Exactly how the distributions extrapolate down is less well con- strained (e.g., Schlichting et al. 2009). However, dust is seen in the inner Solar System that migrates inwards past the Earth due to Poynting-Robertson drag (Leinert & Grun 1990; Der- mott et al. 2001), though it is debated as to whether this dust has an asteroidal or cometary origin (Durda & Dermott 1997; Nesvorn´y et al. 2010). One big difference with extrasolar debris disks is that, to be detected above the light from the star, their dust has to be orders of magnitude more abundant than dust in the Solar System. This means that it has been possible to ignore the effects of P-R drag when studing their evolution, since for such disks collisional timescales (that scale inversely with total dust mass) are much shorter than P-R drag timescales (that are independent of dust mass) (Wyatt 2005). The regime of such collision-dominated disks is now (relatively) well under- stood. Early analytical models treated systems in which the size distribution is defined by a single power law, and showed that disk luminosity is expected to stay constant, but to fall off inversely with age at late times (Dominik & Decin 2003; Wyatt et al. 2007a), in a manner consistent with the observed debris disk population (Wyatt et al. 2007b). More recent analyt- ical models allow for three phases in the size distribution to allow for planetesimals having a size-dependent dispersal threshold (Lohne et al. 2008; Heng & Tremaine 2009), and pro- vide a better fit to the size distribution expected from more detailed numerical models that solve for the evolution of an input size distribution given assumptions about the outcome of collisions between different sized particles (Krivov et al. 2006; Th´ebault & Augereau 2007). However, the analytical models cannot fit details in the expected size distribution, which is not just defined by power laws, and neither do they consider effects other than collisions. There is now a need to understand how loss processes affect the size distribution, and so the detectability, of a debris disk. Current instrumentation is probing dust levels at which drag may start to become important, and as searches for Earth-analogs intensify, it is in- evitable that lower levels of dust will be probed and the dynamics in this regime needs to be properly understood (Beichman et al. 2006). It is also becoming clear that stellar wind drag, which acts in a similar manner to P-R drag, may be important for young stars with high mass loss rates (Plavchan et al. 2005; Reidemeister et al. 2010). Although the competition between collisions and P-R drag has been considered in the Solar System (Leinert et al. 1983; Grun et al. 1985; Ishimoto 2000), such studies focus on explaining the currently observed dust distribution, and are not readily applicable to earlier epochs or to the debris disks of other stars. Collisional models that do study the long-term evolution of the asteroid belt typically exclude small particles and focus on fitting the large body size distribution (Bottke et al. 2005), whereas models for structures in the zodiacal cloud often only treat the dynamics of individual particles essentially ignoring collisions (Dermott et al. 1994; Grogan et al. 2001). However, more recently the balance of collisions and P-R drag has been incorporated into codes that evolve size distributions with time and so far results have been presented for two systems (Vitense et al. 2010; Reidemeister et al. Debris disk size distributions 3 2010). Other models are also being developed that incorporate a (less detailed) prescription for collisions into N-body simulations (Stark & Kuchner 2009; Kuchner & Stark 2010). Here we present a new scheme that can be used to determine the size distribution of a planetesimal belt undergoing destructive collisions as well as loss processes. This scheme is not expected to provide a more faithful measure of the size distribution than would be found by evolving a distribution forward in time. Rather its value is in the speed with which the distribution can be determined (in seconds), with its ability to reproduce details in the steady state size distribution, including when loss processes are acting, and with the simplicity of the analytical presentation which permits an analytical understanding of how various pro- cesses affect the size distribution. Although the scheme presented here is in its most basic form, considering the 1-dimensional size distribution (rather than, e.g., including spatial di- mensions), and treating collisional destruction with a 1-dimensional redistribution function (essentially averaging the outcome over all possible collisions), extension to extra dimen- sions is possible in principle. §2 outlines the development of the scheme by application to the collision-dominated regime, then §3 continues by detailing the effect of P-R drag; §4 considers other loss processes, and conclusions are given in §5. 2 Steady state in the collisional regime Consider a belt of planetesimals in which mk is the total mass in the k-th bin, where Dk is the typical size of planetesimals in the bin, and bins are logarithmically spaced in size, with k = 1 being the largest bin and working to smaller sizes with increasing k so that Dk+1/Dk = 1 −d where d is approximately the logarithmic bin width, since it is also assumed that d ≪ 1. Collisions between planetesimals cause the mass to be redistributed amongst the bins, and furthermore loss processes may cause mass to be removed from the belt, and there may be sources that add mass to the belt. From mass conservation for the k-th bin mk = m+c k − m−c k − m−l k + m+g k , (1) where the superscripts +c, −c, −l, +g refer to the mass gained from collisions in other bins, the mass lost from collisions in this bin, the mass removed by loss processes, and the mass gained from other sources, respectively. Note that our 1-dimensional prescription of the belt does not have a radial dimension and so considers loss processes such as P-R drag as a sink, rather than as a diffusive process (e.g. Gorkavyi et al. 1997). For now we consider the situation where m+g k = 0, but return to loss processes in §3. k = m−l 2.1 Mass loss is independent of size We define the redistribution function f (i, k) to be the fraction of mass leaving the i-th bin from collisions that goes into the k-th bin. Since mass must be conserved (cid:229) k=1 f (i, k) = 1. In this paper the model is being applied to situations in which collisions are destructive, so that f (i, k ≤ i) = 0 and (cid:229) k=i+1 f (i, k) = 1, though it may be possible to extend this to situations where accretion can also take place (see, e.g., Birnstiel et al. 2011). Let us further assume that the redistribution function is scale independent, so that f (i + n, k + n) = f (i, k), where n is any integer, which we can implement by defining a new redis- tribution function needing just one parameter F(k −i) = f (i, k). Putting this into equation (1) ¥ ¥ 4 M.C.Wyatt et al. and giving collisional gains and losses in terms of the redistribution function we find mk = k−1 i=1 i F(k − i) − m−c m−c k . (2) If we assume that the size distribution is in quasi-steady state, so that mk ≈ 0, and append s to any subscripts of quantities that are calculated in steady state, we find that m−c ks = k−1 i=1 m−c is F(k − i). (3) Since (cid:229) that l=1 F(l) = 1, then as long as (cid:229) l=k F(l) ≪ 1, one solution to equation (3) must be m−c ks = m−c is = C, (4) where C is a constant; i.e., the mass loss rate in logarithmic bins is independent of size. Note that the main assumption to arrive at this result was that the redistribution function is scale independent. However, there is a further assumption regarding whether the size distribution is infinite in extent (e.g., Dohnanyi 1969). Truncated distributions are considered in more detail in following sections, but here we note that for distributions truncated at large sizes, equation (4) is expected to apply at sizes for which the mass that would have been replenished from sizes above the truncation is negligible (i.e., it applies for k ≫ 1 and redistribution functions that are weighted toward larger sizes). 2.2 Mass loss rate Equations (3) and (4) can be used to determine the steady state size distribution, mks. To do so, we consider that the mass loss rate from the k-th bin due to collisions is k = mkRc m−c k, (5) where Rc k is the rate at which individual objects of size Dk undergo catastrophic collisions. A catastrophic collision is defined as one in which the mass of the largest remnant after the collision has less than half of the mass of the original object. This definition is used for our typical collision, not because we believe that catastrophic collisions are the only important collisions -- indeed it has been shown that cratering collisions with much lower energy may also result in significant mass loss (e.g., Kobayashi & Tanaka 2010) -- but because we expect the overall mass loss rate (i.e., including both catastrophic and cratering collisions) for all bins to scale with the rate of catastrophic collisions within them, at least for distributions that are infinite in extent. It may be that mass loss in fact scales with the rate of collisions in which the largest remnant has slightly more or less than 50% the mass of the original object. The catastrophic collision rate for a planetesimal of size Dk can be determined using the particle-in-a-box approach and integrating over the size distribution of impactors that have sufficient specific incident energy (Q) to cause a catastrophic collision, which is defined as the dispersal threshold Q⋆ D. For collisions that occur at relative velocity vrel the smallest impactors that cause catastrophic destruction are those of size XcDk, where Xc = (2Q⋆ D/v2 rel)1/3, (6) (cid:229) (cid:229) ¥ ¥ Debris disk size distributions 5 which we denote in the following equation as those with an index ick. Thus the collision rate, both in its discrete and continuous forms, is Rc k = ick i=1 3mi 2rp D3 i (Dk + Di)2Pik, = Z D1 XcDk n(Di)0.25(Dk + Di)2PikdDi, where r is particle density, Pik is the intrinsic collision probability used by other authors (e.g., Wetherill 1967) between particles i and k that is equal to p vrel/V for all particle sizes for applications in this paper, and V is the volume through which the planetesimals are moving, n(D) is the number of planetesimals in the size distribution per unit diameter, and the gravitational focussing factor has been neglected for this calculation. (7) (8) (10) (12) (13) 2.3 Analytical steady state method The continuous form of the collision rate (equation 8) can be used to determine the steady state size distribution analytically, for certain assumptions. By integrating the mass distribu- tion (pr D3/6)n(D) over the bin we find that for small d mk ≈ (pr /6)n(Dk)D4 d , (9) k and so m−c k Further assuming that (cid:181) n(Dk)D4 n(Di)(Dk + Di)2dDi. XcDk kZ D1 n(D) = KD−a (11) and that a > 3 and Xc ≪ 1 so that collisions close to the catastrophic collision threshold dominate the rate (i.e., so we can take only the term with D2 k in the integral and ignore the upper limit), gives , If Q⋆ D (cid:181) Db then we find that m−c k (cid:181) X 1−a c D7−2a k . m−c k (cid:181) D7−2a +(1−a )b/3 k , and for this to be independent of size as required for steady state (equation 4) a = (7 + b/3)/(2 + b/3). (14) This reduces to the canonical a = 3.5 distribution for the situation where strength is inde- pendent of size (Dohnanyi 1969; Tanaka et al. 1996), and replicates the distribution found analytically by other authors when strength depends on size (O'Brien & Greenberg 2003; Lohne 2008; Lohne et al. 2008; Heng & Tremaine 2009; Birnstiel et al. 2011; Belyaev & Rafikov 2011). 2.4 Numerical steady state method The discrete form of the collision rate (equation 7) can also be used to determine the steady state size distribution numerically. (cid:229) 6 2.4.1 Method 1 M.C.Wyatt et al. If we impose a constant mass loss rate per logarithmic bin, as argued to be the case in §2.1, then we can rearrange equation (5) and combine with equation (4) to get mks = C/Rc k. (15) This can be solved iteratively, and converges quickly to find the steady state distribution. The shape of the resulting distribution is independent of the total mass in the distribution, because multiplying the whole distribution mks by N would result in the mass loss in the k-th bin m−c k both increasing by N2. Application of this method is k discussed in §2.6.2. and the mass gain m+c 2.4.2 Method 2 Alternatively we can relax the assumption of equation (4) and apply a similar method more directly to equation (3), and so equate mass loss with mass gain in each bin by iterating mks = Ck/Rc k, Ck = k−1 i=1 m−c is F(k − i). (16) (17) In this case, by the arguments of §2.1, we would expect to end up with a solution for which Ck is approximately constant (at least for most particle sizes). To implement this method requires the redistribution function F(l) to be defined (see §2.5). Without further constraints this method would converge to a distribution with zero mass. This is because the lack of larger particles means that the mass input rate to the top bin is zero, which can only be balanced by having zero mass there; if that is the case then the mass input to the bin below this is zero, and so on. This issue can be remedied by con- straining the largest bin(s) to have either a fixed mass distribution, or a fixed mass input rate (e.g., by adding m+g to Ck). Further numerical tricks are also required to achieve conver- 0 gence. If numeric subscripts indicate the iteration number, or more specifically a variable calculated using the mass distribution from that iteration, then we found convergence with the following two-step scheme mks1.5 = [(Ck1/Rc mks2 = [(Ck1/Rc k1) + mks1]/2, k1.5) + mks1]/2. (18) (19) This method is employed in §2.9, and further extended to incorporate loss due to P-R drag in §3. 2.5 Redistribution function Where it is necessary in this paper, we assume that the fragments from the destruction of objects of size Dk follow the power law of equation (11), with parameters identified by subscript r, from sizes of h rmaxDk down to infinitesimally small particles. Unless otherwise stated we consider h rmax = 2−1/3 (i.e., the largest objects have half the mass of the original), which corresponds to catastrophic collisions dominating the redistribution of mass. In this paper we also consider a function with h rmax = 1, which corresponds to cratering collisions (cid:229) Debris disk size distributions 7 dominating mass redistribution. For h rmax < 1 we also include a small but finite fraction of mass in particles in the h rmaxDk to Dk size range using a steep power law (with a = −20)1. The slope is always assumed to have a r < 4 so that the mass in fragments is weighted toward large objects. These assumptions mean that the fractional mass distribution of these fragments is ¯m(D) ≈ (4 −a r)D3−a r (h rmaxDk) a r −4 across the appropriate range. Integrating over the size range of bin Dk+l, and assuming d ≪ 1, gives F(l) ≈ h a r−4 rmax (4 − a r)d (1 − d )l(4−a r). (20) Although it is the redistribution function F(k − i) that is employed throughout this pa- per, it is worth considering a more general form of the redistribution function, since the techniques presented herein may also applicable in this instance. Most numerical models of collisional evolution define the redistribution function as f2(i, j, k), the fraction of mass leaving the i-th bin from collisions with the j-th bin that goes into the k-th bin. In general such redistribution functions also have some form of scale independence, and can, e.g., be written as F2( j − jci, k − i), where jci is the index of the size that causes catastrophic de- struction of objects of size Di. This is because a collision with the jci-th bin ends up, by definition, with a largest remnant that has half the mass of the original, and collisions with other bins end up with largest remnants that scale as a function of the ratio Q/Q⋆ D, and so have redistribution functions that only depend on j − jci. If we assume that F2( j − jci, k − i) is the true redistribution function, then we have im- plicitly averaged over j, weighted by the frequency of such collisions, to get F(k − i) as the redistribution of mass in a typical collision. Thus we can write F(k − i) = j=1 m j(Di + D j)2D−3 j F2( j − jci, k − i) jci j=1 m j(Di + D j)2D−3 j . (21) This illustrates that the F(k − i) redistribution function may not be scale independent, even if F2( j − jci, k − i) is, because it includes information about the size distribution (see also Belyaev & Rafikov 2011). In particular, it may be expected to be different at locations where there are sharp transitions in the size distribution. These issues may also contribute to the timescale of mass loss in a typical collision being different to the timescale for catastrophic collisions, as discussed in §2.2. Nevertheless, we retain the F(k − i) redistribution function for this paper, since it is expected to be valid over most of the size distribution, and is shown in later sections to be in agreement with more detailed models. 2.6 Application to realistic dispersal laws As reported by several authors studying collision outcomes (e.g., Durda et al. 1998; Benz & Asphaug 1999), the size dependence of the dispersal threshold can be expressed as D = QaD−a + QbDb, Q⋆ (22) where the subscripts a and b refer to contributions from the planetesimal's material strength and from its self-gravity, respectively, and a and b are positive. The weakest planetesimals then have size Dw =(cid:18) aQa bQb(cid:19)1/(a+b) ; (23) 1 This is necessary to allow the size distribution to be calculated at late times in §2.9 when Dt > h rmaxD1. (cid:229) ¥ (cid:229) 8 M.C.Wyatt et al. Table 1 Summary of (some of) the parameters used in the paper: power law indices that refer to equation (11), and diameters. Diameters (e.g., Dx) also have corresponding indices (e.g., ix) to denote bin number. Power law index a p a b a a a pr a l a r Diameter D1 Dt Dw Dpr Dl Dbl Regime Primordial Collisional (gravity) Collisional (strength) PR-dominated General loss-dominated Redistribution function Meaning (or transition) Largest planetesimals Primordial-collisional Strength-gravity scaling Collisional-PR-dominated Collisional-loss-dominated Radiation pressure-dominated Equation given 14 24 41 53 given Equation given 29 23 37 51 26 e.g., for planetesimals made of basalt impacting at 3 km s−1 (Benz & Asphaug 1999), rea- sonable values are Qa = 790 J kg−1, a = 0.38, Qb = 0.017 J kg−1, b = 1.36, for which Dw = 0.23 km. Here we adopt slightly different values of Qa = 620 J kg−1, a = 0.3, Qb = 5.6 ×10−3 J kg−1, b = 1.5, to allow comparison with the results of Lohne et al. (2008) in later sections. 2.6.1 Analytical 2-phase distribution A reasonable approximation is that the only contribution to Q⋆ D is from the strength compo- nent for D ≪ Dw (the strength regime) and from the gravity component for D ≫ Dw (the gravity regime). Application of the analysis of §2.3 then suggests that the size distribution should have two phases described by power laws (equation 11) with a a = (7 − a/3)/(2 − a/3), (24) in the strength regime and a b = a given by equation (14) in the gravity regime. Power law indices in this paper, like a a, that refer to equation (11) are summarised in Table 1, along with a summary of diameters that define where these apply. For the dispersal law assumed here, a a = 3.63 and a b = 3.0. Although one might expect the distribution to be continuous at Dw, closer inspection shows that this cannot be the case, since the condition given by equation (4) should hold regardless of the dispersal law, and so should hold on both sides of (and at) the transition. The discontinuity at Dw was pointed out by O'Brien & Greenberg (2003). The relative heights of the two distributions, and so the magnitude of the discontinuity, can be determined by including the constants of proportionality in equation (13). These constants are (cid:181) K2 a (1 − a a)−1(v2 for the strength regime, and likewise except with subscript b for the gravity regime, resulting in a a/3Q(1−a a)/3 rel/2) a Ka Kb =r 1 − a a 1 − a b (cid:18) v2 2 (cid:19) rel (a b−a a)/6 Q(1−a b)/6 Q(1−a a)/6 b a ! . (25) Scaling the total mass in the 2-phase distribution to that required, say Mtot, means that the distribution is now completely defined. In Figure 1 we show with a dashed line the Debris disk size distributions 9 Fig. 1 Steady state size distribution for particles between 1m m and 1000km undergoing catastrophic colli- sions. The y-axis is the mass in logarithmic bins centred on sizes given on the x-axis, where bin width is given by d = 0.01. Three distributions are shown for relative velocities of 0.6, 1.9, 6.5 km s−1 using the same law for the dispersal threshold. The distributions are scaled to total masses of 0.01, 0.1 and 1 times the mass of the Earth to allow these to be distinguished (though the shape of the distribution is independent of total mass). The distributions are calculated using the analytical 2-phase (dashed line) and numerical method 1 (solid line). predicted distributions for the assumed dispersal law for different values of relative velocity. Note that we did not change the dispersal law with collision velocity, although this might be expected (Benz & Asphaug 1999; Stewart & Leinhardt 2009). The total mass in the distributions were scaled to allow them to be distinguished. The main thing to note from the analytical distribution (for now) is that the discontinuity is bigger for larger relative velocities, as expected from equation (25). It should also be remarked that for low collision velocities, Xc can be greater than 1 for the largest particles, so that our simplistic view of collisions in §2.2 implies zero mass loss rate for some particles near the top end of the distribution. In practise these particles can still lose mass through cratering collisions, illustrating how our assumption that mass loss rates scale only with the catastrophic collision rate breaks down in certain regions. Such particles could not reach steady state if our assumptions were valid, and a more detailed treatment is needed to assess evolution in this regime (e.g., §2.5). Rather than assume some primordial distribution for such particles, here we set the mass to zero in all bins for which Xc > 1 at the top end. Note that, although Xc can also be greater than 1 at small sizes, this does not prevent the destruction of such grains, since they can be destroyed by particles larger than themselves. 2.6.2 Numerical 2-phase with ripples There were many approximations in the analytical 2-phase model that are overcome in the numerical model. Notably the numerical model should be valid regardless of the shape of the dispersal law, and should also be applicable (within the limitations of the assumptions) in the situation where the size distribution is finite in extent, being truncated at small and/or large sizes. Figure 1 also shows with solid lines the result of numerical method 1 (§2.4.1) that solves is indeed equation (15). Convergence of this scheme was confirmed by checking that m−c k 10 M.C.Wyatt et al. the same for all bins, as expected by construction. The first thing to note is the excellent agreement with the 2-phase analytical model. There is slight quantitative disagreement on the slope of the distributions, which we attribute to the terms that were discarded from equation (10) when deriving equation (12). Thus, from Figure 1, we expect the slope of the mass distribution in the strength regime to be slightly steeper than predicted by a a, and for the distribution in the gravity regime to be slightly shallower than that predicted by a b. As expected for a realistic distribution, there is no discontinuity at Dw. However, as discussed by O'Brien & Greenberg (2003, see their Fig. 3) and previously noted by Durda et al. (1998), the transition at Dw does cause a ripple that extends to larger sizes, and is captured by the numerical model. The transition does not affect the distribution in the strength regime because, as discussed in O'Brien & Greenberg (2003), the distribution in the strength regime is independent of that in the gravity regime, which follows in our analysis because m−c k depends almost entirely on the number of smaller objects in the distribution with size around XcDk. We find that the magnitude of the ripple (i.e., the departure from the power law slope) depends on the collision velocity, with higher velocities causing larger ripples. The ripple decreases in magnitude to larger sizes, with a peak-to-peak wavelength that also depends on collision velocity and varies with size (decreasing wavelength to larger sizes). All such effects follow naturally from the dependence of Xc with size, which is the parameter that sets the magnitude and wavelength of the ripple, since for particles of a given size the larger Xc is the nearer in size particles have to be to affect their number. In the gravity regime Xc increases with size (and velocity) causing the observed dependence. This also means there should be a dependence on Q⋆ D in that weaker planetesimals should also result in a more pronounced ripple with a longer wavelength. A similar ripple is also seen at small sizes, caused by the fact that the distribution is truncated at 1m m. While this truncation is implemented simply by not considering bins < 1 m m, a truncation is physically motivated by radiation pressure, since grains smaller than Dbl, where Dbl ≈ (2300/r )(L⋆/M⋆) (26) in m m (for r in kg m−3, and L⋆ and M⋆ in Solar units), are placed on hyperbolic trajec- tories as soon as they are created (e.g., Burns et al. 1979). In this case the lack of < 1m m impactors causes the equilibrium number of 1m m grains to be enhanced. The consequence of this is an enhanced destruction rate of objects of sizes typically destroyed by 1m m grains (i.e., those for which XcD = 1m m). This causes a dearth of such grains and so a higher equi- librium number of grains that would have been destroyed by them, and so on. Such ripples have previously been discussed by Campo-Bagatin et al. (1994) and seen in the simulations (Th´ebault et al. 2003; Krivov et al. 2006). There is a similar dependence on collision veloc- ity of the magnitude of the ripple, and its wavelength, in that both increase with vrel. This can again be understood in terms of Xc, which is both smaller for larger velocities (and so particles can be affected by those at much smaller relative sizes), and which decreases with size in the strength regime. This ripple would also be expected to be more pronounced for weaker planetesimals. The details of this ripple should, however, not be overinterpreted. For example, if the truncation is caused by radiation pressure, then grains just above the blow- out limit (equation 26) should also have a higher eccentricity than larger particles, which numerical simulations show would affect the shape of this ripple (Krivov, Mann & Krivova 2000; Th´ebault & Augereau 2007). The ripple may also be less prominent if particles have a range of compositions and so a range of truncation sizes and strengths. In short, all of the features of the steady state size distribution arising from collisions that have been discussed in the literature (power laws and ripples) are reproduced with our Debris disk size distributions 11 simple numerical scheme. This allows the shape of the distribution to be analysed without having to evolve size distributions forward in time until they reach steady state, which can be time-consuming for large particles with low collision rates, if timesteps are set by the small particles that evolve much faster. This should allow us to consider how the steady state distribution depends on parameters such as the shape of the dispersal threshold law and the collision velocity. But more importantly for this paper, allows us to consider how that shape is affected by loss processes. 2.7 Timescale to damp perturbations from steady state The steady state size distributions presented in the preceding sections have some issues at the top end of the distribution. It was already discussed that the numerical scheme employed, to keep mass loss rate independent of size, should be valid for most of the size range, but that for a distribution that is truncated at sizes above D1 this should be invalid for the size range in which the redistribution function means that objects of size D1 (and larger, had they existed) deposit a significant fraction of their mass. Although a numerical scheme was outlined that should be able to account for this (equation 16), we should consider the validity of the assumption that the largest particles are in steady state. First consider a bin to have a mass that departs from that expected when the distribution is in quasi-steady state by a small fraction e k, so that mk = mks(1 + e k). (27) Since m+c tion), substituting equation (27) into equation (1) gives k and Rc k are independent of mk for small enough bins (or a small enough perturba- In other words, perturbations are damped on the collision timescale. e k/e k = −Rc k. (28) 2.8 Analytical 3-phase distribution Another way of interpreting equation (28) is that the collision timescale also defines the timescale for a given bin size to reach steady state. Thus for a given evolution timescale, tage, all particles with collision rates higher than 1/tage (i.e., all of the smallest particles) would be expected to have reached steady state, which we denote with subscript ss, whereas those with smaller collision rates (i.e., the largest particles) would be expected to retain their primordial size distribution, which we define here using equation (11) with a subscript p. We define Dt to denote those particles with index k = it for which Rc k(Dt) = 1/tage. (29) This motivated Lohne et al. (2008) to define a 3-phase size distribution comprised of the 2 phases in steady state given in §2.6.1, in addition to the primordial phase. These authors connected the size distribution in each phase continuously. However, if it is assumed that there are two distinct portions of the size distribution, primordial and steady state, with the destruction of primordial particles feeding those in steady state (in a manner similar to the approach of Dominik & Decin 2003), then there may be a discontinuity at this transition. 12 M.C.Wyatt et al. The reason is that bins in the steady state population have mass input equal to that lost. This means that the rate of mass loss from the steady state population is the sum over this population considering just the mass put into particles small enough to fall outside the cascade imin m− ss = C i=it k=imin F(k − i), (30) where we have used the fact that m−c is = C in the regime considered to take it outside of the sum, and imin = ibl for initial considerations in this paper, but would be the larger of ibl and ipr if a consideration of P-R drag was necessary (§3). Since C scales with the square of the total mass in steady state (for fixed it), so does the mass loss rate. The mass input to the collisional cascade, however, is set by the sum over all bins in the primordial distribution, considering just the mass put into sizes in the collisional cascade m+ ss = it i=1 m−c i imin k=it F(k − i), (31) is set mostly by the primordial distribution. where m−c i Thus in this approximation, the mass in the steady state population represents a balance between the two, i.e., m+ ss = m− ss so that C = it i=1 m−c i imin i=it imin k=it F(k − i) k=imin F(k − i) (32) and does not necessarily lead to a continuous distribution at Dt. This is particularly so if Dt is close to D1, since at this point the mass input to the steady state population from primordial particles is low, and so that population must also have low mass and the discontinuity is large. In fact the transition from primordial to steady state is not expected to be discontinuous, and such a discontinuity may compromise the validity of the scale independence assumption of our redistribution function (§2.5). Nevertheless it is hoped that this approximation still captures the important physics. For example, although the discontinuity is large in the regime in which Dt approaches D1, the primordial population may, in some respects such as the long term evolution of the steady state population, be irrelevant at this point. This is because, in §2.9, the system age is set to be equal to the collisional lifetime of objects of size Dt, which may well have a negligible contribution from the primordial population. If this is the case, then the collisional lifetime would be expected to scale inversely with the mass in the steady state population, leading to disk mass falling off inversely with age, as expected from previous analytical and numerical studies (e.g., Dominik & Decin 2003; Wyatt et al. 2007a; Lohne et al. 2008). 2.9 Evolving the size distribution The scheme of §2.8 allows us, for a given primordial distribution and transition size, to define the discontinuity and so derive a 3 phase distribution analytically. Further equating the collision lifetime of objects of size Dt with the system age would also permit the evolution of an input primordial size distribution to be determined, in a similar manner to Lohne et al. (2008). Here we follow a comparable approach, except that the 3-phase distribution is (cid:229) ¥ (cid:229) (cid:229) (cid:229) (cid:229) (cid:229) (cid:229) (cid:229) ¥ Debris disk size distributions 13 derived using numerical method 2 (§2.4.2; i.e., using equation 16), with the largest size bins (above Dt) fixed at the primordial size distribution. This automatically captures the discontinuity (and its effect on the collision rate of particles at the transition) and achieves steady state in the appropriate size range. To consider how accurately our scheme reflects simulations that solve for the evolution using the full equation (1) (or its equivalent), we compare our results with those of the ACE run ii-0.3 of Lohne et al. (2008). That run considered the evolution of a 7.5-15AU planetesimal belt with a maximum eccentricity of 0.3, maximum inclination of 0.15 rad, and largest planetesimals 148km in diameter, for a Sun-like star. The primordial distribution had 1M⊕ distributed according to equation (11) with a p = 3.61. The result of our scheme for these parameters, with d = 0.02 (and so 1279 diameter bins), is shown in Figure 2. Figure 2 should only be compared qualitatively with figures in Lohne et al. (2008), since that paper mistakenly used a slightly different Qb = 0.014 J kg−1, a factor of 2.5 higher than stated (and used in our nominal runs, §2.6) (T. Lohne, priv. comm.). However, a quantitative comparison is given within Figures 2c, 2d and 2e, that also show ACE results for run ii-0.3 with Qb = 5.6 × 10−3 J kg−1 (A. Krivov, priv. comm.). This comparison shows that our scheme mostly reproduces the same features, both qualitatively and quantitatively, but there are also some qualitative and quantitative differences that merit further discussion. For example, the size distribution at specific ages (Figure 2a) compares well with that of Lohne et al. (2008). Although the details of the ripples close to the blow-out limit are not identical, they are qualitatively similar, and differences are expected due to the models' different assumptions about collisional outcome and in their treatment of the dynamics of grains close to this limit. However, the discontinuity at the transition size is not apparent in the Lohne et al. simulations. Figure 2b illustrates why this discontinuity cannot be physical, since although the mass evolution of different bins shows generally good agreement, all sizes undergo a sudden loss of mass at the time they are assumed to reach steady state, whereas mass evolution should be continuous. For this reason we do not expect our scheme to reproduce details of the size distribution at the transition between primordial and steady state, such as the divot found by Fraser (2009). The most instructive plots for comparative purposes are those for the evolution of total mass (Figure 2c) and dust mass (Figure 2d). Again, these show qualitatively good agree- ment. For example, total mass exhibits a slow fall-off with time from 0.1Myr, but speeds up around 100Myr tending to a fall-off with age (cid:181) t −0.94, close to the inverse fall-off with time predicted above. However, the total mass lost at 20Gyr is ∼ 92% of the initial mass, whereas the ACE run loses closer to 81% on the same timescale, indicating that mass loss is quan- titatively faster in our scheme. The dust mass shows similarly good qualitative agreement, with dust mass falling off between 1-200Myr (cid:181) t −0.41 for ACE. Although it is noticeable that our scheme underpredicts the dust mass by a factor ∼ 2 at most ages, this is a relatively small difference given the many orders of magnitude in mass over which the size distribution is being considered. However, our scheme also predicts a turnover beyond ∼ 2 Gyr, tending to dust mass falling off (cid:181) t −0.97. This is in qualitative agreement with Lohne et al. who predict an inverse fall-off with age at late times, once the largest planetesimals have come to collisional equilibrium, except that the turnover in their plot is not noticeable since it occurs much later. t −0.34 for our scheme and (cid:181) Figure 2e shows that the dependence of dust mass on total mass in our scheme is very similar to that of Lohne et al. (2008), apart from the small underprediction of dust mass noted above. This suggests that a reinterpretation of age could lead to greater agreement between the schemes. Remember that the size distribution is worked out for a given Dt, and the corresponding age is then derived by working out the collision rate of objects at 14 (a) (c) (e) M.C.Wyatt et al. (b) (d) (f) Fig. 2 Evolution of a planetesimal belt with the parameters quoted for ACE run ii-0.3 of Lohne et al. (2008). For comparison, the results for such an ACE run are reproduced on (c), (d) and (f) with a dotted line (A. Krivov, priv. comm.), whereas for (a) and (b) qualitative comparison with figures in Lohne et al. (2008) is possible (see text for details). (a) Mass (mk) in different size bins (Dk) at times of 0.05, 0.5, 5, 50, 500 Myr and 5, 20Gyr (compare with Lohne's Fig. 4 bottom). (b) Mass (mk) in 50 different (but equally-logarithmically- spaced) size bins as a function of time (compare with Lohne's Fig. 4 top), where bin size could be deduced from plot (a). (c) Total mass as a function of age. (d) Dust mass (in particles D < 2mm) as a function of age. (e) Dust mass as a function of total mass. (f) Transition diameter (from primordial to steady state) as a function of age. In practise, the size distribution is calculated for a given Dt, which is then converted into an age by determining the collision lifetime of objects at the transition. Debris disk size distributions 15 the transition and equating that with the inverse of the age (see Figure 2f). In Figure 2 we consider some of the model parameters that may affect evolutionary timescales. First, since collisions with the primordial population may shorten the collision lifetime at late times (when such a population is not physically realistic), the dashed line shows what happens when collision rates are calculated assuming that steady state and primordial particles can only collide with other members of their respective sub-population. This is implemented by changing the indices over which the collision rate is summed in equation (7). As expected this results in slightly longer evolutionary timescales for both total and dust mass. Notably the evolution is now closer to that found by the ACE simulations in that 85% of the total mass has been lost at 20Gyr, and the turnover in the dust mass evolution at 2Gyr is no longer present. We also considered the effect of the redistribution function, by trying a cratering- like function weighted toward objects of size Dk (with h rmax = 1 and a r = 3.9, see dash- triple-dot line). This has very little effect on the evolution of either total mass or dust mass, and it can be concluded that the size distribution in the collisional regime has very little dependence on the redistribution function (but can be used as a probe of the latter in the P-R drag regime, §3). Thus, despite this limited comparison, it seems that our prescription matches the results of more detailed simulations without the need for such time-consuming simulations to deter- mine the evolution on long (>Gyr) timescales. Although the 3-phase analytical prescription given in Lohne et al. (2008) does likewise, our model has the benefit that we reproduce more detailed features of the size distribution in the steady state regime (such as ripples), and moreover it is possible to consider additional effects such as loss processes like P-R drag (see §3). We expect that our scheme would replicate the results of other sets of full simulations (e.g., with different assumptions about collision outcome), though a detailed comparison would be needed to quantify (and improve through refining the implementation of the scheme) the level of agreement. 3 Steady state with P-R drag Despite the uncertainties in the details of how the size distribution transitions from primor- dial to steady state, equation (28) demonstrates that the small size end of the distribution is firmly in steady state and so can be treated as such. Consider now a belt in which both col- lisions and loss processes are acting. In this section we will consider the effect of Poynting- Robertson (P-R) drag. Poynting-Robertson drag is the dominant loss process for dust in the asteroid belt. 3.1 Analytical fourth phase To work out the loss rate from the belt we consider the timescale for dust particles to migrate from rmid to rmid − dr/2. Clearly, such particles have not yet been removed from the system, but are then fed into an inner region which should then be considered in some other way; our prescription means that we are only considering the size distribution of particles in the belt region. The loss rate by P-R drag from the belt for individual particles is Rpr k = AprD−1 k Apr = 2.9 × 10−6L⋆/(M⋆r drrmid(1 − 0.25dr/rmid )), , (33) (34) 16 M.C.Wyatt et al. where rmid and dr are in AU, L⋆ and M⋆ in solar units, and r rate in yr−1 with Dk in m. Thus the loss rate from the k-th bin is in kg m−3, to get a collision Because mk (cid:181) D4−a k m−pr k = mkAprD−1 k . (35) (from equations 9 and 11), then in the steady state collisional regime where mkRc k is independent of Dk, the loss rate from collisions is Rc k (cid:181) D a −4 k . (36) Thus in both the strength and gravity regimes this drops slower than P-R drag, which falls off as Rpr . Thus it is possible that at some size, which we call Dpr (with a corresponding k index ipr), these two are equal, and so (cid:181) D−1 k k(Dpr) = Rpr Rc k (Dpr). (37) For sizes smaller than Dpr the loss is dominated by P-R drag, whereas for sizes larger than Dpr losses are dominated by collisions, an assumption that will be justified later on. For some disks Dpr is smaller than the size for which particles are removed by radiation pressure (Dbl), and the mass of particles removed by P-R drag is insignificant. In fact this is necessarily the case in most debris disks that are bright enough to be detectable (Wyatt 2005). However, it is not the case for the Solar System's zodiacal cloud, and it will not be the case for the expected population of fainter extrasolar debris disks we have yet to discover. First let us work out what the steady state distribution is for the size range in which P-R drag is the dominant removal mechanism. If we assume that the rate of collisional loss in this regime m−c and so can be ignored, then in steady state the equivalent to equation (3) is k ≪ m−pr k m−pr ks = m−c is F(k − i), (38) ipr i=1 where the sum only extends up to i = ipr, since mass loss from i > ipr is assumed not to be collisional. Since m−c is = Ck is approximately independent of size in the collisional regime (i.e., Ck = C), this can be taken outside the sum giving the steady state distribution as mks = A−1 pr CDk ipr i=1 F(k − i). (39) Thus the size distribution in the P-R drag regime has a slope that is set by the redistribu- tion function modified by an additional factor of Dk. This prescription for the distribution is continuous at Dpr because at that location the sum in equation (39) is unity and so the mass pr) = C/Rpr calculated in the P-R drag regime at that location is mks(D− k , which is the same pr) = C/Rc as that calculated in the collisional regime mks(D+ k. For the redistribution function a r −4, which means defined in §2.5 by a r, the sum works out to be (cid:229) i=1 F(k − i) (cid:181) ipr that pr CDpr(Dk/Dpr)5−a r . mks ≈ A−1 (Dpr/Dk) (40) Or in terms of equation (11) with subscript pr, a pr = a r − 1. (41) This is important because it means that the size distribution of small particles in the P-R drag regime is an indirect measure of the slope in the redistribution function. (cid:229) (cid:229) Debris disk size distributions 17 Since a r < 4 the size distribution in the P-R drag regime is depleted in small particles. The main consequence of this is that the rate of collisional losses decreases toward smaller particles. This is because the mass loss rate due to collisions given by equation (13) applies regardless of whether the slope in the size distribution a is set to make the exponent in this equation equal to zero (as is the case in the collisional steady state regime). Given that mk (cid:181) D4−a this means that, for particles that are small enough to be in the strength regime, k Rc k (cid:181) D3−a −(1−a )a/3 k . (42) So in the P-R drag regime where a is set by the redistribution function, for the parameters . Thus, as we have assumed that a r < 4, collisional loss rates assumed here, Rc k increase with size in the P-R drag regime, meaning that collisional losses can be ignored for all particles smaller than Dpr, justifying the earlier assumption. (cid:181) D3.8−0.9a r k 3.2 Numerical size distribution The steady state distribution, including the influence of P-R drag, can be found numerically by extension of method 2 (§2.4.2). Again we equate mass loss with mass gain in each bin, to derive a revised version of equation (16) mks = Ck/(Rc k + AprD−1 k ), (43) where Ck is again given by equation (17). This method is used in Figure 3 to illustrate the points raised in §3.1 regarding the resulting distribution. To simplify the discussion, these simulations use exactly the same parameters as those used in the comparison with run ii-0.3 of Lohne et al. (2008) that was discussed in §2.9 and Figure 2. The main difference is that here we include the effect of P-R drag, and we also vary the parameter a r. Since the effect of P-R drag only becomes notice- able with these parameters for ages ≫ 10Gyr, its exclusion from the previous section (and from Lohne et al.'s simulations) does not compromise those results. However, this means that we are considering here the distribution at ∼ 500 Gyr ages that are greater than the age of the Universe. Although it seems ridiculous to consider such ages, this is not a major concern, since the same features would appear on shorter timescales in belts with different parameters (e.g., with smaller D1 or rmid), or in those started with low total mass, and effects such as dynamical depletion can be modelled using collisional evolution timescales longer than the system age (e.g., Bottke et al. 2005; §4.1.2). Figure 3a shows how P-R drag causes a turnover in the size distribution toward smaller particles, and how the slope in the P-R drag regime is set by the slope of the redistribution function, which is exactly as expected (remembering that the slope on this figure is expected to be 5 − a r). It also shows how insensitive the rest of the size distribution is to the redis- tribution function. It is evident that the ripples in the size distribution at small sizes have disappeared, which is expected above the blow-out limit, as the loss of particles just above this limit is no longer controlled by collisions. However, the turnover at Dpr has the potential to cause a ripple in the size distribution just above this size, which is not seen presumably because the turnover is a gradual rather than sharp feature. As for the location of the turnover, Figure 3b shows that the diameter for which Rc k = Rpr k is Dpr = 660, 430, 260 m m for a r = 3.0, 3.5, 3.9, respectively, noting that Rc k is calculated from the size distribution of Figure 3a rather than that with no consideration of P-R drag (which from extrapolation of the power law in the collisional regime would have resulted in 18 M.C.Wyatt et al. (a) (b) (c) Fig. 3 Steady state size distribution of a planetesimal belt with the same parameters as assumed in run ii-0.3 of Lohne et al. (2008), but including loss due to P-R drag. Different assumptions about the slope in the redistribution function, a r = 3.0,3.5,3.9 are shown on all plots with dotted, solid and dashed lines, respectively. The evolutionary age was set to give the same mass of 1m particles for the different assumptions, corresponding to an evolutionary age of 290, 540 and 2100 Gyr for the different a r respectively. (a) Mass (mk) in different size bins (Dk). (b) Loss timescales for particles of different sizes in these distributions due to collisions and P-R drag. (c) Mass loss rate from the inner edge of the belt due to P-R drag. The dash- dot line shows the polynomial fit to the accretion rate of dust onto the Earth from the LDEF satellite (Love & Brownlee 1993) scaled up by a factor 1/(40Dk) to account approximately for the accretion efficiency to recover the size distribution in the vicinity of the Earth. Debris disk size distributions 19 a smaller value of Dpr ≈ 150 m m that is independent of a r). For all values of a r, this is a factor of 1.9-2.0 times larger than the sizes obtained from the intersections of power law size distributions describing the P-R drag and collisional regimes, and a factor of 1.5-10 times larger than the peaks in the cross-sectional area distributions. The final Figure 3c shows the mass loss from the belt due to P-R drag. Since this mass can be considered as a gain term for a population of particles that are evolving due to P-R drag we denote this as m+ , and so from equations (9) and (11) is also (cid:181) D3−a . Thus in the P-R drag regime, D < Dpr, this distribution is set by the redistribution k function kpr which is (cid:181) mkD−1 k (44) (45) kpr(D < Dpr) (cid:181) D4−a r m+ k , whereas in the collisional regime, D > Dpr, it is set by the strength law kpr(D > Dpr) (cid:181) D(−3−2a)/(6−a) m+ k . As mentioned above, this mass is not lost from the system, but continues to evolve toward the star due to P-R drag, also suffering collisions on the way. It is beyond the scope of this paper to include the extra radial dimension needed to consider the evolution of this material properly. However, we can consider what happens if collisions no longer amend the size distribution, and the particles also pass a planet onto which they may be accreted. In this case the size distribution of accreted particles would simply be given by Figure 3c modified by a size dependent accretion efficiency Pacc. Accretion efficiencies may plausibly vary Pacc (cid:181) Dk due to the slower motion of larger particles past the planet (e.g., equation B2 of Wyatt et al. 1999), though the detailed dynamics may be more complicated (Kortenkamp & Dermott 1998). The assumption that collisions can be ignored is likely too crude to be useful for specific applications; in particular, collision timescales may be expected to be shorter than P-R drag timescales for particles larger than Dpr. However, it is notable that this size distribution is not so dissimilar to that derived for the accretion of dust onto the Earth by LDEF (Love & Brownlee 1993), a scaled version of which is also shown on Figure 3c. This suggests that, with further work, it may be possible to learn about the redistribution function from such ob- servations. For example, if the collisionless and Pacc (cid:181) Dk assumptions held, then Figure 3c shows that the LDEF accretion rate for < 200 m m particles would imply a redistribution function with a r slightly below 3.5, perhaps with a turnover to a smaller value below a few 10s of m m. Grun et al. (1985) also noted that the size distribution in the zodiacal cloud at 1AU may be indicative of the redistribution function, from a consideration of collisional gains and losses due to collisions and P-R drag, though these authors favoured the inter- pretation that the zodiacal cloud is not in steady state to altering the redistribution function from their assumed value of a r = 3.5. In fact the redistribution function is poorly constrained. This distribution for the break- up of large objects can be measured from asteroid families and is known to depend on impact energy, partly due to geometrical considerations (Tanga et al. 1999), but is typically found to range from a r = 3.5 − 4.5. Collisional experiments of (smaller) rocks (Fujiwara et al. 1989) give distributions that have a r = 2.8 − 5.8 (though 3.7-4.0 is more common). However, it is noted that the distributions can turn over to lower values of a r < 3.5 at fragment sizes smaller than 1mm (e.g., Flynn et al. 2009), i.e., in exactly the size range of interest, also changing to even lower values for < 100 m m. Thus, although the distribution implied by Figure 3c would appear to have a relatively small value of a r, this is not without precedent within the context of previous studies, especially for this size range. 20 3.3 Evolution in P-R regime M.C.Wyatt et al. The evolution in the P-R drag regime was discussed in Dominik & Decin (2003) who found that the dust luminosity should decrease (cid:181) 1/t2 in this regime. We find that the situation is substantially more complicated than the prescription given by these authors, since we treat an evolving size distribution in which the numbers of particles of different sizes evolve differently. Nevertheless, for certain assumptions we recover the same dependence for the evolution of luminosity. To consider the evolution in the P-R drag regime, we first need to determine the evolution of Dpr. Here we will consider the situation in which the largest particles are in equilibrium, so that the mass in all bins above Dpr are being depleted with age as 1/t. The implications for Figure 3b are that the collision lifetimes of all particles (above Dpr) increase (cid:181) t, and so the turnover evolves to larger sizes. Here we ignore the effect of the turnover on the calculation of the collision time at Dpr; this means that we miscalculate Dpr, but by a fraction that should be the same at all ages for a given a r, and so should not affect conclusions about the evolution. Thus equating collision and P-R loss rates we find that Dpr (cid:181) t(6−a)/(3+2a), (46) or Dpr (cid:181) t1.58 for the parameters of our simulation. For particles larger than Dpr, their number continues to deplete inversely with age. How- ever, for particles that are now in the P-R drag regime, the increase of Dpr with age means that they are being replenished by a population for which the size at which its bottom end is truncated is also increasing with age. The number of particles in the P-R drag regime thus depletes much faster mks(D < Dpr) (cid:181) t[a r(6−a)−30]/(3+2a), (47) or (cid:181) t −2.79 for the parameters of our simulation with a r = 3.5. The evolution that is observed then depends on what particle sizes are being probed. For example, the evolution of dust mass defined as the mass in D < 2mm particles would be expected to fall off inversely with age at early times when Dpr ≪ 2mm, but tend to a fall off (cid:181) t −2.79 by the time Dpr ≫ 2mm, and likewise for similar definitions of dust mass. However, for certain observations the contribution to the emission from the disk is dom- inated by the total cross-sectional area in the distribution. For the distributions considered here, with a r < 4, that area is dominated by particles of size ∼ Dpr. Although this size evolves in a way that depends on the strength law (equation 46), it turns out that the evolution of total cross-sectional area does not, since s tot (cid:181) mk(Dre f )D(3+2a)/(a−6) , where mk(Dre f ) is the mass in particles at some fixed reference diameter Dre f that is above Dpr. In the regime we are considering in which the largest particles have reached collisional equilib- rium mk(Dre f ) (cid:181) t −2 as predicted by Dominik & Decin (2003). Neverthless, all observations are prone to bias toward probing certain particle sizes (see, e.g., Fig. 5 of Wyatt & Dent 2002), and so one would expect all observations to tend to fall off at a rate given by equation (47) on long enough timescales, and it is important to consider the particle sizes contributing to a particular observation to understand how disk brightness is expected to evolve. t −1, and so from equation (46) we find that s tot (cid:181) pr The above points are illustrated in Figure 4, which shows the evolution from an age of 10Gyr of the planetesimal belt considered in Figure 2, but including P-R drag at all ages (note that a r = 3.5 in these simulations). Figure 4a shows how the size distribution at 10Gyr is unaffected by P-R drag (justifying its exclusion from the simulations shown in Figure 2), but that soon after this the turnover due to P-R drag appears at a size that gets larger with Debris disk size distributions 21 (a) (b) (c) Fig. 4 Evolution of the planetesimal belt considered in Figure 2 on timescales long enough for P-R drag to affect the evolution of the dust. Although these timescales are unrealistically long for this parameter set, shorter timescales may apply for different belt parameters, and dynamical depletion can result in evolutionary timescales that are longer than the system age (§4.1.2). (a) Mass (mk) in different size bins (Dk) at ages of 10, 30, 100, 300, 1000, 3000, 10000 Gyr. (b) Loss timescales for particles of different sizes at these ages due to collisions and P-R drag. (c) Evolution of total mass, dust mass (both defined as mass in particles smaller than 2 mm, and smaller than 10 m m), and cross-sectional area. 22 M.C.Wyatt et al. Table 2 Evolutionary stages for the belt discussed in Figs. 2 and 4, with the ages over which the stage lasted, the phases present in the size distribution (P=Primordial, S=Strength regime, G=Gravity regime, PR=P-R drag regime), and the dependences on time of total disc mass, dust mass and total cross-sectional area. Stage 0 I II III IV Ages 0 < 0.5Myr 0.5Myr-2Gyr 2 − 30Gyr > 30Gyr Phases P S-P S-G-P S-G PR-S-G mtot - flat gradual t −1 t −1 mdust - flat t −0.34 t −1 t −2.8 s tot - flat t −0.34 t −1 t −2 age. It is also evident that the ripple above Dpr that was noted to be absent from Figure 3 is now apparent at late ages. This is because at larger values of Dpr, particles at this transition have lower values of Xc, i.e., they are destroyed by particles that are much smaller relative to themselves than those with smaller values of Dpr, which means that the turnover appears more like a truncation to them. Figure 4b shows how the turnover size is indeed set at all ages by the location where collision and P-R drag timescales are equal. This justifies the analytical work in the last few paragraphs, which also provides excellent quantitative agreement with the evolution of mass and cross-sectional area seen in Figure 4c; indeed at late times the simulations find mtot (cid:181) t −1.0, mdust (cid:181) t −2.8, s tot (cid:181) t −2.0. A word of caution is necessary on the evolution implied by equations (46) and (47), which is that these apply when the largest objects in the distribution are in collisional equi- librium and so are being depleted (cid:181) 1/t. This may not be the case for belts that were simply born with low mass, or have undergone dynamical depletion (like the asteroid belt), or have enhanced levels of drag (e.g. §4.3). For such systems it would be necessary to consider the rate of depletion of particles with sizes above Dpr, and substitute that into the above analysis in place of the (cid:181) 1/t dependence assumed there. 3.4 Summary of evolutionary stages To summarise, if we ignore the ripples in the distribution, then the evolution of the size distribution shown by the runs of Figs. 2 and 4 show four main stages of evolution that are summarised in Table 2. The evolution of planetesimal belts with different initial parameters would be expected to show similar behaviour. However, there would be quantitative differences. For example, the ages of the transitions between the stages would depend on the initial parameters, since they correspond (e.g., Table 1) to the collisional lifetime of objects of size Dw (transition I-II), and of size D1 (transition II-III), and the age at which the collisional lifetime of objects of size Dbl is equal to their P-R drag lifetime (transition III-IV). The fall-off of dust mass and area with age during stage II also depends on the primordial size distribution and on the dispersal law in the gravity regime (see equation 43 of Lohne et al. 2008), and the dust mass in stage IV depends on the dispersal law in the strength regime and the redistribution function (equation 47). There may also be qualitative differences. For example, if the pri- mordial distribution contained no particles in the gravity regime, i.e., if D1 < Dw, then there could be no "G" phase in the size distribution, and stage II would be skipped. Similarly, if the primordial distribution was made up of different power laws, then there would be differ- ent phases associated with the timescale for collisional equilibrium to reach the transitions between the primordial power laws. And it was already noted that it may be possible for P-R (cid:181) (cid:181) (cid:181) (cid:181) (cid:181) (cid:181) (cid:181) (cid:181) Debris disk size distributions 23 drag to become important during stage II in which case an evolutionary stage involving the 4 phases "PR-S-G-P" would be reached. 4 Steady state with generalised loss processes The method of §3 for considering the effect of P-R drag on the steady state size distribution of planetesimal belts undergoing a collisional cascade can be applied to other loss processes in a similar manner. In particular, if the loss process is described by k = mkRl m−l k, (48) then its effect on the steady state size distribution can be determined numerically by extend- ing method 2 (§2.4.2) to solve mks = Ck/(Rc k + Rl k). (49) Here we consider several loss processes, and their qualitative effect on the size distribution (and its evolution), but defer a detailed treatment to later work. Treatment of other loss processes such as Yarkovsky forces and gas drag are also possible within this framework, as are erosive loss processes, such as sublimation and sputtering, by invoking the gain term m+g in equation (1) (since loss from one size bin results in gain into another). However, k such considerations are again beyond the scope of this paper. 4.1 Power law loss rate Consider the action of a loss process that can be defined by equation (48) with Rl k (cid:181) D−g l k , (50) which is a generalisation of loss by P-R drag for which g l = 1. 4.1.1 Large g l As long as g l is large enough, the analytical treatment is identical to that for P-R drag, in that there is a transition at the size Dl for which Rc k(Dl) = Rl k(Dl). (51) For D < Dl the removal of particles is dominated by the loss process, while for D > Dl removal is dominated by collisions. A consideration of the analogous Figure 3b shows that large enough here means g l > g lim where g lim is determined from a comparison of the slope in the loss regime (equation 50) with that in the collisional regime (equation 36). Assuming the transition occurs in the strength regime (i.e., Dl < Dw), equation (24) means that g lim = 4 − a a = (3 − 3a)/(6 − a), (52) which for a = 0.3 means g l > 0.37. Thus the evolution is qualitatively similar, and the size distribution in the loss regime is still indicative of the redistribution function, except that the index is now a l = a r − g l. (53) 24 M.C.Wyatt et al. k (cid:181) D4−a r The mass loss from the belt by loss processes is m−l ks m−l ks that particles in the collisional regime are being depleted (cid:181) collisional loss Rc k the transition size follows for D > Dl, but is still for D < Dl. Assuming the largest particles are in collisional equilibrium so t −1, then by equating the rate of t −1 with the rate from loss processes (equation 50), we find that a a−4 k (cid:181) D (cid:181) D g lim −g l k Dl (cid:181) 1 l −g g t lim . (54) For similar reasoning to that used to derive equation (47), this means that the mass in small particles below the transition evolves mks(D < Dl) (cid:181) t 4.1.2 Dynamical depletion a r −4−2(g l −g g l −g lim lim ) . (55) However, if g l < g lim, then it is not small particles that are most affected by loss processes, but particles above Dl. This is the situation for most dynamical loss processes for which loss rates are independent of size and so g l = 0. The effect on the size distribution is more complicated in this case. For example, consider the belt of Figure 2 that has evolved to tdep = 0.5Gyr without dynamical mass loss, but then undergoes impulsive and significant dynamical depletion by a factor fdep. This would cause the size distribution to be retained, but to drop at all sizes by the same fraction. This would mean that additional evolutionary time would be needed for the smallest objects in the primordial distribution (∼ 27km in diameter) to come to collisional equilibrium and continue to evolve. In the meantime the size distribution in the portion that was originally in steady state (D < 27km) would be retained since the depletion would not have affected the balance of mass gain or loss from such bins. This example illustrates two ways in which dynamical depletion can affect the size dis- tribution of what appears to be primordial at late times: some parts of the distribution may represent a scaled down version of the original primordial distribution, whereas others may represent the steady state distribution attained during an earlier high mass phase of evolu- tion, even though their collisional lifetime is longer than the current system age. This point was recognised by Bottke et al. (2005), who showed how dynamical depletion can be mod- elled by considering collisional evolution that takes place over timescales longer than the system age. This is in agreement with our analysis, since we expect the size distribution of the above example following the depletion to look like that expected for a primordial dis- tribution that is scaled down by fdep then evolved for a time 1/ fdep times longer than the current system age. As noted by Bottke et al., it is not possible to infer fdep from the current distribution (of the asteroid belt in their application), i.e., we cannot determine if the original distribution was very massive and short-lived, or less massive and longer-lived. However, it is possible to model the evolution of the size distribution for depletion by a known amount fdep at a known time tdep. 4.2 Realistic radiation forces The prescription for the effect radiation forces in the preceding sections is an approximation of the true physics. For example, the truncation of the size distribution by radiation pressure, which occurs for particles with D < Dbl is implemented by simply not including such bins in the analysis. However, a more realistic approach would be to include radiation pressure as Debris disk size distributions 25 a loss process with an associated (short, i.e., less than orbital) timescale for bins below Dbl. Thus it would be possible to account for the possibility that the blow-out population collides with bound grains (Krivov et al. 2000), and for a gradual rather than sharp truncation due to a finite eccentricity of the parent particles (which means that b > 0.5 particles can remain bound) and/or due to a range of particle properties. It was already noted that the increase in eccentricity for particles close to Dbl is not accounted for in this model, but again this could be incorporated using an appropriately modified intrinsic collision probability Pik in equation (7) (e.g., Krivov et al. 2005; Wyatt et al. 2010). b , and although b Another approximation is the inverse linear dependence of the P-R drag loss rate on par- (cid:181) 1/D for large black body particles, this function ticle size. In fact Rpr k peaks at 0.1-1 m m particles (i.e., at particle sizes that are comparable to the dominant wave- length of stellar radiation), turning over for smaller particles and tending to b independent of size for particles ≪ 0.1 m m (e.g., Gustafson 1994). For particles around high luminosity stars this turnover is inconsequential, since it occurs at sizes below Dbl. However, for low luminosity stars it is so important that there may be no particles that can be removed by radiation pressure, i.e., b < 0.5 for all particle sizes (e.g., Sheret et al. 2004). The conse- quence of this is that, in the absence of other loss processes, P-R drag is the only removal mechanism for dust from the cascade, and so it is important to consider the effect of the turnover in b vs D on the steady state size distribution. A reconsideration of equation (40) shows that the size distribution in this regime should really be mks (cid:181) D4−a r k /b . (56) This goes some way to explaining the shape of the size distribution seen in Fig. 4 of Rei- demeister et al. (2010), since equation (56) would predict a distribution of drag dominated particles that is fairly flat in the 0.04-10 m m size range with a dip at ∼ 0.4 m m. 4.3 Stellar wind Although the discussion has focussed on radiation forces, it is notable that P-R drag only be- comes important for debris disks that are so low in density that they are no longer detectable with current instrumentation (Dominik & Decin 2003; Wyatt 2005). However, stellar wind forces act in exactly the same way as radiation forces. Their pressure component (which is analogous to radiation pressure) can usually be ignored, but their drag component (which is analogous to P-R drag) can significantly exceed P-R forces (e.g., Plavchan et al. 2005; Reidemeister et al. 2010). Thus stellar wind drag can be considered as an enhanced form of P-R drag using the same analysis but with a constant Asw in equation (33) that means its effect on the size distribution can become apparent at much younger evolutionary ages (i.e., for much higher, and potentially detectable, disc masses) than its radiation counterpart. As per the discussion of §4.2, it may also be the dominant loss mechanism for the belts of low luminosity stars. No change in the qualitative picture of disc evolution presented in §3.4 is necessary, except to note that the turnover in the size distribution may become apparent in earlier evolutionary phases (e.g., in phase II as already noted in §3.4). 5 Conclusion In this paper we presented a new scheme for determining the shape of the size distribution, and its evolution, for collisional cascades of planetesimals undergoing destructive collisions (cid:181) 26 M.C.Wyatt et al. and loss processes like Poynting-Robertson drag. This scheme, outlined in §2, considers mass gain and loss in different size bins, and treats the steady state portion of the cascade by solving for the assumption that mk = 0, and assumes that larger objects retain their primor- dial distribution. Many results can be obtained analytically, providing insight into the origin of the shape of the size distribution. For example, for particles in the steady state collisional regime this prescription was shown to lead to mass loss rate being independent of size in log- arithmic size bins; thus the steady state size distribution is independent of mass gain into the bins in this regime. However, perhaps most notably this scheme provides an efficient and readily implemented numerical method for calculating the size distribution. Although the scheme as presented assumes a simple scale independent redistribution function that defines the average mass distribution of fragments produced in collisional destruction of objects of a given size, and considers only the size distribution in one radial bin, it may be expanded to more complex redistribution functions, and to include a radial dimension (or to include dimensions of orbital elements). Our scheme, and its application to consider the evolution of the size distribution in the collision-dominated regime, was shown to reproduce qualitatively all of the features found by more detailed (and computationally more time-consuming) models that step size distri- butions forward in time. Such features in the size distribution include ripples above the ra- diation pressure blow-out limit, the change in slope at the transition from strength to gravity regime, the ripple above this transition, and the transition to primordial regime. For evo- lution, the scheme reproduces the initially slow fall-off of dust mass with time (cid:181) t −0.34 followed by steeper decline (cid:181) t −1, as well as an evolution of total disk mass that is much flatter in the initial phases but also tends to (cid:181) t −1 at late ages. One of the main purposes of the paper was to consider the effect of loss processes on the steady state size distribution, which was done in detail for P-R drag in §3. This was found, both analytically and confirmed numerically, to cause a turnover at small sizes to a size distribution that is set by the redistribution function. This means that information about the redistribution function may be recovered by measuring the size distribution of particles undergoing loss by P-R drag. This was illustrated by comparison with the size distribution of particles accreted by the Earth, though it was noted that further work on the collisional evolution of material inside the source belt and on the accretion efficiency as a function of particle size would be needed before definitive conclusions on the redistribution function could be reached. The evolution of the size distribution when P-R drag is included is sum- marised in §3.4 and Table 2. Notably it is found that although cross-sectional area drops with age (cid:181) t −2.8, underlining the importance of understanding which particle sizes contribute to an observation when considering how disk detectability evolves. t −2 in the PR-dominated regime, dust mass falls (cid:181) Other loss processes are readily incorporated into this scheme. In §4, we also discussed loss rates that can be defined by a power law in particle size, including dynamical depletion, a more realistic prescription for radiation loss rates that are not simply inversely proportional to paticle size, and stellar wind drag. Steep power law loss rates lead to size distributions and evolutions that are directly analogous to P-R drag (but with modified exponents), but loss rates that are less biased to small particles are harder to treat. However, the application to dynamical depletion agrees with the results in the literature. The way in which P-R drag becomes independent of size for the smallest particles causes a ripple and flattening of this distribution at such sizes in the P-R drag regime. The application to stellar wind drag results in distributions and evolutions that are the same as those for P-R drag, but if this effect becomes important at all, it does so at higher disc masses meaning that it has the potential to be relevant for disks that are detectable with current technology. Debris disk size distributions 27 There are now several surveys for debris disks that show how disk brightness drops with age (see review in Wyatt 2008), and further surveys are currently underway (e.g., Matthews et al. 2010). Interpretation of these surveys requires efficient schemes so that populations of large numbers of debris disks can be modelled (e.g., Wyatt et al. 2007b; Lohne et al. 2008). As such surveys push to lower disk masses, it is also important to be able to incorporate consideration of loss processes. The results presented here suggest that when disks drop to a level at which P-R drag becomes important, which is close to the limit of detectability with Spitzer (see Fig. 4 of Wyatt et al. 2007a), their brightness begins to drop rapidly, as fast as (cid:181) t −2.8. Such a fast fall off is also expected for disks affected by stellar wind drag, which may occur for disks well above the current detectability threshold. In short, this paper provides an efficient model with which to interpret observations of the debris disks of nearby stars, and their evolution, including in regimes where loss processes are important for small (or large) particles. Acknowledgments The authors are grateful to the Isaac Newton Institute for Mathematical Sciences in Cam- bridge, where the initial work on this paper was carried out during the Dynamics of Discs and Planets research programme, and to Torsten Lohne for discussions on his ACE simula- tions. We are also grateful to the reviewers for their careful reading of the manuscript and to Alexander Krivov for providing the results of the ACE simulation presented in Figure 2. Mark Booth is grateful to UK STFC and to CSA for financial support. References 1. Backman D. E. and Paresce F.: 1993, 'Main-sequence stars with circumstellar solid material - The VEGA phenomenon', In Levy E. H. and Lunine J. I. (eds), Protostars and Planets III, Univ. of Arizona Press, Tucson, pp. 1253 -- 1304. 2. Beichman C. A., et al.: 2006, 'New Debris Disks around Nearby Main-Sequence Stars: Impact on the Direct Detection of Planets', Astrophys. J. 652, 1674 -- 1693. 3. Belyaev M. and Rafikov R.: 2011, 'Non-Power Law Behavior in Fragmentation Cascades', Icarus, submitted, arXiv:1012.3220. 4. Benz W. and Asphaug E.: 1999, 'Catastrophic Disruptions Revisited', Icarus, 142, 5 -- 20. 5. Birnstiel T., Ormel C. W. and Dullemond C. P.: 2011, 'Dust size distributions in coagula- tion/fragmentation equilibrium: numerical solutions and analytical fits', Astron. Astrophys., 525, A11. 6. Bottke W. F., Durda D. D., Nesvorny D., Jedicke R., Morbidelli A., Vokrouhlicky D. and Levison H.: 2005, 'The fossilized size distribution of the main asteroid belt', Icarus, 175, 111 -- 140. 7. Burns J. A., Lamy P. L. and Soter S.: 1979, 'Radiation forces on small particles in the solar system', Icarus, 40, 1 -- 48. 8. Campo-Bagatin A., Cellino A., Davis D. R., Farinella P. and Paolicchi P.: 1994, 'Wavy size distribu- tions for collisional systems with a small-size cutoff', Planet. Space Sci., 42, 1079 -- 1092. 9. Dermott S. F., Jayaraman S., Xu Y.L., Gustafson B. A. S. and Liou J. C.: 1994, 'A circumsolar ring of asteroidal dust in resonant lock with the Earth', Nature, 369, 719 -- 723. 10. Dermott S. F., Grogan K., Durda D. D., Jayaraman S., Kehoe T. J. J., Kortenkamp S. J. and Wyatt M.C.: 2001, 'Orbital evolution of interplanetary dust', In Grun E. et al. (eds), Interplanetary Dust, Springer-Verlag, Berlin, pp. 569 -- 639. 11. Dohnanyi J. S.: 1969, 'Collisional Model of Asteroids and Their Debris', J. Geophys. Res., 74, 2531 -- 2554. 12. Dominik C. and Decin G.: 2003, 'Age Dependence of the Vega Phenomenon: Theory', Astrophys. J., 598, 626 -- 635. 13. Durda D. D. and Dermott S. F.: 1997, 'The Collisional Evolution of the Asteroid Belt and Its Contri- bution to the Zodiacal Cloud', Icarus, 130, 140 -- 164. 28 M.C.Wyatt et al. 14. Durda D. D., Greenberg R. R. and Jedicke R.: 1998, 'Collisional Models and Scaling Laws: A New Interpretation of the Shape of the Main-Belt Asteroid Size Distribution', Icarus, 135, 431 -- 440. 15. Flynn G. J., Durda D. D., Sandel L. E., Kreft J. W. and Strait M. M.: 2009, 'Dust production from the hypervelocity impact disruption of the Murchison hydrous CM2 meteorite: Implications for the disruption of hydrous asteroids and the production of interplanetary dust', Planet. Space Sci., 57, 119 -- 126. 16. Fraser W. C.: 2009, 'The Collisional Divot in the Kuiper Belt Size Distribution', Astrophys. J., 706, 119 -- 129. 17. Fuentes C. I., Holman M. J., Trilling D. E. and Protopapas P.: 2010, 'Trans-Neptunian Objects with Hubble Space Telescope ACS/WFC', Astrophys. J., 722, 1290 -- 1302. 18. Fujiwara A., Cerroni P., Davis D. R., Ryan E., di Martino M.: 1989, 'Experiments and scaling laws for catastrophic collisions', In Binzel R. P. et al. (eds), Asteroids II, Univ. of Arizona Press, Tucson, pp. 240 -- 265. 19. Gladman B. J., et al.: 2009, 'On the asteroid belt's orbital and size distribution', Icarus, 202, 104 -- 118. 20. Gor'kavyi N. N., Ozernoy L. M., Mather J. C. and Taidakova T.: 1997, 'Quasi-Stationary States of Dust Flows under Poynting-Robertson Drag: New Analytical and Numerical Solutions', Astrophys. J., 488, 268 -- 276. 21. Grogan K., Dermott S.F. and Durda D.D.: 2001, 'The Size-Frequency Distribution of the Zodiacal Cloud: Evidence from the Solar System Dust Bands', Icarus, 152, 251 -- 267. 22. Grun E., Zook H. A., Fechtig H. and Giese R. H.: 1985, 'Collisional balance of the meteoritic com- plex', Icarus, 62, 244 -- 272. 23. Gustafson, B. A. S.: 1994, 'Physics of Zodiacal Dust', Annu. Rev. Earth Planet. Sci., 22, 553 -- 595. 24. Heng K. and Tremaine S.: 2009, 'Long-lived planetesimal discs', Mon. Not. Roy. Astron. Soc., 401, 867 -- 889. 25. Ishimoto H.: 2000, 'Modeling the number density distribution of interplanetary dust on the ecliptic plane within 5AU of the Sun', Astron. Astrophys., 362, 1158 -- 1173. 26. Kobayashi H. and Tanaka H.: 2010, 'Fragmentation model dependence of collision cascades', Icarus, 206, 735 -- 746. 27. Kortenkamp S. J. and Dermott S. F.: 1998, 'Accretion of Interplanetary Dust Particles by the Earth', Icarus, 135, 469 -- 495. 28. Krivov A. V., Lohne T. and Sremcevic M.: 2006, 'Dust distributions in debris disks: effects of gravity, radiation pressure and collisions', Astron. Astrophys., 455, 509 -- 519. 29. Krivov A. V., Mann I. and Krivova N. A.: 2000, 'Size distributions of dust in circumstellar debris discs', Astron. Astrophys., 362, 1127 -- 1137. 30. Krivov A. V., Sremcevic M. and Spahn F.: 2005, 'Evolution of a Keplerian disk of colliding and fragmenting particles: a kinetic model with application to the Edgeworth Kuiper belt', Icarus, 174, 105 -- 134. 31. Kuchner M. J. and Stark C. C.: 2010, 'Collisional Grooming Models of the Kuiper Belt Dust Cloud', Astron. J., 140, 1007 -- 1019. 32. Leinert C. and Grun E.: 1990, 'Interplanetary Dust', In Schewenn R. and Marsch E. (eds), Physics of the Inner Heliosphere I, Springer-Verlag, Germany, pp. 207 -- 256. 33. Leinert C., Roser S., Buitrago J.: 1983, 'How to maintain the spatial distribution of interplanetary dust', Astron. Astrophys., 118, 345 -- 357. 34. Lohne T.: 2008, 'Models of rotationally symmetric, collision-dominated debris discs', PhD Disserta- tion, Friedrich Schiller University of Jena. 35. Lohne T., Krivov A. V. and Rodmann J.: 2008, 'Long-Term Collisional Evolution of Debris Disks', Astrophys. J., 673, 1123 -- 1137. 36. Love S. G. and Brownlee D. E.: 1993, 'A Direct Measurement of the Terrestrial Mass Accretion Rate of Cosmic Dust', Science, 262, 550-553. 37. Matthews B. C., et al.: 2010, 'Resolving debris discs in the far-infrared: Early highlights from the DEBRIS survey', Astron. Astrophys., 518, L135 -- L140. 38. Morbidelli A., Bottke W. F., Nesvorn´y D. and Levison H. F.: 2009, 'Asteroids were born big', Icarus, 204, 558 -- 573. 39. Nesvorn´y D., Jenniskens P., Levison H. F., Bottke W. F., Vokrouhlick´y D. and Gounelle M.: 2010, 'Cometary Origin of the Zodiacal Cloud and Carbonaceous Micrometeorites. Implications for Hot Debris Disks', Astrophys. J., 713, 816 -- 836. 40. O'Brien D. P. and Greenberg R.: 2003, 'Steady-state size distributions for collisional populations: analytical solution with size-dependent strength', Icarus, 164, 334 -- 345. 41. Plavchan P., Jura M. and Lipsky S. J.: 2005, 'Where Are the M Dwarf Disks Older Than 10 Million Years?', Astrophys. J., 631, 1161 -- 1169. Debris disk size distributions 29 42. Reidemeister M., Krivov A. V., Stark C. C., Augereau J.-C., L'"ohne T. and Muller S.: 2011, 'The cold origin of the warm dust around epsilon Eridani', arXiv:1011.4882. 43. Schlichting H. E., Ofek E. O., Wenz M., Sari R., Gal-Yam A., Livio M., Nelan E., Zucker S.: 2009, 'A single sub-kilometre Kuiper belt object from a stellar occultation in archival data', Nature, 462, 895 -- 897. 44. Sheret I., Dent W. R. F. and Wyatt M. C.: 2004, 'Submillimetre observations and modelling of Vega- type stars', Mon. Not. Roy. Astron. Soc. 348, 1282 -- 1294. 45. Stark C. C. and Kuchner M. J.: 2009, 'A New Algorithm for Self-consistent Three-dimensional Mod- eling of Collisions in Dusty Debris Disks', Astrophys. J., 707, 543 -- 553. 46. Stewart S. T. and Leinhardt Z. M.: 2009, 'Velocity-Dependent Catastrophic Disruption Criteria for Planetesimals', Astrophys. J., 691, L133 -- L137. 47. Tanaka H., Inaba S. and Nakazawa K.: 1996, 'Steady-State Size Distribution for the Self-Similar Collision Cascade', Icarus, 123, 450 -- 455. 48. Tanga P., Cellino A., Michel P., Zappal´a V., Paolicchi P. and dell'Oro A.: 1999 'On the Size Distri- bution of Asteroid Families: The Role of Geometry', Icarus, 141, 65 -- 78. 49. Th´ebault P. and Augereau J.-C.: 2007, 'Collisional processes and size distribution in spatially ex- tended debris discs', Astron. Astrophys., 472, 169 -- 185. 50. Th´ebault P., Augereau J.-C. and Beust H.: 2003, 'Dust production from collisions in extrasolar plan- etary systems. The inner beta Pictoris disc', Astron. Astrophys., 408, 775 -- 788. 51. Vitense Ch., Krivov A. V. and Lohne T.: 2010, 'The Edgeworth-Kuiper debris disk', Astron. Astro- phys., 520, A32. 52. Wetherill G. W.: 1967, 'Collisions in the Asteroid Belt', J. Geophys. Res., 72, 2429 -- 2444. 53. Wyatt M. C.: 2005, 'The insignificance of P-R drag in detectable extrasolar planetesimal belts', As- tron. Astrophys., 433, 1007 -- 1012. 54. Wyatt M. C.: 2008, 'Evolution of Debris Disks', Annu. Rev. Astron. Astrophys., 46, 339 -- 383. 55. Wyatt M. C. and Dent W. R. F.: 2002, 'Collisional processes in extrasolar planetesimal discs - dust clumps in Fomalhaut's debris disc', Mon. Not. Roy. Astron. Soc., 334, 589 -- 607. 56. Wyatt M. C., Booth M., Payne M. J., Churcher L. J.: 2010, 'Collisional evolution of eccentric plan- etesimal swarms', Mon. Not. Roy. Astron. Soc., 402, 657 -- 672. 57. Wyatt M. C., Dermott S. F., Telesco C. M., Fisher R. S., Grogan K., Holmes E. K. and Pina R. K.: 1999, 'How Observations of Circumstellar Disk Asymmetries Can Reveal Hidden Planets: Pericenter Glow and Its Application to the HR 4796 Disk', Astrophys. J., 527, 918-944. 58. Wyatt M. C., Smith R., Beichman C. A., Bryden G., Greaves J. S. and Lisse C. M.: 2007a, 'Transience of Hot Dust around Sun-like Stars', Astrophys. J., 658, 569 -- 583. 59. Wyatt M. C., Smith R., Su K. Y. L., Rieke G. H., Greaves J. S., Beichman C. A. and Bryden G.: 2007b, 'Steady State Evolution of Debris Disks around A Stars', Astrophys. J., 663, 365 -- 382.
1910.04111
1
1910
2019-10-09T16:46:00
Stability of Nitrogen in Planetary Atmospheres in Contact with Liquid Water
[ "astro-ph.EP" ]
Molecular nitrogen is the most commonly assumed background gas that supports habitability on rocky planets. Despite its chemical inertness, nitrogen molecule is broken by lightning, hot volcanic vents, and bolide impacts, and can be converted into soluble nitrogen compounds and then sequestered in the ocean. The very stability of nitrogen, and that of nitrogen-based habitability, is thus called into question. Here we determine the lifetime of molecular nitrogen vis-a-vis aqueous sequestration, by developing a novel model that couples atmospheric photochemistry and oceanic chemistry. We find that HNO, the dominant nitrogen compounds produced in anoxic atmospheres, is converted to N2O in the ocean, rather than oxidized to nitrites or nitrates as previously assumed. This N2O is then released back into the atmosphere and quickly converted to N2. We also find that the deposition rate of NO is severely limited by the kinetics of the aqueous-phase reaction that converts NO to nitrites in the ocean. Putting these insights together, we conclude that the atmosphere must produce nitrogen species at least as oxidized as NO2 and HNO2 to enable aqueous sequestration. The lifetime of molecular nitrogen in anoxic atmospheres is determined to be >1 billion years on temperate planets of both Sun-like and M dwarf stars. This result upholds the validity of molecular nitrogen as a universal background gas on rocky planets.
astro-ph.EP
astro-ph
DRAFT VERSION OCTOBER 10, 2019 Typeset using LATEX preprint2 style in AASTeX62 Stability of Nitrogen in Planetary Atmospheres in Contact with Liquid Water∗ RENYU HU1, 2 AND HECTOR DELGADO DIAZ1 1Jet Propulsion Laboratory California Institute of Technology Pasadena, CA 91109, USA 2Division of Geological and Planetary Sciences California Institute of Technology Pasadena, CA 91125, USA (Received XXX; Revised XXX; Accepted XXX) Submitted to ApJ ABSTRACT Molecular nitrogen is the most commonly assumed background gas that supports habitabil- ity on rocky planets. Despite its chemical inertness, nitrogen molecule is broken by lightning, hot volcanic vents, and bolide impacts, and can be converted into soluble nitrogen compounds and then sequestered in the ocean. The very stability of nitrogen, and that of nitrogen-based habitability, is thus called into question. Here we determine the lifetime of molecular nitro- gen vis-à-vis aqueous sequestration, by developing a novel model that couples atmospheric photochemistry and oceanic chemistry. We find that HNO, the dominant nitrogen compounds produced in anoxic atmospheres, is converted to N2O in the ocean, rather than oxidized to nitrites or nitrates as previously assumed. This N2O is then released back into the atmosphere and quickly converted to N2. We also find that the deposition rate of NO is severely limited by the kinetics of the aqueous-phase reaction that converts NO to nitrites in the ocean. Putting these insights together, we conclude that the atmosphere must produce nitrogen species at least as oxidized as NO2 and HNO2 to enable aqueous sequestration. The lifetime of molecu- lar nitrogen in anoxic atmospheres is determined to be > 1 billion years on temperate planets of both Sun-like and M dwarf stars. This result upholds the validity of molecular nitrogen as a universal background gas on rocky planets. Keywords: Extrasolar rocky planets -- Habitable planets -- Super Earths -- Exoplanet at- mospheric composition -- Exoplanet evolution 1. INTRODUCTION Corresponding author: Renyu Hu [email protected] ∗ c(cid:13)2019. California Institute of Technology. Government sponsorship acknowledged. Nitrogen is the bulk constituent of Earth's at- mosphere and a common constituent of the atmo- spheres of rocky planets in the Solar System. The universality of nitrogen has been extended to extra- solar rocky worlds, as molecular nitrogen (N2) is generally assumed as the background gas in the at- 2 HU & DELGADO mosphere. The standard picture of habitable plan- ets of stars (Kasting et al. 1993) posits that climate and geologic processes on rocky planets regulate the abundance of atmospheric CO2 to maintain a surface temperature that is consistent with liquid water oceans -- but an often overlooked ingredient of this picture is a constant, approximately 1 bar, N2-dominated background atmosphere. The climate-maintaining effect of N2 primarily stems from its higher volatility than CO2 or H2O. As the partial pressure of CO2 is controlled by the silicate weathering cycle (Walker et al. 1981), and that of H2O is controlled by the surface tem- perature, the partial pressure of N2 is not a di- rect function of any climatological parameters. Having a sizable N2 atmosphere, therefore, al- leviates the sensitivity of the planetary climate to subtle changes in forcings, and thus widens the semi-major axis ranges in which the planet can be habitable (Vladilo et al. 2013). No hab- itable climate can be found if the partial pres- sure of N2 is less than 0.015 bar (Vladilo et al. 2013). The actual lower-limit may be even higher, as it is later found that N2 as a non-condensable gas maintains the cold trap of the middle atmo- sphere and prevents water from loss to space (Wordsworth & Pierrehumbert 2013). While N2 is not a greenhouse gas, it causes a warming effect on climate via pressure broadening of CO2 and H2O absorption features (Goldblatt et al. 2009). Due to its strong triple bond, N2 is very close to being chemically inert in the atmosphere. The processes that can break N2 are peculiar (Mancinelli & McKay 1988): on today's Earth it is primarily performed by microbes, and be- fore the rise of nitrogen-fixation microbes it is done in energetic events including lightning (Yung & McElroy 1979; Kasting & Walker 1981; Navarro-González et al. 1998; Wong et al. 2017), bolide impact (McKay et al. 1988), and also hot volcanic vents (Mather et al. 2004). The immedi- ate product of the "atmospheric nitrogen fixation" is NO, and the NO is then converted to HNO3 in -- and NO3 oxygen-rich atmospheres and to HNO in oxygen- poor ones (Kasting & Walker 1981; Wong et al. It has been suggested that the HNO is 2017). -- in the ocean then converted to NO2 (Mancinelli & McKay 1988; Summers & Khare 2007). As such, NO produced in the atmosphere eventually becomes nitrites and nitrates. The en- tire 1-bar N2-dominated atmosphere could be se- questered in the ocean as nitrites and nitrates -- thus creating a potential problem for the stabil- ity of a nitrogen-dominated atmosphere in contact with liquid water oceans. We are therefore motivated to determine the life- time of N2 -- and thus that of N2-based habitability -- on a habitable exoplanet. We focus on anoxic planets without life, because microbes would be able to harvest the nitrites and nitrates in the oceans, reduce them to N2 or N2O, and restore the N2 stability. Without life, the formation of ni- trites and nitrates may well be mostly one-way and become long-term losses of nitrogen. In this paper we calculate the kinetic timescale of this process. We first study the fate of HNO, the dominant ni- trogen compound produced in anoxic atmospheres, when it is deposited into the ocean. We show that HNO does not lead to nitrogen sequestration but rather formation of N2O (Section 2). We then present a novel model that couples an atmosphere photochemistry model (Hu et al. 2012; Hu et al. 2013) and an ocean aqueous-chemistry model, so that the rates of transfer between the atmosphere and the ocean can be self-consistently calculated (Section 3). Using the coupled model we deter- mine the lifetime of N2 in anoxic atmospheres on temperate planets of Sun-like and M dwarf stars (Section 4). We discuss the implications of our findings in Section 5 and conclude in Section 6. 2. AQUEOUS CHEMISTRY OF HNO ON PLANETS 2.1. Aqueous-Phase Reactions and Kinetic Rates HNO is the main atmospheric product of nitro- gen compounds under anoxic conditions, and its LIFETIME OF NITROGEN 3 fate in the ocean has not been clarified. The aque- ous chemistry of HNO, and its conjugate base NO -- , is peculiar because the ground state of HNO is a singlet while that of NO -- is a triplet. This makes the deprotonation reaction to proceed as the forward direction of HNO + OH− −−−−⇀↽−−−− NO− + H2O (R1) a slow, second-order reaction (Miranda 2005). Un- der the pH conditions relevant to planets, most of the dissolved HNO exists in the form of HNO. We note that the excited state HNO is a triplet and it quickly dissociates to NO -- . The transition to the excited state, however, is spin forbidden and has not been observed in experiments. Dissolved HNO can be removed by rapid dehy- drative dimerization HNO + HNO −−−−→ N2O + H2O (R2) with its rate constant determined by the flash pho- tolysis technique (Shafirovich & Lymar 2002). NO -- is rapidly oxidized to nitrate when free oxy- gen is available NO− + O2 −−−−→ ONOO− −−−−→ NO3 − (R3) or polymerized by NO via NO− + NO −−−−⇀↽−−−− N2O2 − (R4) − N2O2 + NO −−−−→ N3O3 (R5) − N3O3 −−−−→ N2O + NO2 − (R6) The polymerization can also start from HNO HNO + NO −−−−⇀↽−−−− HN2O2 (R7) HN2O2 + NO −−−−→ HN3O3 (R8) HN3O3 −−−−→ N2O + HNO2 (R9) Both polymerization reactions eventually form N2O and nitrite, and their rate constants have been measured using pulse radiolysis and NO-rich flu- ids (Gratzel et al. 1970; Seddon et al. 1973). These polymerization pathways have been adopted as the pathway to convert HNO to nitrite and nitrate in planetary oceans (Mancinelli & McKay 1988; Summers & Khare 2007; Wong et al. 2017). In summary, the removal pathways of HNO in the aqueous phase are dehydrative dimerization (Reaction R2), deprotonation (Reaction R1) fol- lowed by either oxidation (Reaction R3) or poly- merization (Reactions R4-R6), and direct polymer- ization (Reactions R7-R9). Relevant rate constants are tabulated in Table 1. Rate Constant Reaction R1 forward 5 × 104 M−1 s−1 R1 reverse R2 1.2 × 102 s−1 8 × 106 M−1 s−1 4 × 109 M−1 s−1 R3 R4 forward 2 × 109 M−1 s−1 R4 reverse 3 × 104 s−1 3 × 106 M−1 s−1 R5 R7 forward 2 × 109 M−1 s−1 R7 reverse R8 R10 R11 8 × 106 s−1 8 × 106 M−1 s−1 2 × 108 M−1 s−1 1 × 108 M−1 s−1 Table 1. Rate constants for HNO, NO, and NO2 re- actions in the aqueous phase. Compiled from Miranda (2005) and Lee (1984). The rate constants are provided at the room temperature. 2.2. Reaction Rates under Planetary Conditions Using the kinetic constants from experiments, we calculate the reaction rates of the HNO removal pathways under typical planetary conditions. After deprotonation, NO -- can be either oxidized (Reaction R3) or polymerized (Reactions R4-R6). We first compare the two sub-pathways. The rate 4 HU & DELGADO of Reaction (R3) is RR3 = kR3[NO−][O2], (1) and the overall rate of Reactions (R4-R6) is RR4−R6 = kR4 f [NO−][NO] kR5[NO] kR4r + kR5[NO] , (2) where the additional f and r in the subscript denote the rate constant of the forward and the reverse di- rections, respectively, and quantities in [X] denote the concentration of the species X in the aqueous phase, usually in the unit of M (i.e., mole per liter). The reaction rate R has the unit of M s−1. Species Typical Upper Limit 3 × 10−6 1 × 10−8 fNO [NO] fO2 [O2] fHNO [HNO] 4 × 10−11 8 × 10−14 M 6 × 10−9 M 2 × 10−15 2 × 10−18 M 1 × 10−11 M 2 × 10−11 2 × 10−10 M 2 × 10−9 M 2 × 10−10 Table 2. Typical and anoxic upper-limit concentrations for evaluating and comparing the reaction rates of HNO removal pathways. For each gas (X), the mixing ratio at the bottom of the atmosphere ( fX) and the concentration in the surface ocean ([X]) are provided. These quanti- ties are consistent with the converged photochemistry models shown in Section 4. For a typical anoxic condition, [NO] ∼ 8 × 10−14 M and [O2] ∼ 2 × 10−18 M (Table 2). These quanti- ties are from the atmospheric photochemistry mod- els under terrestrial lightning rates (Section 4) and have factored in the Henry's law constants for re- spective gases. When the lightning rate is very high (i.e., 100× the terrestrial rate), the upper lim- its are [NO] ∼ 6 × 10−9 M and [O2] ∼ 10−11 M. Note that these upper limits do not include the oxygen-rich scenarios that would be produced on planets of M dwarf stars (see Section 4). Based on these concentrations, RR4−R6 = 10−15 ∼ 10−10[NO−] M s−1 and RR3 = 10−8 ∼ 4 × 10−2[NO−] M s−1. Therefore, even under the anoxic condi- tions, RR3 ≫ RR4−R6, and the same is true for oxygen-rich conditions. The overall rate of the re- moval path starting with deprotonation (Reaction R1) is thus RR1 = kR1 f [HNO][OH−] kR3[O2] kR1r + kR3[O2] . (3) The rate of dehydrative dimerization (Reaction R2) is RR2 = kR2[HNO][HNO], (4) and the overall rate of direct polymerization (Re- actions R7-R9) is RR7−R9 = kR7 f [HNO][NO] kR8[NO] kR7r + kR8[NO] . (5) Under typical and limiting anoxic conditions, [HNO] ∼ 2 × 10−10 − 2 × 10−9 M (Section 4). For a neutral pH, we estimate RR1 = 7 × 10−23 ∼ 3 × 10−15 M s−1, RR2 = 3 × 10−13 ∼ 3 × 10−11 M s−1, and RR7−R9 = 3 × 10−27 ∼ 10−16 M s−1. Com- paring the three rates, we have RR2 ≫ RR1 > RR7−R9. RR1 is proportional to the concentration of OH -- in the ocean, and for RR1 to be greater than RR2, the ocean must be highly alkaline with pH > 11. Such a pH value is well higher than the pH of Earth's ocean currently or in the Archean (Halevy & Bachan 2017; Krissansen-Totton et al. 2018). Therefore, under anoxic conditions relevant for planetary atmospheres, dehydrative dimeriza- tion (Reaction R2) is the dominant removal path- way of HNO deposited in the ocean. Under oxygen-rich conditions, including in the oxygen-rich atmospheres produced by CO2 pho- tolysis on planets of M dwarf stars (see Section 4), little HNO is produced in the atmosphere, and thus the dissolved concentration is very small. In this case, the rate of Reaction (R2) is very small, and deprotonation followed by oxidation (Reac- tions R1 and R3) dominates. However, that HNO oxidation pathway is still not important to the over- all removal flux of nitrogen, because little HNO is produced in the atmosphere in the first place. LIFETIME OF NITROGEN 5 2.3. Consistency with Summers & Khare (2007) The main finding of this section is that under planetary conditions the deposited HNO in the ocean does not mainly become nitrite or nitrate. This finding might be perceived as contradictory to the experimental result of Summers & Khare (2007), where a gas mixture of CO2 and N2 with 1% NO and 1% CO in contact with liq- uid water was irradiated by ultraviolet light. Summers & Khare (2007) found that nitrate and nitrite to a lesser extent were formed and the NO was depleted in approximately 1 hour. A smaller amount of N2O was also produced. The interpre- tation was that HNO was formed and dissolved, and Reactions (R4-R6) or Reactions (R7-R9) took place dominantly in the system. The experimental result of Summers & Khare (2007) is consistent with our model of the ki- netics of HNO aqueous chemistry, as it show- cases the outcome from a NO-rich fluid. The experimental vessel was filled to a pressure of approximately 1 bar, which means that in equi- librium [NO] ∼ 2 × 10−5 M. Therefore, the fluid was more NO-rich than planetary oceans by or- ders of magnitude. Applying this concentration and re-evaluating all reaction rates in this section, we find that the rate of Reaction (R1) followed by Reactions (R4-R6) is 2 × 10−3[HNO] M s−1, the rate of Reactions (R7-R9) is 8 × 10−1[HNO] M s−1, and the rate of Reaction (R2) is 8 × 106[HNO]2 M s−1. The concentration of HNO in the system is unknown, but NO has a lifetime of 1 hour and yet [HNO] has a lifetime of at most ∼ 1 s. As an upper limit, we assume that HNO is the only intermediary in the removal of NO and that all HNO in the system (a 110-ml gas cell) is in the aqueous phase (15-ml water, Summers & Khare (2007)). We estimate [HNO] < 7 × 10−7 M. To- gether, we find that even at this upper limit, the reaction rate of direct polymerization (Reactions R7-R9) is on the same order of magnitude as the reaction rate of dehydrative dimerization (Reaction R2). In reality, the concentration of HNO should be smaller and polymerization becomes the dom- inant pathway, with the N2O-producing dimeriza- tion the secondary pathway. This is what was seen in the experiment, and our kinetic model is thus consistent with the experiment. 2.4. The Fate of HNO in Planetary Oceans To summarize, the analysis in this section shows that under planetary conditions most of the de- posited HNO undergoes dehydrative dimerization, and becomes N2O. The dehydrative dimerization is kinetically favored over oxidization to nitrate or polymerization to nitrite by at least four orders of magnitude under anoxic conditions, and in most cases, by ten orders of magnitude. The insight we obtain here by evaluating the kinetic rates of HNO removal pathways clarifies the fate of HNO produced in anoxic atmospheres and deposited in the oceans. Models of the at- mospheric evolution for Earth and planets have assumed that the HNO would quickly become ni- trite and nitrate in the ocean (Mancinelli & McKay 1988; Wong et al. 2017; Laneuville et al. 2018; Ranjan et al. 2019). The experimental basis for this early assumption was the pulse radiolysis experiments for Reactions R4-R6 and R7-R9 (Gratzel et al. 1970; Seddon et al. 1973) and the experiment of Summers & Khare (2007). These experiments used NO-rich fluids, and thus to ap- ply their results one must evaluate the implied ki- netic rates for reasonable planetary conditions and compare with other potential reaction pathways. Here we show that for anoxic atmospheres, de- hydrative dimerization is the dominant pathway, and for oxygen-rich atmospheres, deprotonation followed by oxidation is the dominant pathway. These results are also testable by experiments in the laboratory. It is therefore reasonable to consider Reaction (R2) the sole reaction of HNO in the aqueous phase. The produced N2O, because of its low solu- bility, is released to the atmosphere and eventually photolyzed to become N2. The formation of HNO in the atmosphere is thus not an effective path to- 6 HU & DELGADO ward nitrite or nitrate, and does not lead to seques- tration of molecular nitrogen in the aqueous phase. measured spectrum in the ultraviolet (France et al. 2016) in the photochemistry model. 3. COUPLED ATMOSPHERE-OCEAN MODEL We develop an ocean chemistry module and cou- ple it with the atmospheric photochemistry model of Hu et al. (2012); Hu et al. (2013) to determine the lifetime of N2 in anoxic atmospheres in con- tact with liquid-water oceans. The photochemistry model has been validated by computing the at- mospheric compositions of present-day Earth and Mars, as the outputs agreed with the observations of major trace gases in Earth's and Mars' atmo- spheres (Hu 2013). The model includes a com- prehensive reaction network for O, H, C, N, and S species including sulfur and sulfuric acid aerosols, and its applications to simulating anoxic atmo- spheres and maintaining the redox flux balance of the atmosphere and the ocean have been well- documented (James & Hu 2018) and compare well with other photochemical models (Gao et al. 2015; Harman et al. 2018). For this work, we choose to simulate a 1-bar atmosphere of 95% N2 and 5% CO2, as this kind of anoxic atmosphere is akin to the O2-poor and CO2-rich environment of the Archean Earth, and is often adopted as the archetype for anoxic exoplanet atmospheres (e.g., Tian et al. 2014; Domagal-Goldman et al. 2014; Harman et al. 2015). We assume a surface temperature of 288 K and a stratospheric temperature of 200 K and include volcanic outgassing of CO, H2, SO2, and H2S in the same way as in James & Hu (2018). We use the entire reaction network of the atmospheric pho- tochemistry model of Hu et al. (2012); Hu et al. (2013), except the organic compounds that have more than two carbon atoms and their reactions. The outgassing rate adopted here is not high enough to produce a H2SO4 aerosol layer in the atmosphere. We include both a Sun-like star and an M dwarf star as the parent star. For the M dwarf star, we use GJ 876 as the representing case and apply its To simulate the effect of atmospheric nitro- gen fixation, we start from the terrestrial pro- duction rate of NO by lightning, 6 × 108 cm−2 s−1 (Schumann & Huntrieser 2007). Changing the main oxygen donor from O2 to CO2 and H2O would lead to approximately one-order-of- magnitude less NO, but the lightning rate also depends on how convective the atmosphere is (Wong et al. 2017; Harman et al. 2018). Be- sides, bolide impacts and hot volcanic vents may also contribute substantially to the source of NO (McKay et al. 1988; Mather et al. 2004). We there- fore explore the effect of changing NO flux by three orders of magnitude from the terrestrial lightning value to cover these varied scenarios. Also, assuming the oxygen comes from CO2, each molecule of NO produced is accompanied by an- other molecule of CO. We include this conjugate CO source in the model, and in this way, no net redox change is introduced to the atmosphere. 3.1. Ocean Chemistry and Deposition Velocities of Nitrogen Species Chemical reactions in the ocean affect the at- mospheric photochemistry model by adjusting the rate of gas exchange between the atmosphere and the ocean. Conceptually, the transfer flux from the atmosphere to the ocean can be expressed as φ = vmax(n − MC/H) = vdepn where vmax is the max- imum deposition velocity and vdep is the effec- tive deposition velocity, n is the number density at the bottom of the atmosphere, M is the con- centration at the surface ocean, H is Henry's law constant, and C is a unit conversion factor de- pending on the definition of Henry's law constant (Kharecha et al. 2005). The effective deposition velocity depends on how fast the ocean can "pro- cess" the deposited gas: if the ocean removes the gas quickly, then M → 0, and vdep → vmax; whereas if the ocean cannot remove the gas, Henry's law equilibrium could be established, and in this case, M → nH/C and vdep → 0. vmax can be approxi- LIFETIME OF NITROGEN 7 mated by the speed for the gas to diffuse through laminar layers at the interface between the atmo- sphere and the ocean, aka. the two-film model (Broecker & Peng 1982) and is sensitive to the sol- ubility of the gas, the wind speed, and the tempera- ture (Domagal-Goldman et al. 2014; Harman et al. 2015). For highly soluble species vmax ∼ 1 cm s−1 and for weakly soluble ones, vmax ∼ 10−4 − 10−3 cm s−1. as the boundary condition, and the N2O is readily photodissociated in the atmosphere. For NO, NO2, and HNO, we solve for their con- centrations in the ocean, using the rates of Reac- tion (R2) and the following reactions in the aque- ous phase: NO + NO2 −−−−→ 2 NO2 − + 2 H+ (R10) Species Deposition velocity cm s−1 0 calculated iteratively calculated iteratively 1 1 calculated iteratively 1 1 1 N2O NO NO2 NO3 N2O5 HNO HNO2 HNO3 HNO4 Table 3. Effective deposition velocities for nitrogen species. Table 3 lists the effective deposition velocities for nitrogen species. We do not include any pro- cess that removes N2O in the ocean, and thus its deposition velocity is zero. For HNO2 and HNO3, the ocean's capacity to store them is vast, and thus we assume that they are permanently lost to the ocean once deposited, and their deposition veloci- ties approach vmax. NO3, N2O5, and HNO4 quickly -- , and thus they are react or decomposes to NO3 also considered permanently lost once deposited. -- and Over geologic timescales the dissolved NO2 +, or to NO, N2O, NO3 and N2 and released back to the atmosphere, by cycling through hydrothermal vents (Wong et al. 2017; Laneuville et al. 2018), and ultraviolet pho- tolysis and reduction by Fe2+ (e.g., Stanton et al. 2018; Ranjan et al. 2019). This potential source of gaseous NO and N2O is not included in the current model since we explore a wide range of NO flux -- can be reduced to NH4 NO2 + NO2 −−−−→ NO2 − + NO3 − + 2 H+ (R11) The rate constants of Reactions (R10) and (R11) are from Lee (1984) and tabulated in Table 1. For each mixing ratio (or partial pressure) of NO, NO2, and HNO at the bottom of the atmosphere, their steady-state concentrations in the ocean can be cal- culated, assuming homogeneous distribution in the ocean. The results are then expressed in the effec- tive deposition velocities and are shown in Figure 1. Several important observations can be drawn from Figure 1. (1) NO does not substantially trans- fer to the ocean unless the mixing ratio of NO2 is approaching 1 ppm. This is because the removal of NO by Reaction (R10) requires another NO2. The conditions for such a large abundance of NO2 at the surface is rarely achieved. The effective deposition velocity of NO can be large when the mixing ratio of NO is very small. This however does not imply a substantial transfer flux because the flux is the product of the deposition velocity and the mixing ratio. The deposition flux of NO is thus severely limited by the kinetic rate of Reaction (R10). (2) NO2 practically deposits at vmax. Unless the light- ning rate is very small, the partial pressure of NO is always high enough to effectively remove NO2 via Reaction (R10). Even when the mixing ratio of NO is indeed very small (see Figure 1, middle panel, yellow line), Reaction (R11) can efficiently remove the dissolved NO2 and make the deposition velocity to approach vmax for a mixing ratio of NO2 greater than 10−12. Since the deposition flux would always be small at the low end of the lightning 8 HU & DELGADO 10-4 2 ] 1 - s m c [ p e d V 1.5 1 0.5 0 10-14 =3 10-14 =2 10-17 =2 10-7 X NO 2 X NO 2 X NO 2 10-12 10-10 10-8 10-6 10-4 10-4 x NO = 4 10-11 X NO = 3 10-14 = 5 10-5 X X NO NO 10-16 10-14 10-12 x NO 2 10-10 10-8 10-6 ] 1 - s m c [ p e d V 11 10 9 8 7 6 5 0.9 0.85 0.8 0.75 0.7 0.65 0.6 ] 1 - s m c [ p e d V 0.55 10-15 10-14 10-13 10-12 x HNO 10-11 10-10 10-9 Figure 1. Effective deposition velocities of NO, NO2, and HNO as a function of the partial pressure at the bot- tom of the atmosphere. The dashed lines show vmax. The deposition velocities of NO and NO2 depend on the partial pressure of the other gas. For this reason, three cases are shown, with typical (blue lines), low (yellow lines), and high (purple lines) abundances of the other gas. The calculations are performed for a 3-km deep, homogeneous ocean. rate, Figure 1 indicates that in practice the deposi- tion of NO2 is always efficient. (3) The deposition of HNO is generally quite efficient, with vdep close to vmax. But as shown in Section 2, this deposition leads to a return flux of N2O to the atmosphere. Because the effective deposition velocities de- pend on the partial pressure at the bottom of the atmosphere, we need to solve the coupled atmosphere-ocean chemistry model iteratively. For each scenario, we typically start with vmax. Once a steady-state solution is found for the atmospheric chemistry, we use the mixing ratio of NO, NO2, and HNO at the bottom of the atmosphere to cal- 108 106 104 102 100 ) r y G ( e m i t e f i L n e g o r t i N 10-2 10-4 105 G star (Sun) M star (GJ 876) 106 107 108 109 1010 1011 1012 NO flux (molec. cm -2 s-1 ) Figure 2. Lifetime of molecular nitrogen in plane- tary atmospheres in contact with liquid water oceans. The dashed line shows 1 billion years for comparison. Lightning in Earth's atmosphere produces a NO flux of 6 × 108 molecule cm−2 s−1. The lifetime is well greater than 1 billion years unless the NO source flux is partic- ularly strong. culate their effective deposition velocities. We also add the corresponding return flux of N2O as part of the revised boundary conditions. We then re- launch the atmospheric chemistry calculation and find a new steady-state solution. This procedure is repeated until the steady-state mixing ratios of NO, NO2, and HNO no longer change. Typically only a handful of iterations are required. As such, we can found self-consistent solutions that satisfy both the atmosphere and ocean chemistry. To summarize, the analysis presented so far indi- cates that the deposition of NO or HNO cannot be a net sink for molecular nitrogen in the atmosphere, because NO does not deposit efficiently and HNO deposition leads to a return flux of N2O. There- fore, to sequester nitrogen in the ocean, the atmo- sphere must oxidize nitrogen compounds to at least as oxidized as NO2 and HNO2. With this insight, we will show in Section 4 that this required oxi- dization is quite slow in anoxic atmospheres and molecular nitrogen is therefore kinetically stable. 4. RESULTS The lifetime of molecular nitrogen in planetary atmospheres in contact with a liquid-water ocean for varied NO fluxes from lightning and other en- ergetic processes is shown in Figure 2. The life- time is calculated from the deposition fluxes of ] a P [ e r u s s e r P 100 102 104 ] a P [ e r u s s e r P 100 102 104 10-20 10-15 10-10 10-5 100 10-5 Mixing Ratio ] a P [ e r u s s e r P 100 102 104 10-14 10-12 10-10 Mixing Ratio O N 2 NO ] a P [ e r u s s e r P 100 102 104 ] a P [ e r u s s e r P 100 102 104 ] a P [ e r u s s e r P 100 102 104 10-15 10-10 Mixing Ratio 10-5 10-20 10-15 10-10 Mixing Ratio 10-5 NO 3 O N 2 5 ] a P [ e r u s s e r P 100 102 104 ] a P [ e r u s s e r P 100 102 104 10-25 10-20 10-15 Mixing Ratio 10-10 10-50 10-40 10-30 10-20 10-10 Mixing Ratio ] a P [ e r u s s e r P 100 102 104 HNO 2 ] a P [ e r u s s e r P 100 102 104 HNO 3 LIFETIME OF NITROGEN 9 O 2 CO O 2 CO 10-4 10-3 Mixing Ratio 10-2 HNO ] a P [ e r u s s e r P 100 102 104 ] a P [ e r u s s e r P 100 102 104 10-4 10-3 10-2 10-1 100 10-3 Mixing Ratio 10-2 10-1 Mixing Ratio ] a P [ e r u s s e r P 100 102 104 O N 2 ] a P [ e r u s s e r P 100 102 104 100 HNO 10-8 10-25 10-20 10-15 10-10 10-5 10-10 10-9 10-8 10-7 10-30 Mixing Ratio NO 2 Mixing Ratio NO 10-20 10-10 Mixing Ratio 100 NO 2 ] a P [ e r u s s e r P 100 102 104 ] a P [ e r u s s e r P 100 102 104 10-20 10-15 10-10 10-5 100 10-20 Mixing Ratio 10-15 10-10 Mixing Ratio 10-5 NO 3 O N 2 5 ] a P [ e r u s s e r P 100 102 104 ] a P [ e r u s s e r P 100 102 104 10-25 10-20 10-15 Mixing Ratio 10-10 10-25 10-20 10-15 10-10 10-5 Mixing Ratio HNO 2 HNO 3 ] a P [ e r u s s e r P 100 102 104 ] a P [ e r u s s e r P 100 102 104 10-25 10-20 10-15 10-10 10-5 10-30 10-25 10-20 10-15 10-10 Mixing Ratio Mixing Ratio 10-20 10-15 10-10 Mixing Ratio 10-5 10-18 10-16 10-14 10-12 10-10 Mixing Ratio Figure 3. The atmospheric abundances of O2, CO, and main nitrogen species in an N2-dominated atmosphere on a temperate rocky planet of a Sun-like star. Note that the horizontal axis of each panel is different. Dotted, solid, dashed, and dash-dot lines are from converged atmosphere-ocean chemistry models with an NO flux of 6 × 107, 6 × 108 (terrestrial value), 6 × 109, 6 × 1010 molecule cm−2 s−1, respectively. The source strength of NO has a variety of impact and feedback on the nitrogen chemistry in the atmosphere (see text). NO, NO2, NO3, N2O5, HNO2, HNO3, and HNO4, from the converged atmosphere-ocean chemistry solutions. The atmospheric abundances of these species are shown in Figures 3 and 4. The deposi- tion flux of HNO is not included in the calculation of the lifetime, as it is returned to the atmosphere in the form of N2O (Section 2). With the effective Figure 4. The same as Figure 3 but with GJ 876 as the parent star. The atmosphere becomes O2-rich at the steady state due to CO2 photolysis, and this effect has a strong impact on the nitrogen chemistry (see text). deposition velocities calculated self-consistently from the ocean-chemistry models (Figure 5), the deposition fluxes of weakly soluble species (NO and NO2) represent how fast the ocean can process them. The lifetime of molecular nitrogen is well longer than 1 billion years unless the NO flux is > 100 times larger than the present-day Earth's lightning production rate. Interestingly, we see that the life- time of nitrogen on planets around Sun-like stars is longer than that on planets around M dwarf stars. For instance, the lifetime under the lightning rate 10 HU & DELGADO 100 ] 1 - l s m c [ s e i t i c o e V n o i t i s o p e D e v i t c e f f E 10-2 10-4 10-6 10-8 10-10 105 NO NO2 HNO 106 107 108 109 1010 1011 1012 NO Production Rate [molecule cm -2 s-1 ] Figure 5. Effective deposition velocities in the con- verged atmosphere-ocean chemistry solutions. Solid lines are from the Sun-like star cases, and dashed lines are from the GJ 876 cases. The effective deposition ve- locities are self-consistently calculated, and are differ- ent from case to case. of present-day Earth is ∼ 2 billion years on an M dwarf's habitable planet, and that on a Sun- like star's habitable planet is 4-order-of-magnitude longer. The atmospheric nitrogen chemistry is substan- tially modified with the inclusion of the oceanic feedback, i.e, the inability to deposit NO and the return flux of N2O. For a Sun-like star as the par- ent star, the atmosphere is always poor in O2 (Fig- ure 3), and thus oxidizing NO is difficult. For a higher NO production rate, the steady-state mixing ratios of NO, NO2, and HNO increase, and so is the return flux of N2O. The steady-state mixing ra- tio of N2O thus also increases. The upper limit of the N2O mixing ratio obtained from our model is ∼ 10−8, still much smaller than that in present-day Earth's atmosphere (∼ 3 × 10−7). The dominant form of nitrogen deposition is HNO3 when the NO flux is ≤ 6 × 108 molecule cm−2 s−1, and it becomes HNO2 when the NO flux is ≥ 6 × 109 molecule cm−2 s−1. The surface abundance and thus the deposition rate of HNO is larger than HNO2 and HNO3 -- it is however not counted as a net loss of atmospheric nitrogen. The steady-state mixing ra- tio of NO can accumulate to a quite high level, and this is made possible by its very small effective de- position velocity (Figure 5). In other words, the ocean cannot process the NO so quickly. For the same reason, even a large surface abundance NO does not imply a major deposition pathway. The situation is more complex when the par- ent star is an M dwarf. Because M dwarfs emit strongly in the far-ultraviolet bandpass but weakly in the near-ultraviolet bandpass, their rocky planets in the habitable zone tend to accumu- late O2 from photolysis of CO2 (Tian et al. 2014; Domagal-Goldman et al. 2014; Harman et al. 2015). The NO-NO2 catalytic cycle initiated by lightning cannot remove the photochemical O2 on an M dwarf's planet either (Hu et al. 2019, ApJ, sub- mitted). Here we find the same phenomenon of abiotic O2 accumulation, and the exact amount of O2 has to do with the assumed NO flux from light- ning (Harman et al. 2018, and Hu et al. 2019, ApJ, submitted). The accumulation of abiotic O2 is not the focus of this paper, but the availability of free oxygen does impact the nitrogen chemistry and greatly reduces the lifetime of N2. With the free oxygen, the atmosphere has up to 10 ppm of O3 in the stratosphere and is thus able to efficiently oxidize NO via NO + O3 −−−−→ NO2 + O2 · (R12) Compared to the Sun-like star cases, the M star cases have higher abundances of NO2, NO3, and HNO2 at the steady state. The higher abundance of NO2 also helps deposition of NO via Reaction (R10). HNO is practically not produced in the atmosphere unless the NO flux from lightning is ≥ 6 × 109 molecule cm−2 s−1. When it is pro- duced, the corresponding return flux of N2O can drive the atmospheric N2O to up to 3 × 10−8. To compare, the terrestrial (biological) emission rate of N2O would lead to a much higher abundance of ∼ 10−6 (Segura et al. 2005). The response of HNO and N2O to an increasing lightning rate is not monotonic, and this reflects the competing effects of free oxygen, and a low level of near-ultraviolet irradiation and low abundances of OH and HO2 in the atmosphere. 5. DISCUSSION LIFETIME OF NITROGEN 11 5.1. Lifetime of Nitrogen on Archean Earth We can apply the results to Archean Earth as the modeled atmosphere irradiated by a Sun-like star has an oxidation state similar to Earth before the rise of oxygen. Except for bolide impact that con- centrated in the earliest time (McKay et al. 1988), the production rate of NO from lightning and hot volcanic vents would be in the range of 6 × 107 ∼ 6 × 108 molecule cm−2 s−1 (Mather et al. 2004; Wong et al. 2017; Harman et al. 2018). With this input, we find that the total flux of nitrogen depo- sition would in the range of 1.6 × 104 ∼ 1.5 × 105 molecule cm−2 s−1. In other words, only ∼ 0.03% of the reactive nitrogen produced in the atmosphere is permanently lost to the ocean. The lifetime of ni- trogen is 104 billion years or larger, implying that the N2 atmosphere is stable without any help from nitrate-consuming microbes. Of the deposition flux of nitrogen species, ap- proximately 80% is HNO3 and 20% is HNO2. The flux of nitrate deposition we calculate is consis- tent in the ballpark with Wong et al. (2017) but we clarify the oceanic feedback to the gas depo- sition and we remove HNO from effective deposi- tion. Assuming that the residence time of this ni- trite and nitrate is determined by the ocean cycling through high-temperature hydrothermal vents (∼ 0.4 billion years, Wong et al. 2017), and an aver- age ocean depth of ∼ 3 km, we estimate the con- centration of nitrate to be 0.9 -- 9 µM, and that of nitrite to be 0.2 -- 2 µM in the Archean ocean. If circulation through all hydrothermal vents causes the removal of nitrite and nitrate (Laneuville et al. 2018), the residence time reduces to ∼ 10 million years and the nitrate and nitrite concentrations fur- ther reduce by two orders of magnitude. Cycling through hydrothermal vents is proba- bly not the only way to remove nitrite and ni- trate in the ocean. Ranjan et al. (2019) compares the kinetic loss rate of oceanic nitrite and ni- trate due to hydrothermal vents, ultraviolet pho- tolysis (Zafiriou 1974; Carpenter & Nightingale 2015), and reactions with reduced iron (Jones et al. 2015; Buchwald et al. 2016; Grabb et al. 2017; Stanton et al. 2018). The loss rates due to photoly- sis and reactions with reduced iron can be greater than that due to hydrothermal vents by orders of magnitude. This implies that the concentrations of nitrite and nitrate we estimate in this section is an upper limit and the actual concentrations can be much lower. 5.2. Abiotic N2O in Anoxic Atmospheres In this work we show that HNO produced in the atmosphere would become N2O when an aqueous environment exists. One might ask if this source of N2O constitutes a "false positive" for using N2O as a biosignature gas (e.g., Des Marais et al. 2002). With the coupled atmosphere-ocean model, we find that the abundance of N2O produced by HNO dehydrative dimerization is always smaller than the abundance of N2O that would be pro- duced from a source strength of current Earth's biosphere, by more than one order of magni- tude, but it can be comparable to a lower biolog- ical N2O production in Earth's anoxic past (e.g. Rugheimer & Kaltenegger 2018). This is true for either a Sun-like star or an M star as the parent star. The difference in the N2O mixing ratio by more than one order of magnitude causes an ap- preciable difference in the N2O spectral features in the infrared (e.g. Rugheimer & Kaltenegger 2018). The use of N2O as a biosignature gas thus requires the detection of its source strength at the level of current Earth's biosphere. 6. CONCLUSION We present a coupled atmosphere-ocean chem- istry model to study the lifetime of molecular nitro- gen (N2) in planetary atmospheres in contact with a liquid-water ocean. The question of lifetime exists because nitrogen is the background gas for canoni- cal planetary habitability scenarios and because ni- trogen could be sequestered in the ocean when it is chemically converted to soluble compounds like nitrites and nitrates. 12 HU & DELGADO We clarify several important features of nitro- gen's aqueous-phase chemistry for planetary appli- cations. First, we find that dehydrative dimeriza- tion is the main loss pathway of HNO, the dom- inant nitrogen species produced in anoxic atmo- spheres. This reaction produces N2O, which is then released to the atmosphere and photodissoci- ated to become N2. This finding corrects the long- standing assumption that the HNO would even- tually become nitrate in the ocean. Second, we find that the deposition flux of NO is always very small under anoxic conditions. These findings col- lectively indicate that sequestering nitrogen in the ocean requires atmospheric oxidation to at least as oxidized as NO2 and HNO2. We determine that the lifetime of molecular ni- trogen is well longer than 1 billion years unless the NO flux is > 100 times larger than the present- day Earth's lightning production rate. As such, N2 atmospheres on Archean Earth and habitable ex- oplanets of both Sun-like and M dwarf stars are kinetically stable against aqueous-phase sequestra- tion. This result affirms the nitrogen-based habit- ability on rocky planets. This research was supported by NASA's Exo- planets Research Program grant #80NM0018F0612. The research was carried out at the Jet Propulsion Laboratory, California Institute of Technology, un- der a contract with the National Aeronautics and Space Administration. REFERENCES Broecker, W., & Peng, T. 1982, Lamont-Doherty Geological Observatory, Columbia University, Palisades, New York, 10964, 690 Harman, C. E., Schwieterman, E. W., Schottelkotte, J. C., & Kasting, J. F. 2015, ApJ, 812, 137 Hu, R. 2013, PhD thesis, Massachusetts Institute of Buchwald, C., Grabb, K., Hansel, C. M., & Wankel, Technology S. D. 2016, Geochimica et Cosmochimica Acta, 186, 1 Carpenter, L. J., & Nightingale, P. D. 2015, Chemical reviews, 115, 4015 Des Marais, D., O Harwit, M., Jucks, K., et al. 2002, Astrobiology, 2, 153 Domagal-Goldman, S. D., Segura, A., Claire, M. W., Robinson, T. D., & Meadows, V. S. 2014, ApJ, 792, 90 France, K., Loyd, R. O. P., Youngblood, A., et al. 2016, ApJ, 820, 89 Gao, P., Hu, R., Robinson, T. D., Li, C., & Yung, Y. L. 2015, ApJ, 806, 249 Goldblatt, C., Claire, M. W., Lenton, T. M., et al. 2009, Nature Geoscience, 2, 891 Hu, R., Seager, S., & Bains, W. 2012, ApJ, 761, 166 Hu, R., Seager, S., & Bains, W. 2013, ApJ, 769, 6 James, T., & Hu, R. 2018, ApJ, 867, 17 Jones, L. C., Peters, B., Lezama Pacheco, J. S., Casciotti, K. L., & Fendorf, S. 2015, Environmental science & technology, 49, 3444 Kasting, J. F., & Walker, J. C. 1981, Journal of Geophysical Research: Oceans, 86, 1147 Kasting, J. F., Whitmire, D. P., & Reynolds, R. T. 1993, Icarus, 101, 108 Kharecha, P., Kasting, J., & Siefert, J. 2005, Geobiology, 3, 53 Krissansen-Totton, J., Arney, G. N., & Catling, D. C. 2018, Proceedings of the National Academy of Sciences, 115, 4105 Laneuville, M., Kameya, M., & Cleaves, H. J. 2018, Grabb, K. C., Buchwald, C., Hansel, C. M., & Wankel, Astrobiology, 18, 897 S. D. 2017, Geochimica et Cosmochimica Acta, 196, 388 Gratzel, M., Taniguchi, S., & Henglein, A. 1970, Ber Bunsenges Phys Chem, 74, 1003 Lee, Y.-N. 1984, Atmospheric aqueous-phase reactions of nitrogen species, Tech. rep., Brookhaven National Lab., Upton, NY (USA) Mancinelli, R. L., & McKay, C. P. 1988, Origins of Halevy, I., & Bachan, A. 2017, Science, 355, 1069 Harman, C., Felton, R., Hu, R., et al. 2018, ApJ, 866, Life and Evolution of the Biosphere, 18, 311 Mather, T. A., Pyle, D. M., & Allen, A. G. 2004, 56 Geology, 32, 905 LIFETIME OF NITROGEN 13 McKay, C. P., Scattergood, T. W., Pollack, J. B., Shafirovich, V., & Lymar, S. V. 2002, Proceedings of Borucki, W. J., & Van Ghyseghem, H. T. 1988, Nature, 332, 520 Miranda, K. M. 2005, Coordination Chemistry Reviews, 249, 433 Navarro-González, R., Molina, M. J., & Molina, L. T. 1998, Geophysical Research Letters, 25, 3123 Ranjan, S., Todd, Z. R., Rimmer, P. B., Sasselov, D. D., & Babbin, A. R. 2019, Geochemistry, Geophysics, Geosystems, 20, 2021 Rugheimer, S., & Kaltenegger, L. 2018, The Astrophysical Journal, 854, 19 the National Academy of Sciences, 99, 7340 Stanton, C. L., Reinhard, C. T., Kasting, J. F., et al. 2018, Geobiology, 16, 597 Summers, D. P., & Khare, B. 2007, Astrobiology, 7, 333 Tian, F., France, K., Linsky, J. L., Mauas, P. J., & Vieytes, M. C. 2014, Earth and Planetary Science Letters, 385, 22 Vladilo, G., Murante, G., Silva, L., et al. 2013, The Astrophysical Journal, 767, 65 Walker, J. C., Hays, P., & Kasting, J. F. 1981, Journal of Geophysical Research: Oceans, 86, 9776 Wong, M. L., Charnay, B. D., Gao, P., Yung, Y. L., & Schumann, U., & Huntrieser, H. 2007, Atmospheric Russell, M. J. 2017, Astrobiology, 17, 975 Chemistry and Physics, 7, 3823 Wordsworth, R. D., & Pierrehumbert, R. T. 2013, The Seddon, W., Fletcher, J., & Sopchyshyn, F. 1973, Canadian Journal of Chemistry, 51, 1123 Segura, A., Kasting, J. F., Meadows, V., et al. 2005, Astrobiology, 5, 706 Astrophysical Journal, 778, 154 Yung, Y., & McElroy, M. 1979, Science, 203, 1002 Zafiriou, O. C. 1974, Journal of Geophysical Research, 79, 4491
1511.09211
1
1511
2015-11-30T09:21:20
Dynamical considerations for life in multihabitable planetary systems
[ "astro-ph.EP" ]
Inspired by the close-proximity pair of planets in the Kepler-36 system, we consider two effects that may have important ramifications for the development of life in similar systems where a pair of planets may reside entirely in the habitable zone of the hosting star. Specifically, we run numerical simulations to determine whether strong, resonant (or non-resonant) planet-planet interactions can cause large variations in planet obliquity---thereby inducing large variations in climate. We also determine whether or not resonant interactions affect the rate of lithopanspermia between the planet pair---which could facilitate the growth and maintenance of life on both planets. We find that first-order resonances do not cause larger obliquity variations compared with non-resonant cases. We also find that resonant interactions are not a primary consideration in lithopanspermia. Lithopanspermia is enhanced significantly as the planet orbits come closer together---reaching nearly the same rate as ejected material falling back to the surface of the originating planet (assuming that the ejected material makes it out to the location of our initial conditions). Thus, in both cases our results indicate that close-proximity planet pairs in multihabitable systems are conducive to life in the system.
astro-ph.EP
astro-ph
Draft version October 8, 2018 Preprint typeset using LATEX style emulateapj v. 5/2/11 5 1 0 2 v o N 0 3 . ] P E h p - o r t s a [ 1 v 1 1 2 9 0 . 1 1 5 1 : v i X r a DYNAMICAL CONSIDERATIONS FOR LIFE IN MULTIHABITABLE PLANETARY SYSTEMS Jason H. Steffen1,2 University of Nevada, Las Vegas, Department of Physics and Astronomy, 4505 S. Maryland Pkwy., Las Vegas, NV 89154-4002 Harvard Smithsonian Center for Astrophysics, Institute for Theory and Computation, 60 Garden Street, Cambridge, MA 02138 Gongjie Li Draft version October 8, 2018 ABSTRACT Inspired by the close-proximity pair of planets in the Kepler-36 system, we consider two effects that may have important ramifications for the development of life in similar systems where a pair of planets may reside entirely in the habitable zone of the hosting star. Specifically, we run numerical simulations to determine whether strong, resonant (or non-resonant) planet-planet interactions can cause large variations in planet obliquity -- thereby inducing large variations in climate. We also determine whether or not resonant interactions affect the rate of lithopanspermia between the planet pair -- which could facilitate the growth and maintenance of life on both planets. We find that first- order resonances do not cause larger obliquity variations compared with non-resonant cases. We also find that resonant interactions are not a primary consideration in lithopanspermia. Lithopanspermia is enhanced significantly as the planet orbits come closer together -- reaching nearly the same rate as ejected material falling back to the surface of the originating planet (assuming that the ejected material makes it out to the location of our initial conditions). Thus, in both cases our results indicate that close-proximity planet pairs in multihabitable systems are conducive to life in the system. Subject headings: Celestial Mechanics,Planets and Satellites -- terrestrial planets, dynamical evolution and stability 1. INTRODUCTION NASA's Kepler mission has discovered several thou- sand candidate exoplanet systems including many hun- dreds with multiple planets (Borucki et al. 2010; Stef- fen et al. 2010; Batalha et al. 2013; Burke et al. 2014; Rowe et al. 2015; Mullally et al. 2015). Among the many interesting systems identified by this mission is Kepler- 36, which has two planets orbiting near the 7:6 mean- motion resonance (MMR) (Carter et al. 2012). These two planets have orbital periods of 13.8 and 16.2 days, have masses of 4.5 and 8.1 M⊕, and radii of 1.5 and 3.7 R⊕ for the inner and outer planets respectively. It is striking to note that these two planets (with densities differing by a factor of 8) orbit at distances that differ by only 10%. And, while Kepler-36 is among the most extreme cases currently known, several other pairs of planets lie near first-order MMRs with similarly high indices such as the 6:5 and 5:4. Given the context of a mission designed to measure the number of habitable planets in the galaxy and the alien nature of many of the systems that Kepler has found (e.g., circumbinary planets (Doyle et al. 2011; Welsh et al. 2012), planetary systems orbiting each component of a binary pair Kepler-132 (Lissauer et al. 2014), sys- tems with planets near chains of MMR such as Kepler- 80 (Xie 2013) and Kepler-223 (Lissauer et al. 2011), and these systems with pairs of planets in strongly interacting (and for the case of Kepler-36 manifestly chaotic orbits (Deck et al. 2012)) it is not much of a stretch to consider 1 Northwestern University, CIERA, 2131 Tech Drive, Evanston, IL 60208 2 Lindheimer Fellow systems with multiple planets that orbit in the habitable zone of their host stars, including pairs in or near MMR. Here we consider two issues that have been raised re- garding life on Earth but in the context of these Kepler- 36-inspired multihabitable systems. First, whether or not close planetary orbits or resonant interactions cause any significant variations in planet obliquities -- the an- gle between the planet's spin and orbital axes. Obliquity variation plays a major role in the modulation of climate, since it determines the latitudinal distribution of solar ra- diation. For the case of Mars (an ocean-free, atmosphere- ice-regolith system), the obliquity changes would result in drastic variations of atmospheric pressure caused by runaway sublimation of CO2 ice (Toon et al. 1980; Fanale et al. 1982; Pollack and Toon 1982; Francois et al. 1990; Nakamura and Tajika 2003; Soto et al. 2012). For Earth-like planets (planets partially covered by oceans) the change of climate depends on the specific land-sea distribution and on the position within the hab- itable zone around the star. The ice ages on the Earth, for example, are closely associated with the variation in insolation at high latitudes, which depends on the orbital eccentricity and orientation of the spin axis according to the Milankovitch theory (e.g. Weertman 1976; Hays et al. 1976; Imbrie 1982). While it is debatable whether the variation in obliquity truly renders a planet unin- habitable (obliquity variations may, in some cases, ac- tually increase the habitability of a planet (Armstrong et al. 2014), though civilizations that rely on agriculture may struggle), it is clear that the climate can change drastically as the obliquity varies (Williams and Kast- ing 1997; Chandler and Sohl 2000; Jenkins 2000; Spiegel et al. 2009). 2 Steffen & Li The spin-axis dynamics of planets in the solar system has been extensively studied in the literature. At present, the obliquity variation of the Earth is regular and only undergoes small oscillations between 22.1◦ and 24.5◦ with a 41000 year period (e.g. Vernekar 1972; Laskar and Robutel 1993). Without the Moon, the obliquity of the hypothetical Earth is chaotic, but is constrained between 0 − 45◦ over billion year timescales (Laskar et al. 1993; Lissauer et al. 2012; Li and Batygin 2014a) -- though Lis- sauer et al. (2012) showed some conditions where the obliquity of the Earth can be stable in the absence of the Moon. The Earth's obliquity remained stable as the Moon moved outward before Late Heavy Bombardment (Li and Batygin 2014b), yet Martian obliquity is thought to have been chaotic throughout the solar system's life- time (Ward 1973; Touma and Wisdom 1993; Laskar et al. 1993; Brasser and Walsh 2011). If strong planet-planet interactions in multihabitable systems cause or preclude large obliquity variations, then the development of intel- ligent life would be hampered or aided by its effects on climate. The second issue we address, perhaps more rich in con- sequences, is to understand how biological material might be exchanged between planets in a multihabitable sys- tem through the process of lithopanspermia (hereafter simply "panspermia"). Panspermia in the solar system has been studied at some length (e.g., Melosh and Tonks 1993; Gladman et al. 1996; Worth et al. 2013, see this last reference for a good historical review). And, from obser- vations of meteorites, some hypothesize that a fraction of life on Earth may have originated on Mars (which may have supported life sooner than the Earth (e.g., McKay et al. 1996)). The survival of life on collision ejecta, both at the time of collision (Melosh 1988) and over the subse- quent, extended trajectory was studied by Mileikowsky et al. (2000), who found that successful transfer (espe- cially from Mars to Earth) was highly probable. Two Earth-like planets in Kepler-36-like orbits would likely have a much greater opportunity to exchange such material than the terrestrial planets in the solar system. The planets would subtend over 25× the solid angle at conjunction than the Earth does from Mars and the rela- tive velocities of the planets, and hence the ejecta parti- cles, could be much less. This scenario allows more of the ejected particles (especially with low relative velocities) to successfully make the trip between planets. The close proximity of the two planets reduces the reliance on the effects of secular resonances to successfully transfer par- ticles (Dones et al. 1999). (For the solar system, secular resonances excite orbital eccentricities of the ejected par- ticles and facilitate their transfer over large distances.) Consider the scales over which biological material may be transmitted via collision ejecta (see Figure 1), on the shortest scales, individual habitable planets may have barriers such as oceans and mountain ranges that could be traversed by ejecta from a planet falling back to onto the same planet, "auto-panspermia" (which has been shown as a viable means of seeding life across the Earth (Wells et al. 2003)). The distances involved would be ∼ 0.1 → 104km. Binary planets or a single planet with multiple habitable moons are on a somewhat larger scale, perhaps 105 → 107km. On the largest scales one can (and has) considered interstellar panspermia within Figure 1. Logarithm of various length scales over which life could be seeded via panspermia. This paper is primarily concerned with multihabitable systems. a star cluster or within the galaxy (1013 → 1019km) (Melosh 2003; Belbruno et al. 2012). Between these extremes lie habitable planets orbiting different compo- nents of a stellar binary (1010 → 1012km). And, finally, multihabitable systems (108 → 109km), which we con- sider here. This paper is organized as follows. In the next section we briefly discuss the properties of the model systems we consider. Section 3 presents the evolution of planet obliq- uities in the model systems. Section 4 shows the results of our panspermia simulations. We discuss some of the implications of our work in section 5 and give concluding remarks in section 6. 2. PLANETARY SYSTEM MODELS For this investigation, we construct a sample of systems in first-order MMR along with systems that have nearly identical period ratios but that are not in MMR. The systems in MMR were created following the methods of Lee and Peale (2002); Batygin and Brown (2010). Specif- ically, we set the planets slightly outside of their resonant location, and integrate the system using a Burlisch-Stoer integrator. We evolve the planets under semi-major axis and eccentricity damping in addition to the Newtonian N-body interaction. In these simulations, the central star has a mass of one solar mass, and the planets each have masses of 1.0×10−6 M(cid:12) (about 1/3 of an Earth mass). The planet masses were chosen arbitrarily, and our results should not de- pend strongly on this quantity. We ran a large number of simulations with different semi-major axis and eccen- tricity decay rates, and selected the final state of the systems where the planets are in MMR at the end of the simulations. The corresponding non-resonant configura- tions are obtained by setting the longitude of pericenter, longitude of ascending node, and the mean anomaly to zero for both planets. A third set of non-resonant initial conditions was also chosen for the purposes of differenti- ating the effects due to resonance and the effects due to initial conditions. All of our initial conditions are given in Table 2 in the Appendix. 3. OBLIQUITY EVOLUTION The obliquity variation in a planet arises largely as a consequence of the underlying resonant structure (Laskar 1996). Specifically, the spin-axis of the planet may ex- hibit complex behavior if its precession resonates with the inclination variation. The former is controlled pri- marily by the stellar torques, whereas the latter is forced by planet-planet interactions. When the precession fre- quency of the planetary spin axis from stellar torque co- incides with the frequency of inclination variation from planetary interactions, the resonance occurs. For our systems, the stellar torque for both the res- onant and the non-resonant systems are essentially the Multihabitable Systems 3 Figure 2. The Fourier spectrum of (ieiΩ) for the 3:2 resonance (where i ≡ √−1). The blue dotted line represents the case when the two planets are in resonance, the red dashed line represents the case when the two planets are not in resonance and the an- gle variables (longitude of pericenter, longitude of ascending node, and mean anomaly) are set to 0, 120, and 240 degrees respectively, and the purple line represents the case when the angle variables are set to zero. The maximum high-amplitude frequencies are ap- proximately the same for the three cases. The black lines show the precession of the inner and the outer planets assuming the planets have Earth-like properties. same, since the two systems only differ in the orbital orientation. To calculate the frequency of the orbital inclination variation for the systems, we integrate the orbital elements numerically using the HYBRID inte- grator in the MERCURY package (Chambers 1999). The Fourier spectrum of the planetary orbital param- eter (ieiΩ), where i ≡ √−1, i is the orbital inclination, and Ω is the longitude of ascending node, are shown in Figure 2. The amplitudes and frequencies of the incli- nation variations are approximately the same. This fact indicates that first-order mean motion resonances do not affect obliquity variations -- otherwise the resonant case would be different from the nonresonant ones. The black lines show the precession of the inner and the outer plan- ets assuming the planets have Earth-like properties. To illustrate how the obliquity variation occurs, we cal- culate the obliquity as a function of time following the Hamiltonian that describes its evolution (e.g. Colombo 1966; Kinoshita 1972; Laskar et al. 1993; Touma and Wisdom 1993; Neron de Surgy and Laskar 1997; Arm- strong et al. 2004). (The detailed equations are given in the Appendix.) We set the planetary system configura- tion to be the case of "3:2 nonresonant I" (the red-dashed line) shown in Figure 2 and described in section 4.2, the dynamical ellipticity of the planet is set to that of Earth (Ed = 3 × 10−3), and the rotation period of the planet is 12 hrs. This rotation period is somewhat arbitrary as we only require that the precession rate of the inner planet's spin axis, α cos () (where α is the precession co- efficient and  is the obliquity -- see the appendix for de- tails), matches the inclination variation frequency when  ∼ 62◦. Figure 3 shows the maximum, the minimum and the mean of the obliquity variation as a function of the initial obliquity for m1 in ∼ 10 Myr. This scenario, where the inclination and precession rates coincide, cor- responds to the region where the obliquity of the planet varies with moderate amplitude. The evolution of the orbital elements for the resonant case is shown in Figure 4. Figure 3. The obliquity variation for the 3:2 resonance in 30 Myr. The precession frequency matches that of the inclination variation at  ∼ 62◦, and causes the obliquity variation at around 62◦. We briefly consider two additional scenarios relating to obliquity variation in a planetary system, though a de- tailed study of these scenarios lies beyond the scope of this work. First, we consider the case where the Earth, embedded in the solar system, is replaced by a pair of near resonant planets with equal masses in order to de- termine whether or not the size of the obliquity variations differ significantly form those of a moonless Earth. Sec- ond, we consider the case where a habitable, Mars-mass planet lives in a system with a near-resonant planet pair. To illustrate the obliquity variation for systems involv- ing closely separated planets, we compared the variation of a moonless Earth in the solar system with an analo- gous solar system where the Earth is replaced with two planets of mass m1 = m2 = 1 × 10−6 M(cid:12) located at a1 = 0.974 AU, and a2 = 1.068 AU (with period ra- tio 6/7). Different from the case of the solar system, one of the frequencies of the inclination modes peaks at ∼ 15.5(cid:48)(cid:48)/yr, and the mode at ∼ 20(cid:48)(cid:48)/ yr is lower in am- plitude for m1 and m2. The obliquity variation of the planets depends upon their precession coefficients, which we assume to be the same for both planets. The obliquity variations of this "split Earth" scenario are shown in figure 5 and are com- pared with variations of the true Earth. When the pre- cession coefficient of the planets are 21.24(cid:48)(cid:48)/yr, the varia- tion is stronger for the true Earth where the initial obliq- uity is in the range of ∼ 0 − 20 degrees, and is weaker when in the range of ∼ 45 − 65 degrees. The variation is quite similar between the true and split Earth scenarios when the initial obliquity is between ∼ 20 − 45 degrees. Using the same split-Earth modification to the solar system, we also tested the variation of the obliquity of Mars. This investigation addresses the issue of a habit- able planet residing in the same system as a close pair of planets. In this case we found that the obliquity vari- ations of Mars were not strongly affected by the near resonant pair. These results show that climate change due to obliq- uity variations for multihabitable systems embedded in the solar system would be comparable to, but not larger than, the variations of a moonless Earth. The same Frequency (''/yr)010203040Amplitude10-510-23:2 res3:2 nonres 1.3:2 nonres 2.precession rate (ǫ = 45o)ǫ0020406080100ǫmax/mean/min020406080100ǫminǫmeanǫmax 4 Steffen & Li Figure 4. The evolution of the orbital elements for the 3:2 resonance case. Here λ is the longitude of the planet in its orbit,  is the longitude of pericenter, e is the orbital eccentricity, a is the orbital semi-major axis, and i is the orbital inclination. The libration of these quantities demonstrates the resonance behavior. is true for habitable planets living in a system with a close-proximity pair where the planet-planet interactions do not affect obliquity variations in any material way. Only in relatively rare instances, where orbital parame- ters are somewhat fine-tuned (so that the timescales of inclination variation and spin precession coincide), does a planet's obliquity vary significantly. Brasser et al. (2014) showed that spin-orbit resonances could exist in exoplanet systems. And, systems may be driven to such couplings through some damping mecha- nism (e.g., tides). However, the scenarios that more read- ily produce spin-orbit couplings would likely be closer to the host star than the typical habitable zone (e.g., near the orbit of Mercury rather than Earth). Thus, barring other influences, we can expect stable climates for plan- ets in systems with resonant or near-resonant planet pairs at the same rate as stable climates in systems without them. 4. LITHOPANSPERMIA Now we consider panspermia in these systems -- the transmission of collision ejecta containing biological ma- terial from one planet to the other. As we study pansper- mia in this context, there are a few dynamical points worth making. First, since collision ejecta comes from a highly localized source and since the particles are colli- sionless (widely dispersed) following the initial impact, the available phase space for the particles is restricted (by Liouville's theorem). Consequently, collision ejecta is dynamically very cold -- having a very low, and ever de- creasing, velocity dispersion. The stream of debris will therefore be concentrated in phase-space sheets that fold and twist as the system evolves (see, e.g., Sikivie 1998) -- forming caustics or cold flows or, the term we will use, "caustic flows". The fact that the debris evolves in this manner means that when one piece of debris strikes a planet, there is ex- cess probability that additional pieces will also strike the planet and within a relatively short amount of time. A consequence of this effect would be seen in the distribu- tion of intercollision times. If the debris were randomly spread in the vicinity of the planets then the distribu- tion of intercollision times would be approximately ex- ponential (the collisions being a Poisson process). Here, the deviations from Poisson behavior would come pri- marily from long-term reduction of the ejecta population through collisions or ejections. On the other hand, the collision time intervals from caustic flows would have an excess of collisions separated by short times. An overall Poisson behavior would be visible only over long times as the planets pass through the various flows. We inves- tigate these effects below. 0246810a1 (AU)0.99610.99620.99630246810a2 (AU)1.30531.30541.30550246810e10.0210.02150.0220.02250246810e20.02250.0230.02350.0240246810i1 (degree)0.60.811.20246810i2 (degree)0.60.81time (Myr)02468103λ2-2λ1-1-10010time (Myr)02468103λ2-2λ1-2170180190 Multihabitable Systems 5 the source planet into 768 locations using Hierarchical Equal Area isoLatitude Pixelisation (HEALPix, G´orski et al. 2005). Each location is the source of three ejecta particles with randomly assigned velocities directed ra- dially outward from the surface. The starting position of each particle is located one Hill radius from the planet above the geometric center of each pixel. The choice to start at one Hill radius is to eliminate potential nu- merical artifacts related to the planet surface. Thus, for each planetary system and for each source planet in that system we had 10 initial planet locations, each with 768 ejecta starting locations, each with 3 initial velocities giv- ing 23,040 particles per simulation. For each velocity, we choose a uniformly distributed number between 0.5 and 0.5 v(cid:63)/vp where vp is the es- cape velocity from the planet (beginning at one Hill ra- dius) and v(cid:63) is the escape velocity from the star at the orbital distance of the inner planet (all set to 1 AU). This random number R is then substituted into the formula: (1) 1 + R2. (cid:112) v = vp Figure 5. The obliquity variation of planets. The upper panel shows the maximum, minimum and the mean obliquity in 20 Myr of a moonless Earth versus the initial obliquity, and the lower panels show those of a planet in the system where the Earth is substituted by two closely separated planets. Specifically, the middle panel shows the obliquity variation of the inner substituted planet (m1), and the lower panel shows that of the outer substituted planet (m2). The precession coefficient α is set to be 21.24(cid:48)(cid:48)/yr. At this value, the obliquity variation of the planet pair is larger than that of the moonless Earth when the initial obliquity is ∼ 0− 20 degrees and weaker when it is ∼ 45 − 65 degrees. Another issue to consider is the strong dependence of the results of any panspermia calculation on the chosen distribution of initial velocities. With relative ease one can choose a velocity distribution where all of the par- ticles are ejected from the system -- leaving only a single epoch for dynamical encounters between the destination planet and the ejecta (effectively eliminating the prob- ability of a successful transfer). However, if the typical ejection velocities put the particles on orbits with periods comparable to the destination planet (or slightly more or less than the escape velocity from the originating planet) then successful transfers are more likely to occur. 4.1. Setup For our panspermia simulations we consider both when the inner planet and the outer planet act as the source. We take each planetary system and evolve it over one synodic time to select a set of 10 initial configurations of the system that are roughly equidistant in the rela- tive longitudes of the two planets. We then rotate these initial configurations so that the longitude of the source planet is always equal to zero. This transformation gives a common substellar point on the source planet for anal- ysis purposes. For each initial configuration we divide the surface of This particular form for the function that defines the ve- locities is quite arbitrary and was motivated by the desire to have roughly the appropriate energy scale for the es- caped particles in the system (between just escaping from the planet and just escaping from the star). When added to the orbital motion of the planet, some of the highest velocity particles along the direction of orbital motion es- cape the system, especially when the outer planet is the source. Throughout the simulations the collision ejecta are treated as test particles and did not affect the orbits of the planets. We integrated each system for 10 mil- lion years using MERCURY -- employing the HYBRID integrator. 4.2. Resonant and non-resonant planet pairs The first test to identify any differences between reso- nant and nonresonant systems. We ran simulations as de- scribed above for both resonant and nonresonant planet pairs and for both inner and outer planets as the source. The four resonances we consider are the 7:6, 6:5, 4:3, and 3:2. For these systems the time dependence of the successful transfers were very similar, generally giving of order 1000 events. However, we noticed that the nonres- onant systems consistently had a ∼ 20% higher success rate regardless of planet source or MMR. We ran a set of 5 simulations for the 6:5 nonresonant case in order to estimate the variations that we could expect from different initial conditions for the velocities of the ejected particles. The results show that the vari- ation over the course of the whole integration is typical of Poisson fluctuations in the successful transfers (statis- tical variations of only a few percent given our ∼ 1000 successes). This fact implies that the difference between the resonant and nonresonant success rates are quite significant -- generally between 2.5 and 6 sigma for the different MMRs. This result may have been due to the fact that the angle variables (longitude of pericenter, longitude of as- cending node, and mean anomaly) were all set to zero -- giving a highly specialized initial condition. Conse- quently, we ran additional simulations for the 4:3 and 3:2 MMRs with nonresonant planet pairs where the lon- gitudes of pericenter and ascending node, and the mean 020406080obliquity variation020406080020406080obliquity variation020406080initial obliquity020406080obliquity variation020406080 6 anomaly for the inner planet were set to 0, 120◦, and 240◦ respectively (these are the "I" cases as shown in Figure 2). For these new nonresonant cases the number of successful transfers were either similar to, or somewhat less than, the resonant cases (and significantly below the original nonresonant case). That the transfer rate from the second set of nonres- onant simulations differed significantly from the origi- nal set shows that the planetary initial conditions play a strong role in panspermia (imagine two planets with very different orbital inclinations). The variations that arise from changing a few orbital angles are comparable to or larger than the variations between resonant and non- resonant systems. This fact suggests that the primary cause of the observed differences between the resonant and nonresonant simulations may be due primarily to the initial conditions of the planets rather than their res- onant behavior, or lack thereof. A definitive statement in this regard, since its effects must be very small ((cid:46) 1%), would require a study that lies beyond our scope. Nev- ertheless, since the ejecta themselves are unlikely to be in resonance with the destination planet, it would not be surprising that the resonance behavior of the planets is of little consequence. 4.3. Variations with Longitude In our simulations we found that the successful trans- fers depended somewhat on the initial longitude of the particles. When the outer planet was the source, more successfully transferred particles were ejected opposite the direction of motion. When the inner planet was the source, more successfully transferred particles were ejected from the substellar and antistellar points. This longitudinal variation should depend upon the distri- bution of initial velocities. Consider the inner planet source, in the extreme where the particles barely escape the planet, only the particles in the direction of motion would have sufficient energy to reach the orbit of the outer planet. In the other extreme, with high velocity ejecta, only particles that directly face the outer planet at the time of the collision would transfer successfully. To demonstrate this effect, we ran a suite of simula- tions using a different initial velocity distribution. In this case the velocities were again assigned using Equa- tion (1) but with the random number being chosen from an exponential distribution with a scale parameter of λ = 0.1v(cid:63)/vp. We call this the "restricted" velocity dis- tribution as it has a much smaller range of values for the assigned particle velocities. The motivation for this distribution was simply to have a contrasting example. Using these simulations we examine three observable con- sequences for successfully transferred particles and com- pare them with the results of the "standard" velocity distribution from before. These observables are: 1) the longitudinal dependence of the successful transfers, 2) the transfer rate as a function of the separation of the planetary orbits, and 3) the inter-arrival time of the col- lisions (testing the Poisson nature of the collision rate). Figure 6 shows an example of the original spatial lo- cations of the transferred particles for both the standard and the restricted velocity distributions. The bottom two panels show the locations of the initial ejecta on the inner planet for the two velocity distributions. The top panel, which is a histogram of the initial longitudes only, Steffen & Li Anderson Darling p-values for the velocity distributions compared with a uniform distribution and with each other. Table 1 MMR Source 3:2 3:2 4:3 4:3 6:5 6:5 7:6 7:6 1 2 1 2 1 2 1 2 Standard Restricted 0.22 0(cid:63) 0.03 0 3 × 10−6 0 2 × 10−5 0 0.34 0.28 0.23 0.78 0.35 0.23 0.73 0.32 S v. R 0.19 0 0.02 0 7 × 10−5 0 8 × 10−4 0 (cid:63) Zero values indicate that the p-value was less than machine pre- cision. had a small random number added to the true longi- tude values. The reason for this addition was to en- able more reliable statistical tests -- since the initial lo- cations of the ejecta were discrete, statistical tests that rely on the empirical distribution (e.g., the Anderson- Darling test) are compromised because there are large jumps whenever multiple particles from the same lon- gitude are considered. These large jumps yield anoma- lously large differences between the empirical and com- parison distributions. The added random numbers were normally distributed with zero mean and a standard de- viation of 2π/64 (there are 32 initial locations around the equator of the source planet and this quantity is half of their separations). Table 1 shows the Anderson-Darling p-values compar- ing the distribution of the initial longitudes for the two initial velocity distributions both to a uniform distribu- tion and to each other. For all MMRs, and for both planet sources, a uniform distribution can not be ex- cluded for the restricted velocity distribution. For the standard velocity distribution, a uniform distribution is excluded with high confidence for most cases (there is some variation in these values depending upon the ran- dom numbers added to each longitude, but the changes do not affect the order of magnitude of the p-values). For the successful transfers using the standard velocity distribution, there is also a dependence on the velocity that varies with longitude. Figure 7 shows the distribu- tion of the initial longitudes of the particles as a function of their their initial velocities for both the inner planet and the outer planet as source (the case shown is for a pair near the 4:3 MMR). As one may expect, when the outer planet is the source, the particles simply need to lose sufficient energy to cross the inner planet orbit and the range of initial velocities is much larger away from the direction of motion than the range of velocities in the direction of motion. A similar, somewhat more interest- ing, effect can be seen when the inner planet is the source where there is a wide range of successful initial veloci- ties in the substellar and anti-stellar directions. Similar results occur for other MMRs. However, since the re- stricted velocity distribution is much more narrow than the standard distribution, the same structure does not appear. In addition to the differences in longitudes of success- ful transfers between the two velocity distributions, we look at changes in the transfer rate that depend upon the difference in semi-major axis of the two planets. The actual transfer rates are much higher for the restricted velocity distribution (by about a factor of three). This Multihabitable Systems 7 Figure 6. Longitudinal (top) and spatial distribution (middle and bottom) of initial locations of ejecta from the inner planet for suc- cessful transfers from one planet to another near the 4:3 MMR. The middle panel shows the initial locations for the standard ve- locity distribution and produce the outline histogram in the top panel. The bottom panel shows the initial locations for the re- stricted velocity distribution and produce the solid histogram in the top panel. The histograms show that the velocity distribution affects the longitudinal distribution of source points for successful transfers. Here 0 corresponds to the sub-stellar point and 3π/2 is the direction of motion. For the bottom panels, the areas of the circles are proportional to the number of successful transfers from that location (30 total are possible, but typically only one to a few are realized). difference is due in part to the more densely occupied phase space of the particles, there being far less spread in the velocities. However, Gladman et al. (1996) showed that the Martian meteorites on Earth also showed evi- dence for initial velocities near the escape velocity from Mars -- a result consistent with our statement. Regardless of the overall rates, the relative rates for the different MMRs between these two velocity distributions are quite similar. Figure 8 shows the relative fraction of successful transfers for both velocity distributions as a function of the difference in semi-major axes of the two planetary orbits. (That is, for example, the number of successful transfers near the 7:6 or 6:5 MMR divided by the total number of successful transfers for all MMRs -- Nj+1:j/Nall.) Recall that the estimated fractional uncer- tainties in these rates is only a few percent and could not Figure 7. A smoothed density histogram of the fraction of suc- cessfully transferred particles as a function of the scaled velocity and longitude for the inner (top) and outer (bottom) planets for the standard velocity distribution. Note that the range in veloci- ties is larger at the substellar and antistellar points for the inner planet source and it is larger for the direction opposite the orbital motion for the outer planet. explain the observed behavior. Power-law fits to the two sets of points both yield declining efficiency that scales approximately as ∼ (a2 − a1)−2/3 (shown for reference). This apparent similarity between the results of the two velocity distributions is somewhat coincidental since one can construct limiting cases where there are no successful transfers at all (by having very small velocities) or where the rate falls as (a2 − a1)−2 (by having very large ones). Another issue that we investigate with our simulations is the ratio of the ejecta that is recaptured on the source planet to the ejecta that is successfully transferred to the other planet. These results are shown in Figure 9. There we see that the recapture rate is similar to the transfer rate -- the ratio being between unity and a few. The ratio is within a factor of two when the planets are near the 6:5 or 7:6 MMR. This fact implies that for multihabitable systems, biological transfer from one planet to another is similar in importance to auto-panspermia where biolog- ical material is transferred across the surface of a planet by its own collision ejecta -- at least that portion of the ejecta that initially escapes the planet. 4.4. Effects of Caustic Flows One of our earlier claims is that there should be an excess of successive collisions separated by only a short time interval. Figure 10 shows the distribution of separa- 8 Steffen & Li Figure 8. The relative fraction of all successful transfers of col- lision ejecta (Nj+1:j /Nall), where N is the number of successful transfers, from one planet to the other as a function of the differ- ence between the orbital semi-major axes (the total of the y-values is unity for each case). The blue squares correspond to the stan- dard velocity distribution while the orange circles correspond to the restricted distribution. The dotted line shows that for our two ini- tial velocity distributions the relative fraction falls as a2 − a1−2/3 (only the normalization was allowed to float in a fit to the results of the standard distribution). We believe the fact that these two examples are near this line to be a coincidence (see text). Never- theless, these results confirm the expectation that the closer the two planets are to each other, the more likely it is for material to transfer from one to the other. Figure 10. Top: PDF of time differences between impacts for the 3:2 MMR pair, where the outer planet is the source, using our standard velocity distribution (outline) and using the restricted dis- tribution (orange, solid). The dashed line corresponds to the best fit exponential distribution to the standard velocity case. Bottom: Similar plot but for the 6:5 MMR pair and with the inner planet being the source. The excess of collisions at short time intervals in both of these plots shows the non-Poisson nature of the im- pacts. Specifically, many collisions occur in quick succession -- a result consistent with the expected behavior from caustic flows. Results for all other MMR pairs and both sources are qualitatively similar to these. cating a non-Poisson component consistent with caustic flows. The simulations that used the restricted veloc- ity distribution show a very large excess of short time intervals between collisions. This result is due to the much higher concentration of the particle velocities -- giving a phase space that is much more densely filled. Such flows would imply that multiple successful trans- fers are likely to occur in quick succession, spreading any biological material -- which came from the same location of the source planet -- across a range of locations on the destination planet. 5. DISCUSSION Given our results above, we offer a few observations about life in multihabitable systems. Both issues consid- ered here turned out to be favorable for life and its trans- fer among multiple bodies in the nominal habitable zone of a multihabitable system. Only in special situations do the orbital obliquity of the close-proximity planets vary by large amounts. Simultaneously, the probability of material being shared between the planets is much larger than it is for the solar system. The different dependencies that we observed in our Figure 9. Ratio of recaptured partices to transferred particles as a function of the difference in semi-major axis. The blue circles (solid line) are when the inner planet is the source using the standard velocity distribution. The orange squares (dashed) is for standard velocities when the outer planet is the source. The green diamonds (dotted) and red triangles (dot-dashed) are for the restricted ve- locity distribution for the inner and outer planets, respectively, as the source. tion times between successive collisions for two example cases -- the non-resonant 3:2 pair using the outer planet as the source and the 6:5 pair using the inner planet as the source. We examine results from both the standard velocity distribution and the restricted distribution. Also shown is the best-fitting exponential distribution to the standard velocity case. For the most part, near the long-separation tail of the distribution, the time differences do follow an exponen- tial distribution -- showing that a portion of the time dis- tribution follows from Poisson fluctuations. At short separations there is a slight to significant excess, indi- Multihabitable Systems 9 simulations (especially the preference for low ejection ve- locities) bodes well for the successful transfer biological material for several reasons. First, since smaller veloc- ity particles are more readily transferred to the other planet, the collision that produces the ejecta can be less energetic -- reducing the risk of destroying important ma- terial. Second, lower particle velocities mean that more of the surface of the planet is effective as a source for the ejecta (high velocity particles had large changes in suc- cess rates that varied with longitude while low velocity particles were uniformly distributed in longitude). Third, since the ejecta forms caustics with phase-space sheets having high particle densities, the successful trans- fer of one particle implies an increased probability that additional particles will also transfer successfully and within a relatively short amount of time. Thus, one can imagine material from one part of the source planet be- ing distributed almost simultaneously across much of the surface of the destination planet -- giving the seeds of life a greater chance of taking hold. The same effect is less likely to occur in more widely separated planet planetary systems (such as in the solar system), which rely on sec- ular resonances to excite orbital eccentricities, because chaotic behavior near those resonances would, over long timescales, cause neighboring orbits to diverge. Not only will panspermia be more common in a multi- habitable system than in the solar system, but the close proximity of the planets to each other within the hab- itable zone of the host star allows for a real possibility of the planets having regions of similar climate -- perhaps allowing the microbiological family tree to extend across the system. There are many things to consider in mul- tihabitable systems, especially in the cases where intelli- gent life emerges. For panspermia in a multihabitable system, the trans- fer rate of ejecta from one planet to the other is com- parable to the rate of auto-panspermia. In all of our simulations the rate at which material is recaptured by the source planet is a factor of one to a few times the rate that material transfers between the two planets (see Figure 9) but the ratio is not 10 or 100. Thus, in these systems one may find a significant amount of exchange between the two planet surfaces. 6. CONCLUSIONS The prospect of planetary systems containing multiple habitable planets is intriguing. And, given some of the results of the Kepler mission, we may expect that such systems exist. As a full census of habitable-zone planets comes ever closer to reality, we may have the opportu- nity to study one or a few of these systems in some detail. There are many issues that have an impact on life in a planetary system. Here we have only considered the ef- fects of large obliquity variations due to strong dynamical interactions -- which could hamper or preclude the devel- opment of intelligent life due to its associated changes in climate -- and the possibility of frequent lithopanspermia in each system. We note that our models did not con- sider possible effects of tides, spin-orbit resonances, or the potentially strong effect of inclination resonances in the systems. We found that obliquity variations are generally not affected by the close proximity of the planets in a multi- habitable system. Also, obliquity variations of close pairs embedded in the solar system or of potentially habitable planets in a system with a close pair were not sufficient to significantly reduce the probability of having a sta- ble climate. Only in cases where the inclination modal frequencies coincide with the planetary precession fre- quency did large obliquity variation arise. At the same time, we expect to find relatively high rates of panspermia in these systems. The nearer the planets are to one another the higher the success rate of ejecta transfer -- coming quite close to the same rate as ejecta falling back to the surface of the originating planet. Also, we found that the smaller the velocity of the ejected material the more uniformly they can be sourced across the originating planet. With high velocity ejecta, the range of initial longitudes is constrained relative to the direction of motion. Mean-motion resonance did not have as strong of an ef- fect on the collision rate as the variations that occur from changes to the initial orbital parameters of the planets. By way of comparison, changes to the initial conditions of the collision ejecta (within the parameters of the ve- locity distribution) produced variations consistent with Poisson fluctuations over the 10 million year integration time. That is, for a given set of planet initial condi- tions, the transfer rate was robust given our simulation method. Finally, we claimed that the restricted phase space of the ejected particles (due to their originating from a highly localized source and their collisionless nature) implies that the particles should form caustic flows of high density sheets of ejecta. A consequence of such a dynamical state is non-Poisson fluctuations in the trans- fer rate occurring at short time intervals. We observe these fluctuations, which are especially pronounced for the restricted velocity dispersion where the phase space is much more densely sampled by our particles. ACKNOWLEDGEMENTS We thank Konstantin Batygin for many useful dis- cussions in the preparation of this work. J.H.S. ac- knowledges support from the Lindheimer Fellowship at CIERA, Northwestern and from NASA under grant NNX08AR04G issued through the Kepler Participating Scientist Program. REFERENCES J. C. Armstrong, C. B. Leovy, and T. Quinn. A 1 Gyr climate model for Mars: new orbital statistics and the importance of seasonally resolved polar processes. Icarus, 171:255 -- 271, October 2004. doi:10.1016/j.icarus.2004.05.007. J. C. Armstrong, R. Barnes, S. Domagal-Goldman, J. Breiner, T. R. Quinn, and V. S. Meadows. Effects of Extreme Obliquity Variations on the Habitability of Exoplanets. Astrobiology, 14: 277 -- 291, April 2014. doi:10.1089/ast.2013.1129. N. M. Batalha et al. Planetary Candidates Observed by Kepler. III. Analysis of the First 16 Months of Data. ApJS, 204:24, February 2013. doi:10.1088/0067-0049/204/2/24. K. Batygin and M. E. Brown. Early Dynamical Evolution of the Solar System: Pinning Down the Initial Conditions of the Nice Model. ApJ, 716:1323 -- 1331, June 2010. doi:10.1088/0004-637X/716/2/1323. E. Belbruno, A. Moro-Mart´ın, R. Malhotra, and D. Savransky. Chaotic Exchange of Solid Material Between Planetary Systems: Implications for Lithopanspermia. Astrobiology, 12: 754 -- 774, August 2012. doi:10.1089/ast.2012.0825. 10 Steffen & Li W. J. Borucki et al. Kepler Planet-Detection Mission: G. Li and K. Batygin. On the Spin-axis Dynamics of a Moonless Introduction and First Results. Science, 327:977 -- , February 2010. doi:10.1126/science.1185402. Earth. ApJ, 790:69, July 2014a. doi:10.1088/0004-637X/790/1/69. R. Brasser and K. J. Walsh. Stability analysis of the martian G. Li and K. Batygin. Pre-late Heavy Bombardment Evolution obliquity during the Noachian era. Icarus, 213:423 -- 427, May 2011. doi:10.1016/j.icarus.2011.02.024. R. Brasser, S. Ida, and E. Kokubo. A dynamical study on the habitability of terrestrial exoplanets - II The super-Earth HD 40307 g. MNRAS, 440:3685 -- 3700, June 2014. doi:10.1093/mnras/stu555. C. J. Burke et al. Planetary Candidates Observed by Kepler IV: Planet Sample from Q1-Q8 (22 Months). ApJS, 210:19, February 2014. doi:10.1088/0067-0049/210/2/19. J. A. Carter, E. Agol, et al. Kepler-36: A Pair of Planets with Neighboring Orbits and Dissimilar Densities. Science, 337: 556 -- , August 2012. doi:10.1126/science.1223269. of the Earth's Obliquity. ApJ, 795:67, November 2014b. doi:10.1088/0004-637X/795/1/67. J. J. Lissauer, J. W. Barnes, and J. E. Chambers. Obliquity variations of a moonless Earth. Icarus, 217:77 -- 87, January 2012. doi:10.1016/j.icarus.2011.10.013. J. J. Lissauer et al. Architecture and Dynamics of Kepler's Candidate Multiple Transiting Planet Systems. ApJS, 197:8, November 2011. doi:10.1088/0067-0049/197/1/8. J. J. Lissauer et al. Validation of Kepler's Multiple Planet Candidates. II. Refined Statistical Framework and Descriptions of Systems of Special Interest. ApJ, 784:44, March 2014. doi:10.1088/0004-637X/784/1/44. J. E. Chambers. A hybrid symplectic integrator that permits D. S. McKay, E. K. Gibson, Jr., K. L. Thomas-Keprta, H. Vali, close encounters between massive bodies. MNRAS, 304: 793 -- 799, April 1999. doi:10.1046/j.1365-8711.1999.02379.x. M. A. Chandler and L. E. Sohl. Climate forcings and the initiation of low-latitude ice sheets during the Neoproterozoic Varanger glacial interval. J. Geophys. Res., 105:20737 -- 20756, 2000. doi:10.1029/2000JD900221. C. S. Romanek, S. J. Clemett, X. D. F. Chillier, C. R. Maechling, and R. N. Zare. Search for Past Life on Mars: Possible Relic Biogenic Activity in Martian Meteorite ALH84001. Science, 273:924 -- 930, August 1996. doi:10.1126/science.273.5277.924. H. J. Melosh. The rocky road to panspermia. Nature, 332: G. Colombo. Cassini's second and third laws. AJ, 71:891, 687 -- 688, April 1988. doi:10.1038/332687a0. November 1966. doi:10.1086/109983. K. M. Deck, M. J. Holman, E. Agol, J. A. Carter, J. J. Lissauer, D. Ragozzine, and J. N. Winn. Rapid Dynamical Chaos in an Exoplanetary System. ApJ, 755:L21, August 2012. doi:10.1088/2041-8205/755/1/L21. L. Dones, B. Gladman, H. J. Melosh, W. B. Tonks, H. F. Levison, and M. Duncan. Dynamical Lifetimes and Final Fates of Small Bodies: Orbit Integrations vs Opik Calculations. Icarus, 142: 509 -- 524, December 1999. doi:10.1006/icar.1999.6220. L. R. Doyle et al. Kepler-16: A Transiting Circumbinary Planet. Science, 333:1602 -- , September 2011. doi:10.1126/science.1210923. F. P. Fanale, J. R. Salvail, W. B. Banerdt, and R. S. Saunders. Mars - The regolith-atmosphere-cap system and climate change. Icarus, 50:381 -- 407, June 1982. doi:10.1016/0019-1035(82)90131-2. L. M. Francois, J. C. G. Walker, and W. R. Kuhn. A numerical simulation of climate changes during the obliquity cycle on Mars. J. Geophys. Res., 95:14761 -- 14778, August 1990. doi:10.1029/JB095iB09p14761. B. J. Gladman, J. A. Burns, M. Duncan, P. Lee, and H. F. Levison. The Exchange of Impact Ejecta Between Terrestrial Planets. Science, 271:1387 -- 1392, March 1996. doi:10.1126/science.271.5254.1387. K. M. G´orski, E. Hivon, A. J. Banday, B. D. Wandelt, F. K. Hansen, M. Reinecke, and M. Bartelmann. HEALPix: A Framework for High-Resolution Discretization and Fast Analysis of Data Distributed on the Sphere. ApJ, 622:759 -- 771, April 2005. doi:10.1086/427976. J. D. Hays, J. Imbrie, and N. J. Shackleton. Variations in the Earth's Orbit: Pacemaker of the Ice Ages. Science, 194: 1121 -- 1132, December 1976. doi:10.1126/science.194.4270.1121. J. Imbrie. Astronomical theory of the Pleistocene ice ages - A brief historical review. Icarus, 50:408 -- 422, June 1982. doi:10.1016/0019-1035(82)90132-4. G. S. Jenkins. Global climate model high-obliquity solutions to the ancient climate puzzles of the Faint-Young Sun Paradox and low-latitude Proterozoic Glaciation. J. Geophys. Res., 105: 7357 -- 7370, 2000. doi:10.1029/1999JD901125. H. Kinoshita. First-Order Perturbations of the Two Finite Body Problem. PASJ, 24:423, 1972. J. Laskar. Large Scale Chaos and Marginal Stability in the Solar System. Celestial Mechanics and Dynamical Astronomy, 64: 115 -- 162, March 1996. doi:10.1007/BF00051610. J. Laskar and P. Robutel. The chaotic obliquity of the planets. Nature, 361:608 -- 612, February 1993. doi:10.1038/361608a0. J. Laskar, F. Joutel, and P. Robutel. Stabilization of the earth's obliquity by the moon. Nature, 361:615 -- 617, February 1993. doi:10.1038/361615a0. M. H. Lee and S. J. Peale. Dynamics and Origin of the 2:1 Orbital Resonances of the GJ 876 Planets. ApJ, 567:596 -- 609, March 2002. doi:10.1086/338504. H. J. Melosh. Exchange of Meteorites (and Life?) Between Stellar Systems. Astrobiology, 3:207 -- 215, January 2003. doi:10.1089/153110703321632525. H. J. Melosh and W. B. Tonks. Swapping Rocks: Ejection and Exchange of Surface Material Among the Terrestrial Planets. Meteoritics, 28:398 -- 398, July 1993. C. Mileikowsky, F. A. Cucinotta, J. W. Wilson, B. Gladman, G. Horneck, L. Lindegren, J. Melosh, H. Rickman, M. Valtonen, and J. Q. Zheng. Natural Transfer of Viable Microbes in Space. 1. From Mars to Earth and Earth to Mars. Icarus, 145:391 -- 427, June 2000. doi:10.1006/icar.1999.6317. F. Mullally et al. Planetary Candidates Observed by Kepler VI: Planet Sample from Q1-Q16 (47 Months). ArXiv e-prints:1502.02038, February 2015. T. Nakamura and E. Tajika. Climate change of Mars-like planets due to obliquity variations: implications for Mars. Geophys. Res. Lett., 30:1685, July 2003. doi:10.1029/2002GL016725. O. Neron de Surgy and J. Laskar. On the long term evolution of the spin of the Earth. A&A, 318:975 -- 989, February 1997. J. B. Pollack and O. B. Toon. Quasi-periodic climate changes on Mars - A review. Icarus, 50:259 -- 287, June 1982. doi:10.1016/0019-1035(82)90126-9. J. F. Rowe et al. Planetary Candidates Observed by Kepler. V. Planet Sample from Q1-Q12 (36 Months). ApJS, 217:16, March 2015. doi:10.1088/0067-0049/217/1/16. P. Sikivie. Caustic rings of dark matter. Physics Letters B, 432: 139 -- 144, July 1998. doi:10.1016/S0370-2693(98)00595-4. A. Soto, M. A. Mischna, and M. I. Richardson. Climate Dynamics of Atmospheric Collapse on Ancient Mars. In Lunar and Planetary Institute Science Conference Abstracts, volume 43 of Lunar and Planetary Institute Science Conference Abstracts, page 2783, March 2012. D. S. Spiegel, K. Menou, and C. A. Scharf. Habitable Climates: The Influence of Obliquity. ApJ, 691:596 -- 610, January 2009. doi:10.1088/0004-637X/691/1/596. J. H. Steffen et al. Five Kepler Target Stars That Show Multiple Transiting Exoplanet Candidates. ApJ, 725:1226 -- 1241, December 2010. doi:10.1088/0004-637X/725/1/1226. O. B. Toon, J. B. Pollack, W. Ward, J. A. Burns, and K. Bilski. The astronomical theory of climatic change on Mars. Icarus, 44:552 -- 607, December 1980. doi:10.1016/0019-1035(80)90130-X. J. Touma and J. Wisdom. The chaotic obliquity of Mars. Science, 259:1294 -- 1297, February 1993. doi:10.1126/science.259.5099.1294. A. D. Vernekar. a Study of Mean Temperature of the Earth's Surface. In Atmospheric Radiation, page 228, 1972. W. R. Ward. Large-Scale Variations in the Obliquity of Mars. Science, 181:260 -- 262, July 1973. doi:10.1126/science.181.4096.260. Multihabitable Systems 11 J. Weertman. Milankovitch solar radiation variations and ice age ice sheet sizes. Nature, 261:17 -- 20, May 1976. doi:10.1038/261017a0. L. E. Wells, J. C. Armstrong, and G. Gonzalez. Reseeding of early earth by impacts of returning ejecta during the late heavy bombardment. Icarus, 162:38 -- 46, March 2003. doi:10.1016/S0019-1035(02)00077-5. W. F. Welsh et al. Transiting circumbinary planets Kepler-34 b and Kepler-35 b. Nature, 481:475 -- 479, January 2012. doi:10.1038/nature10768. D. M. Williams and J. F. Kasting. Habitable Planets with High Obliquities. Icarus, 129:254 -- 267, September 1997. doi:10.1006/icar.1997.5759. R. J. Worth, S. Sigurdsson, and C. H. House. Seeding Life on the Moons of the Outer Planets via Lithopanspermia. Astrobiology, 13:1155 -- 1165, December 2013. doi:10.1089/ast.2013.1028. J.-W. Xie. Transit Timing Variation of Near-resonance Planetary Pairs: Confirmation of 12 Multiple-planet Systems. ApJS, 208: 22, October 2013. doi:10.1088/0067-0049/208/2/22. APPENDIX HAMILTONIAN EQUATIONS The Hamiltonian we use for our simulations (e.g. Colombo 1966; Kinoshita 1972; Laskar et al. 1993; Touma and Wisdom 1993; Neron de Surgy and Laskar 1997; Armstrong et al. 2004) is given by: (cid:112) H(χ, ψ, t) = 1 2 αχ2 + 1 − χ2 (A(t) sin ψ + B(t) cos ψ)) , (A1) where χ = cos ,  is the obliquity, ψ is the longitude of the spin axis, α is the precession coefficient defined as (Neron de Surgy and Laskar 1997), and α = √ 3G 2ω (a m∗ 1 − e2)3 Ed, A(t) = 2( q + p(q p − p q))/ B(t) = 2( p − q(q p − p q))/ 1 − p2 − q2, 1 − p2 − q2, (cid:112) (cid:112) (A2) (A3) (A4) where p = sin i/2 sin Ω and q = sin i/2 cos Ω. A(t) and B(t) are obtained using the aforementioned results of the orbital elements from our MERCURY simulations. In the expression for α, m∗ is the mass of the star, a and e are the semi-major axis and the eccentricity of the planet's orbit, ω is the spin rate of the planet, Ω is the longitude of ascending node, and Ed is the dynamical ellipticity of the planet. Table 2 shows the initial orbital elements we used for our simulations. Prior to including the test particle ejecta we transformed the positions and velocities such that the orbital distance of the inner planet was 1 AU. In all simulations, the mass of the central star is 1 M(cid:12) and the mass of the planets are 1.0× 10−6M(cid:12) and the planet densities are 1 g/cc. INITIAL CONDITIONS Table 2 Simulation initial conditions. Run 3:2 resonant 3:2 nonresonant N 3:2 nonresonant I 4:3 resonant 4:3 nonresonant N 4:3 nonresonant I 6:5 resonant 6:5 nonresonant N 7:6 resonant 7:6 nonresonant N Planet 1 2 1 2 1 2 1 2 1 2 1 2 1 2 1 2 1 2 1 2 a 0.996185E+00 0.130541E+01 0.996185E+00 0.130541E+01 0.996185E+00 0.130541E+01 0.964821E+00 0.116885E+01 0.964821E+00 0.116885E+01 0.964821E+00 0.116885E+01 0.953590E-01 0.107684E+00 0.953590E-01 0.107684E+00 0.976083E-01 0.108174E+00 0.976083E-01 0.108174E+00 e 0.217249E-01 0.234200E-01 0.217249E-01 0.234200E-01 0.217249E-01 0.234200E-01 0.218355E-01 0.223782E-01 0.218355E-01 0.223782E-01 0.218355E-01 0.223782E-01 0.199269E-01 0.207574E-01 0.199269E-01 0.207574E-01 0.182103E-01 0.189198E-01 0.182103E-01 0.189198E-01 i 0.997871E+00 0.700569E+00 0.997871E+00 0.700569E+00 0.997871E+00 0.700569E+00 0.685909E+00 0.983627E+00 0.685909E+00 0.983627E+00 0.685909E+00 0.983627E+00 0.983127E+00 0.721238E+00 0.983127E+00 0.721238E+00 0.994049E+00 0.705235E+00 0.994049E+00 0.705235E+00  0.346938E+03 0.165285E+03 0. 0. 0. 0.165285E+03 0.207332E+03 0.125833E+02 0. 0. 0. 0.125833E+02 0.232531E+02 0.202503E+03 0. 0. 0.128348E+03 0.307745E+03 0. 0. Ω 0.301425E+00 0.359627E+03 0. 0. 120. 0.359627E+03 0.455775E-02 0.359761E+03 0. 0. 120. 0.359761E+03 0.445968E+01 0.354280E+03 0. 0. 0.358880E+03 0.150213E+01 0. 0. M 0.162046E+03 0.500250E+02 0. 0. 240. 0.500250E+02 0.214100E+03 0.354742E+03 0. 0. 240. 0.354742E+03 0.177217E+03 0.148131E+03 0. 0. 0.105195E+03 0.270803E+03 0. 0.
1807.00063
1
1807
2018-06-29T20:38:40
Fine-grained Material Associated with a Large Sulfide returned from Comet 81P/Wild 2
[ "astro-ph.EP" ]
In a consortium analysis of a large particle captured from the coma of comet 81P/Wild 2 by the Stardust spacecraft, we report the discovery of a field of fine-grained material (FGM) in contact with a large sulfide particle. The FGM was partially located in an embayment in the sulfide, so appears to have been largely protected from damage during hypervelocity capture in aerogel. Some of the FGM particles are indistinguishable in their characteristics from common components of chondritic-porous interplanetary dust particles (CP-IDPs), including glass with embedded metals and sulfides (GEMS) and equilibrated aggregates (EAs). The sulfide exhibits surprising Ni-rich lamellae, which may indicate that this particle experienced a long-duration heating event after its formation but before incorporation into Wild 2. We discuss the relationship of the FGM to the sulfide, to other Wild 2 particles and to the history of the Solar nebula.
astro-ph.EP
astro-ph
Fine-grained Material Associated with a Large Sulfide returned from Comet 81P/Wild 2 July 3, 2018 Z. Gainsforth1,†, A. J. Westphal1, A. L. Butterworth1, C. E. Jilly-Rehak1, D. E. Brownlee2, D. Joswiak2, R. C. Ogliore3, M. E. Zolensky4, H. A. Bechtel5, D. S. Ebel6, G. R. Huss7, S. A. Sandford8, A. J. White9 1Space Sciences Laboratory, University of California, Berkeley, CA 94720, 2Dept. of Astronomy, University of Washington, Seattle, WA 98195, 3Department of Physics, Washington University in St. Louis, St. Louis, MO, 63117, 4ARES, NASA Johnson Space Center, Houston, TX 77058, 5Advanced Light Source, Lawrence Berkeley Laboratory, Berkeley, CA 94720, 6Dept. Earth Planet. Sci., American Museum Natural History, NY, NY 10024, 7University of Hawai'i at Manoa, Honolulu, HI 96822, 8NASA Ames Research Center, Moffett Field, CA 94035, 9Dept. Astro. and Planet. Sci., University of Colorado, Boulder, CO 80309 †e-mail: [email protected] 8 1 0 2 n u J 9 2 . ] P E h p - o r t s a [ 1 v 3 6 0 0 0 . 7 0 8 1 : v i X r a 1 Abstract In a consortium analysis of a large particle captured from the coma of comet 81P/Wild 2 by the Stardust spacecraft, we report the discovery of a field of fine-grained material (FGM) in contact with a large sulfide particle. The FGM was partially located in an embayment in the sulfide, so appears to have been largely protected from damage during hypervelocity capture in aerogel. Some of the FGM particles are indistinguishable in their characteristics from common components of chondritic-porous interplanetary dust particles (CP-IDPs), including glass with embedded metals and sulfides (GEMS) and equilibrated aggregates (EAs). The sulfide exhibits surprising Ni-rich lamellae, which may indicate that this particle experienced a long-duration heating event after its formation but before incorporation into Wild 2. We discuss the relationship of the FGM to the sulfide, to other Wild 2 particles and to the history of the Solar nebula. 2 INTRODUCTION Anhydrous, fine-grained ("chondritic-porous") interplan- etary dust particles (CP-IDPs) are regularly captured by aircraft-borne collectors in the stratosphere. Several lines of evidence point to a cometary origin for a subset of CP-IDPs (Love and Brownlee 1991; Brownlee et al. 1993; Joswiak et al. 2007), and while there is substan- tial evidence that some specific CP-IDPs are cometary (Joswiak et al. 2017), there is no hard link between spe- cific CP-IDPs and specific comets. CP-IDPs commonly contain 100-500 nm amorphous silicates known as GEMS (Glass with Embedded Metals and Sulfides), and crys- talline silicate aggregates of similar size and larger called Equilibrated Aggregates (EAs). The relationship, if any, between GEMS and EAs is unclear, but a subset of EAs have compositions similar to GEMS and it has been sug- gested that many EAs are GEMS that have been heated and annealed (Brownlee et al. 2005; Keller and Messenger 2009). In addition to this fine-grained material (FGM), CP-IDPs also contain larger crystalline silicates, sulfides and, more rarely, metal. The Stardust mission returned ∼300 µg of material from the coma of Jupiter-family comet Wild 2 to ter- restrial laboratories, in a collector composed of aerogel and aluminum foil (Brownlee et al. 2006; Brownlee 2013; Horz et al. 2006). There was an expectation that the Stardust collection would include abundant GEMS, EAs, and pre-solar grains. The collection was found to contain abundant crystalline particles (cid:29) 1 µm in size (Brownlee et al. 2006), but evidence for GEMS and EAs has been elu- sive, probably because of the violence of the hypervelocity capture process and the fragility of FGM. Pre-solar grains have been reported, but the abundance is difficult to esti- mate accurately because of the uncertainties in survival efficiency (Floss et al. 2013). Nevertheless, several particles extracted from the aero- gel collectors have shown evidence for FGM adhering to their peripheries. These include Iris (C2052,12,74), a ≈15 µm chondrule-like object with associated primi- tive sulfides and an enstatite whisker (Stodolna et al. 2014), and Cecil (C2062,2,162), a particle containing a chondrule-like object and fine-grained sulfides, spinels, and pyroxenes. In Cecil, the fine grained material was largely embedded in glass suggesting that some heating alteration may have occurred during aerogel capture, but several of the fine grained components retained primitive signatures despite the apparent heating (Gainsforth et al. 2014b, 2015b). Joswiak et al. (2012) studied a sulfide particle, Febo (C2009,2,57), with FGM adhering to one side. This FGM showed strong evidence for melting, probably caused by aerogel capture. However, 15N-rich organics survived in select locations within the fine-grained material implying that at least some of the material in the Febo FGM retains chemical signatures from before aerogel capture (Matrajt et al. 2008). Here we report the discovery and analysis of a par- ticle named Andromeda (C2086,22,191) returned from comet 81P/Wild 2 by the Stardust spacecraft. It consists of a field of fine-grained material (FGM) in association with a large sulfide particle. Some FGM appears to have been protected from significant damage during hyperve- locity capture in aerogel and has strong affinities with components of CP-IDPs, including GEMS and EAs. METHODS Andromeda is one of at least seven terminal particles of type-B (Burchell et al. 2008) track 191 (C2086,22,191, Fig. 1A and B). It was located at a depth of 8.8 mm from the space-exposed surface of the aerogel collector, and consisted of a 15 x 15 x 20 µm iron sulfide with abundant FGM aggregated in an embayment of the sulfide (Fig. 2). The particle was extracted in an aerogel "keystone" (West- phal et al. 2004), then embedded in EMBED 812 epoxy on the end of an epoxy bullet. We used an ultramicro- tome to cut ≈100 nm thick sections, which we placed onto Cu TEM grids with a 10 nm amorphous carbon substrate (Ladd Research). We studied Andromeda using an FEI Titan transmis- sion electron microscope (TEM) operated at 80 keV and 200 keV at the Molecular Foundry, Lawrence Berkeley National Laboratory (LBL). A 4-element Bruker silicon drift detector with a solid angle of 0.6 sr provided energy- dispersive X-ray spectroscopic (EDS) mapping in the Ti- tan at count rates between 5 and 100 kcps. We used the Cliff-Lorimer approximation (Cliff and Lorimer 1975) with a thin film correction as described in Gainsforth (2016). Most spectra had on the order of 106 counts after back- ground subtraction yielding detection limits <0.03 at% for most elements (Tables 1 – 4). In the case of the large sul- fide given by spectrum Sulfide02 in Table 1 (abbreviated as: Table 1, Sulfide02), a longer spectrum of 3×108 counts allowed for quantification at hundreds of ppm concentra- tion for several elements for which backgrounds were low. For sensitive phases that lose Na during TEM analysis, we acquired a sequence of EDS maps and then extrapo- lated the original Na content based on the observed loss rate. For the sulfide, the dose was sufficiently high that we applied a S volatilization correction (Gainsforth et al. 2014a). In many cases, it was not possible to fully remove the epoxy background from spectra and so we expect some O concentrations are a few percent too high. For quan- tification of Fe and O in iron oxides, the excess O from epoxy could have significantly altered the conclusions, so we subtracted a nearby epoxy spectrum by normalizing to Cl when possible, or C when the Cl signal was too weak. We compared compositions of the fine grained material (FGM) to Solar System abundances from Lodders (2003). In most cases we chose to normalize against Mg rather than Si in order to the avoid possible confusion with SiO2 3 Figure 1. A) Optical mosaic image of track 191 produced from image stack at varying focus. Andromeda was found at the end of an 8.8 mm track. The arrow connects the location of Andromeda in the track to panel B. B) Optical zoom-in of Andromeda. C) XRD topograph of Andromeda showing that the sulfide is a single crystal but with three subgrains spaced in 0.35◦ increments. Due to sample preparation and beamline geometry panel C is not at the same viewing angle as panels A and B. dilution from aerogel. To express elemental ratios with respect to protosolar values, we adopt the notation from astronomy: (cid:18) XA/XB (cid:19) (XA/XB)(cid:12) [A/B](cid:12) = log10 (1) where (cid:12) indicates the accepted protosolar values from Lodders (2003). Therefore, an A/B ratio which is exactly protosolar would mean [A/B](cid:12) = 0, a 10× enhancement in A relative to B gives [A/B](cid:12) = 1 and a 3× depletion in A relative to B gives [A/B](cid:12) = −0.5. We carried out white-beam X-ray microdiffraction (µXRD) analyses on beamline 12.3.2 of the Advanced Light Source (ALS) at LBL (Tamura et al. 2009). The µXRD beam was generated by a superconducting bend magnet and then collimated to a spot size of approxi- mately 0.5 µm × 0.5 µm with Kirkpatrick-Baez mirrors and slits. We recorded the XRD patterns on a Dectris Pilatus 1M camera. By acquiring 2D spatial maps with a diffraction pattern at each position, we were able to ana- lyze the crystal distortion as a function of its position. By tracking the presence and location of a single diffracted beam in each of the diffraction patterns, we were able to produce topographs: images of the crystal locations and distortions much like dark-field electron microscopy does in a TEM. For details on the method in general, see Black and Long (2004), and for details on our specific implementation see Tamura (2014) and Gainsforth et al. (2013a). For estimation of gas temperatures during capture in aerogel, we computed dissociation energies for aerogel and troilite into gases using Density Functional Theory (DFT) implemented in the plane wave formalism as part of the Quantum Espresso suite (Giannozzi et al. 2009). We used Perdew-Burke-Ernzerhof pseudopotentials from the Standard Solid State Pseudopotentials Efficiency database (Perdew et al. 1996; Corso 2014; Vanderbilt 1990; Prandini et al. 2018). We used single-unit-cell quartz and troilite structures for solid SiO2 and FeS, and a 5 Å cubic vol- ume for molecules. Quartz was chosen as a proxy for aerogel since the two are expected to have formation en- ergies within a few hundred meV of each other based on examination of SiO2 structures in the Materials Project (Jain et al. 2013), and examination of thermodynamic constants from the NIST-JANAF thermodynamic tables (Chase 1998). Convergence criteria were set to ∆E ≤ 40 meV in accordance with the expected accuracy of the pseudopotentials (Lejaeghere et al. 2016). Computation was done on the Vulcan, Nano and Etna compute clusters at LBL. 4 Figure 2. A) Mosaic of brightfield TEM images of Andromeda showing the sulfide on the left, and fine-grained material (FGM) on the right. Objects in the FGM are labelled according to their names in Table 2. B) EDS map of the FGM with iron (red), nickel (green), and sulfur (blue). The Ni content in the FGM is significantly higher than the Ni content of the primary sulfide impactor. C) EDS map of the FGM with iron (red), magnesium (green), and silicon (blue). Mg-rich grains include equilibrated aggregates, GEMS and altered objects. The interstitial space is filled with epoxy. RESULTS Large Sulfide µXRD analysis showed that the sulfide was a polygonalized crystal, consisting of three domains rotated by 0.35◦ from each other. Fig. 1C shows a topograph produced from a Laue map that shows the location and size of each domain. The reflections seen in XRD show multiple maxima within each reflection (expected for polygonalized crystals) and also show broadening of the peaks that indicate large internal stresses on the scale of 1◦ over the entire width of the crystal. The diffraction pattern fits pyrrhotite 4C. Because the internal strains are so large, it is not possible to differentiate troilite from pyrrhotite on the basis of the unit cell shape. The (Fe+Ni)/S ratio measured by TEM/EDS was very close to 1:1, as expected for troilite, not pyrrhotite 4C (Table 1). Ni was strongly depleted with respect to Fe: [Ni/Fe](cid:12) < −1. Such Ni depletion is also seen in several other Stardust sulfides (Joswiak et al. 2012; Gainsforth et al. 2013b). [S/Se](cid:12) = 0.02, indicating no significant fractionation of selenium from the protosolar abundance (S/Se = 6.8·103). High-resolution EDS maps showed that Ni is concentrated in bands no more than a few tens of nm thick rather than homogeneously distributed throughout the sulfide (Fig. 3A). The observed variability of Ni-rich band widths may be consistent with varying viewing angles of thin (≤ 30 nm) lamellae within the Ni-poor sulfide matrix. The average of three linescans across Ni-rich bands is shown in Fig. 4 as the difference in composition from ideal troilite. Ni and Fe are anti-correlated in the Ni- rich region. S appears to increase in the Ni-rich bands, but the deviation is no more than the deviation in S content outside the Ni-rich region and could be due to systematic error due to variable thickness of the sulfide. The linescan 5 Figure 3. A) EDS map showing troilite (purple) with Ni-rich bands (green). B) HRTEM image down the 110 troilite axis with a Ni-rich band on the left and a Ni-poor region on the right. The boxes show the regions used to produce FFTs of each region (shown in D). C) SAED of region including Ni-rich and poor regions down the 110 (troilite) or 100 (pyrrhotite 4C) zone axis. Pyrrhotite 4C superlattice reflections (4C) are marked alongside troilite (Tr). D) HRTEM FFT shows the same pattern. Purple spots are from the FFT of the Ni-poor region (box shown in B) and green shows the FFT from the Ni-rich region (box shown in B). The Ni-rich FFT shows the pyrrhotite 4C superlattice reflections. is limited by the spatial resolution of the EDS map at ≈30 nm. Selected Area Electron Diffraction (SAED) showed that the structure is best described by a superposition of troilite (zone 110) and pyrrhotite 4C (zone 100), though the pyrrhotite superlattice reflections are weak in com- parison to those of the troilite (Fig. 3C). HRTEM of the Ni-rich bands down the 110 zone axis of troilite show that they exhibit the pyrrhotite 4C superlattice reflection, and the Ni-poor regions do not (Fig. 3D). We measured the broadening of the 11n family of superlattice reflections after subtracting the instrumental response (22n) using the method of Gainsforth et al. (2017b) and we found the spread to be 0.28 nm−1 in the c* direction. Fine-grained Material Andromeda contained a field of FGM, partially embayed in the large sulfide, with ≈50 Mg-rich particles in the ultramicrotomed section shown in Fig. 2. We estimate that several hundred such particles were probably present Figure 4. Average of three linescans across Ni-rich bands. The Ni content is higher and the Fe content lower in the Ni-rich bands. The variability of S across bands is not clear. in total, considering that the microtomed section is only a 100-200 nm thick sampling of a 3D volume. TEM and STXM analysis showed that the embedding epoxy penetrated the interstial spaces in the FGM during sample preparation, indicating that the material was porous prior to embedding. When normalized to Si and protosolar ratios, we find that the FGM is depleted in Na, Mg, Al, Ca, Cr, Mn and Ni, and enriched in P, S, Fe and K (Table 2, FGM09). If the FGM had protosolar elemental abundances before capture, we calculated an upper limit of 1% contamination by SiO2 and 30% by FeS. K was present at ≈2 times protosolar abundance, normalized to Si, but no K-rich minerals were observed. We studied five classes of particles in the FGM field, discussed in detail below: GEMS-like objects, EA-like objects, Fe oxides, kosmochloric pyroxene and sulfides. GEMS-like objects The mean composition of GEMS differs from that of the Solar System (Keller and Messenger 2011). Keller and Messenger (2011) and Messenger et al. (2015) measured compositions of 287 GEMS, from which we calculated mean compositions and standard deviations for each ele- ment. We used a χ2 test to determine if objects within the FGM were consistent in their elemental composition with GEMS: (cid:88) (X − ¯XGEMS)2 σ2 GEMS (2) χ2 = elements where X is the atomic fraction of an element in the particle, ¯XGEMS and σGEMS are the mean and standard deviation respectively of the atomic elemental fractions, measured in a large number of GEMS from Keller and 6 Table 1. Sulfide compositions from TEM EDS (atomic %). Spectrum S Fe Ni Cr Atomic % ppm Mn Se ΣCations (S + Se) τ (nm g/cm3) Diffraction [Ni/Fe]a(cid:12) [S/Se]a(cid:12) Notes 350 70 0.15 0.10 0.13 0.48 0.97 6.93 0.60 49.99 49.87 49.93 49.52 49.25 39.73 52.12 49.86 49.99 49.95 50.00 49.78 53.35 47.14 Andromeda: Sulfide01 Sulfide02 Sulfide03 Sulfide04 Sulfide05 Sulfide06 Sulfide07 Febo: Sulfide08 160 aRatio normalized to Solar System abundances, see text. bThe thickness correction has been chosen to optimize for troilite. cThis idealized composition of the pyrrhotite bands was calculated in the discussion by combining HRTEM, diffraction and EDS. 1.01 1.00 1.00 1.00 1.01 0.88 1.12 0 279 358 358 358 n/a 300 Troiliteb Pyrrhotiteb -1.3 -1.5 -1.3 -0.8 -0.5 -0.7 -0.5 53.34 45.76 0.79 280 620 0.87 450 nC -0.3 0.02 Cliff-Lorimer High-counts Fe matrix Ni band Highest Ni Calculatedc P detected Table 2. Nanophase object compositions from TEM EDS (atomic %). Atomic % # O Na Mg Al Si P S K Ca Ti Cr Mn Fe Ni τ O/Si (Mg+Al+ Ca+Fe) /Si Notes FGM09 FGM10 FGM11 FGM12 FGM13 FGM14 FGM15 FGM16 FGM17 FGM18 FGM19 FGM20 FGM21 52.92 61.29 64.49 59.59 43.52 66.17 62.90 60.55 62.16 62.37 57.64 65.24 58.75 0.27 0.12 0.41 0.32 0.04 0.80 b.d. 0.38 0.19 0.27 0.14 0.41 0.07 1.33 10.86 13.77 16.91 4.24 14.70 12.41 9.04 10.74 6.29 3.56 9.51 2.60 0.25 1.20 1.15 1.20 0.72 1.19 0.93 1.27 0.77 1.31 0.27 0.89 1.62 17.39 17.03 14.76 16.31 13.98 12.73 12.95 16.14 14.48 18.10 20.31 13.42 21.53 0.17 0.10 0.15 0.20 b.d. 0.22 0.12 0.12 0.44 0.08 0.039 0.25 0.30 11.60 3.47 0.37 0.46 17.51 0.93 0.80 2.68 0.73 1.46 7.65 1.07 1.33 0.13 0.16 0.19 0.17 0.09 0.07 0.07 0.10 0.12 0.14 0.03 0.21 0.38 0.57 0.49 0.49 0.22 0.37 0.26 0.58 1.31 0.86 1.72 0.21 0.94 0.38 b.d. 0.03 0.03 0.03 0.09 0.05 0.04 0.02 0.01 0.03 b.d. 0.02 0.06 0.04 0.14 0.16 1.13 0.06 0.09 0.21 0.16 0.42 0.12 0.04 0.15 0.32 0.03 0.06 0.15 0.15 19.11 b.d. 0.09 0.11 0.10 0.04 0.06 b.d. b.d. 15.17 4.82 3.73 3.29 0.25 2.55 8.62 7.93 8.95 7.74 9.90 7.81 12.08 0.14 0.23 0.15 0.03 0.03 0.19 0.29 0.22 0.02 0.34 0.15 0.06 0.61 300 300 300 300 300 300 300 300 300 300 300 300 300 3.04 3.60 4.37 3.65 3.11 5.20 4.86 3.75 4.29 3.45 2.84 4.86 2.73 1.00 1.02 1.30 1.33 0.40 1.47 1.74 1.21 1.47 0.94 0.69 1.43 0.77 FGM Bulk Zn det. Has Fe-Oxides Has Fe-Oxides Zn det., Has Fe-Oxides Messenger (2011) and Messenger et al. (2015). To com- pute χ2 we used atomic fractions of eight elements and report the reduced χ2 (χ2 ν, ν = 7 degrees of freedom) along with p, the probability that the composition is inconsistent with that of GEMS distribution. Since low-Z elements are suceptible to significant instrumental differences and are often more volatile, we have not included elements with Z ≤ Na in the computation of χ2. For the purpose of computing p, we made the simplifying assumption that the elemental abundances are normally distributed inde- pendent variables (e.g., we ignored the correlation between Fe and S). This assumption is conservative in the sense that it tends to overestimate p. GEMS and some EAs have compositions similar to each other and distinct from other materials. As can be seen in Table 2 we acquired quality EDS maps and images of about 10 objects within the FGM that have compositions close to GEMS. To estimate bias in our selection we used k-means clustering (Arthur and Vassilvitskii 2007) on the EDS map from Fig. 2 to identify 52 regions which could be GEMS-like, so we studied ∼20% of the GEMS-like objects in this field. We observed an object (FGM10) which had the ap- pearance and composition of a GEMS (Fig. 5, Table 2, FGM10, p = 0.2). The top and left edges were bounded by ≈100 nm euhedral sulfides. The core contained Fe-Ni metal inclusions 5-20 nm in size. Within the metal in- clusions, [Ni/Fe](cid:12) was ≈ 0.3, while the external sulfide had [Ni/Fe](cid:12) ≈ 0.0. EDS maps show zoning of Mg and Ca within the core, a feature that has been noted within GEMS in IDPs (Keller and Messenger 2011; Joswiak et al. 1996). The sulfide on the lower right was more rounded and some sulfur was present in the nearby silicate por- tion, penetrating approximately half of the silicate. One metal-core/sulfide-shell structure was present in the sili- cate region containing the diffuse S. EA-like objects Two objects in the FGM field showed morphologies reminis- cent of equilibrated aggregates in CP-IDPs. FGM11 (Fig. 6) contained Mg-rich ≈ 10 nm silicate crystals. The sili- cate crystals showed mosaicity, that is, they were oriented within a few degrees of each other. It also contained Fe-Ni metals, sulfides, Cr-rich hotspots (probably chromite), and 7 Figure 5. A) BF TEM image of FGM10, a GEMS-like object in Andromeda's FGM. Euhedral sulfides are present on one side of the object. B) EDS map showing Fe(red), Mg (green), Si (blue). Mg shows distinct inhomogeneity. C) EDS map of showing Fe (red), Ni (green), S (blue, with a three pixel gaussian filter). Fe-Ni metals are present within the body of the object. Al-rich inclusions. The outline of the object was defined by the euhedral edges of the silicate crystals. There was no attached silica melt, anhedral sulfide, or any other phase to indicate alteration during capture. FGM12 was a second particle with an equilibrated aggregate morphology and contained Mg-rich silicate crys- tals. There were several Fe-Ni metal and sulfide inclusions, and no core-rim structures. The periphery was surrounded by euhedral chromite crystals, and was well defined by their shapes. Heated GEMS-like and EA-like objects Additional objects were present that had characteristics of GEMS or EAs, but showed indications of significant heating. FGM13 is morphologically reminiscent of GEMS but is compositionally inconsistent with GEMS (p > 0.99) because of a large excess of sulfide and Cr relative to the other elements. In the core are Fe-Ni metal inclusions with [Ni/Fe](cid:12) = -0.04. There was one euhedral sulfide on the left face with [Ni/Fe](cid:12) = -0.55. The remaining sulfide resided around the periphery, was morphologically nebulous, and conformally coated the object. A metal- core/sulfide-rim structure was visible in the silicate portion of the object. FGM14, FGM15 and FGM16 were all spherical objects 100 - 200 nm across with p = 0.5–0.8. They contained Fe-Ni metals and sulfides. They did not contain euhedral sulfides around the periphery, with the possible exception of one sulfide adjacent to FGM14. They did not show evidence for the core-rim structures caused by aerogel capture, and they did not show evidence for excess SiO2. FGM16 (Fig. 7) was apparently sintered to amorphous Fe-oxide. Additional Fe oxides containing Mg, Al, Si, P, S, Ca and Ni were present in the vicinity. These oxides are described in the next section and are a primary reason for 8 Figure 6. A) BF TEM image of FGM11, an equilibrated aggregate (EA) in Andromeda's FGM. B) EDS map of EA showing the location of Ni-rich metals (green) and the presence of sulfides (blue). C) EDS Map of EA showing the location of Cr-rich inclusions (green) and an Al-rich inclusion and surrounding Al-rich silicates (blue). considering these to be heated. FGM17 contained crystalline Mg-rich silicates and Cr- rich inclusions. It is associated with an Fe-oxide similar to the oxide adjacent to FGM16. Fe was also visible as a web-like structure weaving around the periphery of the object and outlining the crystals. There are no core-rim structures or excess silica. FGM18 was reminiscent of GEMS but contained a metal-core/sulfide-rim structure. There were no euhedral crystals around the periphery: instead, all crystals within and without were rounded. There was no evidence of crystalline silicate, but the Mg formed a web-like structure throughout the object interconnecting the Fe-Ni metals. The outer periphery of the object was rich in Ca. FGM19 was similar to common aerogel capture prod- ucts seen by Ishii et al. (2008) (Fig. 8). It contained multiple metal-core/sulfide-rim structures. We observed vesicular structures and substantial excess silica. The periphery was lined by amorphous web-like FeS. FGM20 was located at the boundary between the sulfide and FGM, touching the sulfide, and contained amorphous and poorly-crystalline objects. The outlines of Al-rich silicate, Ca-rich silicate, and Mg-rich silicate could be seen but were not crystalline. An amorphous silicate with Fe-Ni metal grains and Fe-Ni-S grains was attached to one side. Several small beam-sensitive, amorphous Fe-oxides were present around the periphery. FGM21 was also located at the sulfide/FGM boundary and contained the highest K and P concentration of any particle we measured. The P was largely concentrated into a ≈ 10 nm diameter Fe-Ni-P hotspot that could have been schreibersite, and appeared to have a core- shell structure in HAADF. Fe-Ni metals were also present, sulfur was diffused throughout the silicate, and no sulfides Figure 7. A) BF image of FGM16. The location marked (1) is the equilibrated aggregate. (2) is one of several Fe-oxides. B) EDS map showing the Si-rich aggregate (cyan) contrasted with the Fe-oxides (yellow). C) EDS map showing the presence of Ni-rich metals inside the equilibrated aggregate (green), the Fe-oxides (red) and sulfides (purple). Figure 8. A) BF image of FGM19. Vesiculated structure and Fe-core/sulfide-rim structures are visible. The periph- ery was surrounded by sulfides. The object is bisected by ultramicrotomy. B) EDS map showing Fe-cores (red) and sulfide rims (blue). The periphery also contains sulfides which have a web-like texture from volatilization during aerogel capture. C) EDS map showing vesicles in the silicate where Mg (green) is absent. were visible. The Mg and Al were present near the center but segregated from each other, and a 5 nm Ti hot spot was present in the Al-rich portion of the EDS map. Iron oxides in FGM Diffraction analyses of the Fe oxides near FGM16 showed that they were amorphous. Comparison of HAADF images taken before and after EDS mapping of the oxides around FGM20 showed that the Fe oxides were beam-sensitive. The Fe oxides were most prevalent at the interface between the FGM and the sulfide, and in the large voids between clumps of material within the FGM. The oxides were also present on the leading edge of the sulfide away from the FGM. They were not frequently found within clumps of FGM material. For comparison, we examined a magnetite rim from a lightly heated CP-IDP (Lambda, Cluster IDP L2071,17). The bright-field image shown in Fig. 9A shows that the rim was nanocrystalline, although it may also have included some amorphous regions. The EDS spectrum in Table 3 (FeOx29) shows abundances of Mg, Al, P, S and Ni ≤ 1 at%. We also studied Fe-O-Si smokes made by Nuth et al. (2000) shown in Fig. 9B. Here, amorphous and crystalline objects a few nm wide aggregated to produce smoke par- ticles. The EDS overlay in the image shows that the Fe-O-Si elements are heterogeneously distributed, unlike in Andromeda oxides. Figure 9. A) BF image of magnetite from the IDP Lambda. B) LRGB image of Fe-smoke made by Nuth et al. (2000). Luminance channel is HAADF, Fe (red), O (green), Si (blue). C) An Fe-oxide from FGM16 to scale for compari- son. All scale bars are 200 nm. Px30. Small amounts of P and S in the spectrum may have been due to neighboring glassy material and neighboring sulfide. Its diffraction was consistent with diopside, as should be expected based on its composition. The Na content was 0.159 cations per 6 oxygens, which was well balanced by Ti4+, Al3+ and Cr3+: Al + Cr + 2Ti = 0.174 cations per 6 oxygens, suggesting (Ti, Al, Cr)-Na coupled substitution. A small sulfide-metal appendage was found on one side of the pyroxene (Table 1, Sulfide07). The Ni/Fe ratio was significantly larger than the primary sulfide impactor. Kosmochloric pyroxene We observed a kosmochloric pyroxene (i.e., Na-Cr coupled substitution) in the FGM similar to those commonly found in other Wild 2 samples and in CP-IDPs (Joswiak et al. 2009). The elemental quantification is shown in Table 4 DISCUSSION Wild 2 FGM compared to GEMS Hypervelocity capture of particles in aerogel can produce objects that are morphologically reminiscent of GEMS 9 Table 3. Fe Oxide compositions from TEM EDS (atomic %). Atomic % # FeOx22 FeOx23 FeOx24 FeOx25 FeOx26 FeOx27 FeOx28 FeOx29 1Element for normalizing and subtracting the background. See methods. O 65.65 50.89 63.88 59.70 56.70 63.52 59.40 53.11 K b.d. b.d. b.d. b.d. b.d. 0.32 0.68 b.d. Mg b.d. b.d. b.d. b.d. 0.61 0.60 b.d. 0.41 S 1.64 1.28 2.23 1.10 1.47 1.89 2.60 0.69 Si 0.68 2.33 4.29 0.96 2.42 1.38 1.67 b.d. P det. 0.49 0.69 0.37 0.37 0.50 0.96 0.20 Na b.d. b.d. b.d. b.d. b.d. 0.93 1.46 b.d. Al 0.15 b.d. b.d. b.d. b.d. b.d. b.d. 0.18 Ca 0.34 0.58 0.63 0.47 0.43 1.95 2.16 b.d. Fe 31.49 44.44 28.28 37.40 38.00 28.92 31.07 44.53 Ni 0.06 b.d. b.d. b.d. b.d. b.d. b.d. 0.83 τ 300 300 300 300 300 300 300 300 O/Fe 2.08 1.15 2.26 1.60 1.49 2.20 1.91 1.19 Background1 Cl Cl C Cl Cl C C Cl Notes Near FGM16 Near FGM16 Near FGM16 Near FGM16 Near FGM16 Near FGM17 Near FGM17 Lambda magnetite. Cr det. Table 4. Pyroxene compositions from TEM EDS. Normalized oxide weight % Spectrum SiO2 Px30 56.14 TiO2 0.25 Al2O3 0.95 Cr2O3 MgO 4.23 14.81 CaO 16.26 FeO Na2O P2O5 1.00 3.94 2.28 S 0.14 Phaseb En40Fs07Wo33Ko14Ja06 Cations per 6 oxygens Na Spectrum Si Px30 0.16 aEn=enstatite, Fs=ferrosilite, Wo=wollastonite, Ko=kosmochlor, Ja=jadeite. bnm·g/cm3. Fe 0.12 Al 0.04 Cr 0.12 Ca 0.63 Ti 0.01 Mg 0.80 2.02 P 0.03 S 0.01 ΣCations 3.90 τ b 300 (Leroux et al. 2008a; Ishii et al. 2008). Four common signatures distinguish these objects from GEMS. 1) Excess silica is usually present due to mixing with the aerogel. In some cases the original compositions are still preserved except for the excess SiO2. 2) The rapid heating of sulfides reduces them to iron metal and subsequent rapid cooling recondenses the sulfur on the outside to form sulfide rims on the iron cores (Leroux et al. 2008b). 3) Vesicular structures from rapid volatilization are often observed. 4) Euhedral shapes on crystals, especially sulfides, are lost as they are liquified or volatilized. Keller and Messenger (2011) studied GEMS grains and noted that they had varying degrees of silicate polymeriza- tion. They hypothesized that this was related to nebular interactions with hydrogen and water which resulted in excess oxygen and OH within the silicate matrix. We plot the compositions of the GEMS-like and EA- like objects in Fig. 10 in a fashion modeled after Keller and Messenger (2011), their Fig. 7. Most of our objects show excess O which is likely an artifact of the surround- ing epoxy, so this analysis is not sensitive to the presence of OH. However, we still find a very similar trend to that observed by Keller and Messenger (2011), namely these objects have a variable silicate polymerization between SiO3 and SiO4. Aerogel dilution would push the compo- sition towards SiO2, but the distribution of our objects does not favor compositions more silica-rich than SiO3. The compositions of GEMS measured by Keller and Messenger (2011); Messenger et al. (2015) are shown in Fig. 11 as gray points. The Andromeda GEMS-like objects reported in this paper are also plotted in Fig. 11 and are consistent in elemental composition with GEMS. As expected, the bulk composition of the entire Andromeda FGM is not GEMS-like because it contains pyroxenes, sulfides, and other objects. Aerogel Capture: Dynamics and Thermal Environment Here we characterize the stopping dynamics and thermal environment experienced by Andromeda during capture in aerogel. Following Dominguez (2009), we assume a hydrodynamic model for hypervelocity aerogel capture - that is, the physics is identical to that of a projectile stopping in a gas (Fig. 12). This approach is justified because the hydrodynamic forces are much larger than the mechanical strength of the aerogel. In this regime, the range R of a spherical particle with radius rg, density ρ, and initial speed v, stopping in aerogel with density ρa, is R = λrg ln v vc (3) where vc is the speed at which the mechanical strength of the aerogel becomes important, which we take to be comparable to the sound speed in aerogel, ≈ 100 m s−1 (Dominguez 2009), and λ = (4/3)(ρ/ρa). This calculation underestimates the actual range, be- cause it only calculates the distance from the aerogel surface to the transition away from hydrodynamic stop- ping. After this transition, stopping is dominated by the crushing strength of the aerogel. At 6 km s−1 capture speed, this residual range is short compared to the total range. Because we have no information about the location of Andromeda within the original complex projectile - 10 Figure 10. The silicate polymerization of nanophase ob- jects from Andromeda's FGM is plotted after the style of Keller and Messenger (2011). The nanophase objects include the GEMS and EA objects as well as the thermally modified objects, yet only three of the studied Andromeda nanophase objects plot outside the field between SiO3 and SiO4 as might be caused by aerogel dilution (shown by red arrow). Points above the line show excess oxygen, and points below the line are deficient in oxygen relative to expected stoichiometries. Gray points are from a re- cent collection of IDPs not contaminated with silicone oil (Messenger et al. 2015). Figure 11. Comparison of thermally modified and pristine GEMS-like and EA-like objects within Andromeda's fine grained material (FGM) to GEMS from literature. Black diamonds indicate Andromeda nanophase object compo- sitions normalized to protosolar and Mg. Gray circles indicate GEMS compositions from Keller and Messenger (2011) (K&M) and Messenger et al. (2015) (M&K). Red triangles show the bulk composition of the Andromeda FGM which is distinctly not GEMS-like since it contains pyroxenes, sulfides, and other astromaterials. and the leading edge of the sulfide. In the case of a strong shock (M (cid:29) 1), the post-shock temperature and pressure are T ∼ 2(γ − 1) (γ + 1)2 m k and P ∼ 2 γ + 1 ρav2 = 5 6 ρav2 v2 = 5 36 m k v2 (4) (5) not to mention the difficulty of modeling the stopping of a complex, fragmenting particle in aerogel - we treat Andromeda as an isolated, robust, refractory object. Andromeda was found at the end of the second-longest of at least 7 terminal tracks identified in track 191 (Fig. 1), a large track near one edge of Stardust cometary tile C2086 (ρa = 0.028 g cm−3). Since Andromeda is dominated by sulfide, we take ρ = 4.6 g cm−3, so λ ≈ 220. With vi = 6.1 km s−1 and rg = 8 µm, the predicted range in this aerogel is 7.0 mm. The observed range was 8.8 mm, so the final ≈ 1.8 mm of the range was probably dominated by non- hydrodynamic stopping, but may have included a range increment if Andromeda was not located near the leading edge of the track 191 projectile, so that it was delayed in encountering the aerogel stopping medium. We calculated that 70% of the speed and 90% of the kinetic energy were lost within the first 2 mm of penetra- tion into the aerogel, within ≈0.6 µs after impact. This is the region in which most of the heating occurs. Because the projectile was hypersonic with Mach number (cid:29) 10, a strong shock just ahead of the leading edge of the pro- jectile heated and vaporized the aerogel capture medium where m is the molecular mass, v is the velocity, ρa is the aerogel density, γ = cp/cv is the adiabatic index, k is Boltzmann's constant and ∼ indicates an order of magnitude relationship (Shu 1992). Here we assume an adiabatic index γ = 1.4. This post-shock temperature corresponds to approximately 4 eV/molecule at maximum heating. We assume a bulk aerogel temperature of ∼ 200 K at the time of cometary dust capture. The impactor transfers energy that dissociates atoms and molecules from the aerogel and sulfide surface, and then heats the remaining solid and gas. We computed silicate and FeS dissociation energies using Density Functional Theory (DFT) and from thermodynamic paramterizations in the JANAF database and show the results in Table 5. If we further assume that the dissociation follows a Boltzmann distribution, i.e., exp−Ea/kT where kT = 4 eV, and Ea is the dissociation energy computed by DFT, or derived from JANAF, then we expect one-quarter to one-third of the silica and FeS to dissociate into molecular SiO2 and FeS. We also expect a few percent each of O2, O, Si, Fe and S. This will lead to very high oxygen and sulfur fugacities. 11 Table 5. Dissociation energies and abundances from DFT and JANAF at 0 K. Reaction SiO2(s) → SiO2(g) SiO2(s) → SiO(g) + O(g) SiO2(s) → Si(g) + O2(g) SiO2(s) → Si(g) + 2O(g) FeS(s) → FeS(g) FeS(s) → Fe(g) + S(g) 1Element for normalizing and subtracting the background. See methods. Dissociated % (DFT) 20 5 3 0 28 5 Energy (DFT, eV) 6.5 12.1 13.9 22.5 5.1 12.2 Energy (JANAF, eV) 6.2 10.9 14.0 19.1 4.9 8.2 Dissociated % (JANAF) 21 7 3 1 29 13 Figure 12. Hypersonic flow during capture. Aerogel meets the face of Andromeda's sulfide producing a shock (red) that vaporizes aerogel and FeS. Hot SiO2 and FeS gas flow around the particle. A turbulence region behind Andromeda causes some of the hot gas to enter the fine grained material (FGM) and heat it. Some pockets of FGM are difficult to heat because they are shielded from the hot gas by other FGM. Primitive objects in these pockets can survive capture. We computed the temperature of the hot gas at the shockwave. We started with the energy loss due to de- celeration. We assumed that the energy is partitioned equally between FeS and aerogel, so ≈20% of the silica and ≈30% of the FeS are volatalized, in accordance with the dissociation fractions predicted from DFT. The re- maining solids and gases are heated to higher temperature by the remaining available energy. The result is shown in Fig. 13. The initial gas temperature is about 14000 K at the start of the impact but falls rapidly. The gas temper- ature falls to < 1000 K by 2.5 mm into the track after a time of < 1 µs. We define this initial region as the heating region (see Fig. 13). Although this model makes a number of simplifying assumptions, the conclusion is clear: this particle was exposed to ≈ 104 K gas for < 1 µs. The inertia of the radially-accelerated material and the pressure of the shocked gas were largely responsible for blowing the bulbous cavity in the aerogel (Fig. 1). Figure 13. Vaporized gas temperature as a function of time. The top solid line shows simple heating of gases and serves as an upper limit for the temperature. The bottom dashed curve heats the gases only after accounting for the heat used to volatilize aerogel and sulfide, as well as to dissociate of FeS(g) → Fe(g) + S(g) and SiO2(g) → SiO(g) + O(g) or Si(g) + O2(g). The bottom curve serves as a lower limit for the heating. The shaded region shows the range of possible temperatures as a function of depth per the model. The heating zone marks the region where gas temperatures exceed 1000 K and lasts for < 1 µs and < 3 mm into the track. The travel time from the surface of the aerogel to the end of the hydrodynamic regime was ≈17 µs (Fig. 13). The heating of the sulfide and FGM in response to the < 1 µs thermal pulse depends on their respective thermal inertias. Given radius r, specific heat capacity cp, density ρ, and thermal conductivity k, the thermal time constant τ is τ ∼ ρcp k r2. (6) Using physical constants for sulfide from Tsatis and Theodossiou (1982), we find that the thermal timescale for Andromeda's sulfide was τ > 20 µs. The thermal time constant τ of a particle describes the time required for it 12 to heat up or cool down to 1/e ≈ 36% of the difference between its initial temperature and a new temperature. This was an order of magnitude longer than the duration of its exposure to 104 K gas and limits the internal heating it would have experienced. A 1 µs exposure to 104 K gas with a time constant of 20 µs leads to a maximum heating to ≈700 K, which assumes no heating losses from ablation. Within 1 µs after entering the aerogel, Andromeda was again in a cold environment, and frictional heating during stopping was negligible. There were undoubtedly ablative losses as the sulfide decelerated, and these would have removed heat from the sulfide so the interior may have remained at a few hundred degrees for approximately its thermal time constant, ≈20 µs. This surficial heating is reminiscent of entry of large meteorites by the Earth's atmosphere. High-Ni bands in sulfide: indicator of pre- accretional heating? The primary impactor was troilite with narrow Ni-rich bands of pyrrhotite 4C. In the case of the three bands averaged in Fig. 4, we computed the width by assuming stoichiometry. First we assumed (Fe+Ni)/S = 1 outside the Ni-rich bands, and then calculated (Fe+Ni)/S = 0.994 at the center of the bands. Pyrrhotite 4C has (Fe+Ni)/S = 0.875. Since the resolution of the EDS map was 30nm, the pyrrhotite bands should have been ≤ 2 nm or 1-2 unit cells thick (in the c* direction) and would have had a composition of Fe6NiS8 (Table 1, Sulfide06) to yield the observed composition. Such a thin band would have produced a broadening of the diffraction reflections of ≈ 1 nm−1. Based on SAED we found the peak broadening of the 4C superlattice reflections to be only 0.28 nm−1 which could be produced by bands on the order of 6-7 nm thick. Some regions of the sulfide had Ni content as high as 1 at % (i.e., double the Ni content in Fig. 4) and the HRTEM image shown in Fig. 3 shows a Ni band more than 20 nm thick. The likely explanation is that the sulfide contained Ni-rich pyrrhotite bands of varying thickness averaging around 7 nm. Gainsforth et al. (2017b) related the broadening of pyrrhotite 4C superlattice reflections with varying degrees of disequilibrium. If the sulfide is a nebular condensate (e.g. formed at 773 K), then the disorder would record the initial formation conditions. We follow the discussion of Gainsforth et al. (2017b), based on the experimental sulfidation of metal foils using H2S by Lauretta (2005). Andromeda's [Ni/Fe](cid:12) = -1.3 implies a large depletion of Ni in the source material or a very low partition of Ni into the sulfide phase. During sulfidation of Fe-metal by H2S at 773 K, Ni partitions primarily into the metal phase and forms taenite or can be pulled into schreibersite if P is present. Lower formation temperatures lower the partitioning of Ni into the sulfide but formation out of equilibrium favors placing more Ni into the sulfide. At 773 K, for pyrrhotites showing a similar level of disor- der, and for "nebular metal" with a few at % Ni, Cr, P, and C, we expect a partition between metal and sulfide DN i/F e metal/F eS = 1-2. Therefore, this sulfide did not form under these conditions because its Ni content is one order of magnitude too low. It could have formed under cooler temperatures but for a longer time so as to achieve a similar proximity to equilibrium. At 673 K, DN i/F e metal/F eS = 5. This would still put two to three times too much Ni into the sulfide. Alternately, it could have formed from a Ni-poor metal having only 0.1-0.5 % Ni originally. Another possibility is that Andromeda's sulfide is not a primary nebular condensate but rather formed in an igneous environment or experienced a metamorphic event. The fact that the Ni has partitioned into well defined Ni- rich bands suggests that it was homogeneously dissolved throughout the sulfide at an earlier time, but later condi- tions pushed it out of equilibrium. Such a condition could be obtained from a metamorphic event. To understand the thermal history we need to know what thermal profile could have exsolved the pyrrhotite. To our knowledge, the diffusion rate of Ni as a function of temperature in troilite has not been measured. Condit et al. (1974) measured the diffusion rates of Fe in troilite and pyrrhotite over the temperature range of 500 – 1000 K. We computed the timescales for diffusion of Ni by assum- ing it to have the same diffusivity as Fe. At 500 K, the segregation might occur on the time scale of hours and at 1000 K it could occur on the timescale of 1 second. We did not extrapolate to temperatures below 413 K because of a phase transition at that temperature. We previously calculated that the sulfide was heated to no more than several hundred degrees for a few µs during aerogel capture, so diffusion is far too slow to have allowed the Ni-rich bands to form as a result of capture in aerogel. We conclude that the bands are an original feature, either dating from initial formation or are due to low-temperature metamorphism. High-Ni lamellae from meteoritic sulfides with these characteristics have not previously been reported. In pre- vious analyses with similar sensitivity of four large Wild 2 sulfides, no such lamellae were observed. This provides further evidence against an aerogel capture origin, and also points to a diversity in formation/alteration history of sulfides within Wild 2. It is worth noting that the small variation in Ni content is very difficult to see instrumen- tally. The lack of reports in the literature may be due to the fact that TEM instruments capable of seeing this phenomenon easily have only recently become available. Our initial map of Andromeda missed the Ni-banding - it was only after we acquired a high count map with > 108 counts, for Se quantification, that the bands became obvious. As high-count EDS systems become more abun- dant in the community, more such banded sulfides may 13 be found. FGM Porosity, Density, Modal abundance Here we compare the physical characteristics of the FGM to FGM in CP-IDPs and to in situ cometary observations. We estimated the porosity (φ = fraction of filled space) of the FGM from Si and S EDS maps. We could not determine the modal abundance of carbonaceous material because of the ubiquitous epoxy. We found φ = 0.57 for the FGM. 88% of the material in the FGM was silicate and 12% of the material was sulfide, so we estimated the bulk density of the FGM to be ≈1.4 g/cm3. Gainsforth et al. (2017a) found porosities and densi- ties for several IDPs and found values in approximate agreement with Andromeda. Specifically, the porosity, density and silicate/sulfide abundance of the FGM is com- parable to the silicate/sulfide abundance in the "Nessie" IDP (L2071 Cluster 17). Joswiak et al. (2007) also re- ported densities of 0.7-1.7 g/cm3 for several TEM sections of IDPs, and Fraundorf et al. (1982) reported densities of 0.7-2.2 g/cm3 for IDPs measured on a quartz fiber balance. As noted in Gainsforth et al. (2017a), mea- surements from the Rosetta mission using the GIADA instrument found that dust grains ejected from Comet 67P/Churyumov-Gerasimenko (C-G) had densities be- tween 1-3 g/cm3(Rotundi et al. 2015). Using high-precision Doppler measurements, the Rosetta mission was able to measure the mass of 67P/Churyumov- Gerasimenko nucleus (Sierks et al. 2015). This yields a density of 0.5 g/cm3. The discrepancy between this bulk measurement of the comet nucleus and the higher densities seen in CP-IDPs and C-G particles implies the existence of void spaces on a larger length scale than the size of CP-IDPs or cometary particles collected by Stardust. FGM Heating As discussed earlier, vaporized aerogel and sulfide gas would have been in thermal contact with the leading edge of the particle. Unless the particle was rotating very rapidly (≈ 106 rotations s−1), Andromeda did not rotate significantly during the 1 µs hot thermal pulse, so one side was a leading edge and the opposite side a trailing edge. The FGM in the sulfide embayment was likely on the trailing edge of the sulfide during capture in aerogel. The hydrodynamic pressure (> 1 GPa on the leading edge) was probably much larger than the binding strength between the FGM particles, so it is unlikely that this loosely-bound material would have survived intact if it had been exposed directly to the flow of hot, vaporized gas. In the results section, we reported evidence of variable heating of individual silicate particles in the FGM. Here we explore two possible heating mechanisms, related to aerogel capture. The most likely mechanism was direct thermal contact with gas from the shocked, vaporized mixture of aerogel and ablated upstream cometary materials. Cooler gas in the cavitation region of the supersonic flow could have penetrated the porous FGM and been in thermal contact with it for the duration of the < 1 µs hot pulse (Fig. 12). Near the top of the track, thej SiO2 gas would have been partially dissociated into SiO and O, and the FeS gas would have been partially dissociated into Fe and S, so it would have been oxidizing and sulfidizing. The thermal time constant of ≈300 nm silicates is small, (cid:28) 0.1 µs, so any particles exposed to the gas would have rapidly equilibrated with it, but variations in permeability of the porous FGM could be expected to lead to variable exposures to the hot gas. We would expect some regions of the FGM to be inaccessible to the gas and any GEMS, EAs or other primary objects in these "shielded pockets" would be unheated or minimally heated, and remain unoxidized and unsulfidized. Without more information about the permeability and porosity of the FGM, it is probably not possible to model this process with sufficient fidelity to compare with detailed observations. We also considered radiative heat transfer from a hot, ionized gas, conservatively assuming that the radiating gas was an optically thick blackbody, and the FGM was also a blackbody with 100% absorptivity. We calculated that the heating of FGM was less than 1 K and rule it out as a significant heating mechanism. Excess silica is a common indicator of aerogel mixing in the bulb region of Stardust tracks (Leroux et al. 2008a; Stodolna et al. 2012). The bulk SiO2 component of the FGM is only 1 at% in excess of the SiO2 content of olivine – a relatively SiO2 poor mineral. This means that the amount of SiO2 contamination of the FGM is probably small – on the order of a percent. Since some gas phase SiO2 would have been produced during aerogel capture, this significantly limits the heating dose within the FGM and explains why some primitive components survived. We computed the heating that would be experienced by 200 K quartz (as a proxy for FGM) from thermal equilibration with an SiO2 gas at 104 K. Assuming the gas/solid ratio is 1%, we found that the FGM should heat to ≈300 K. Apparently, some FGM heated more than that. One possibility is that more gas was present and heated the sample, but later diffused out without condensing. If the abundance of gas was an order of magnitude greater, then the FGM could have heated to about 1100 K which is high enough to begin to see alteration – though only the smallest objects (e.g, GEMS) would alter on the timescale of 1 µs. A more likely explanation is that the hot gas heated a smaller fraction of the material because it was only in contact with the pores. Indeed, if we assume that the gas was only in contact with about 10% of the FGM, then it should have heated about 10% of the FGM to 1100 K and then deposited a total mass of 1% Si when it 14 condensed. Brownlee et al. (2005) found that heating GEMS to temperatures ≥ 1000 K for several hours results in sub- solidus crystallization of the silicate and they postulated that equilibrated aggregates could have formed by heating in the nebular environment. If this picture is correct, then equilibrated aggregates should be more robust to transient heating events than GEMS – i.e., GEMS should be more susceptible to destruction during aerogel capture than equilibrated aggregates. With this in mind, the survival of a GEMS would provide the strongest constraint on the temperatures during heating. Primitive FGM FGM10 is the best candidate for a GEMS in Andromeda (Fig. 5). It contains euhedral sulfides, does not contain excess silica, and retains the expected chemical compo- sition of a GEMS. While there is likely some heating on one side of FGM10 where there is one subhedral exter- nal sulfide, and sulfur has bled into the silicate matrix, the appearance of FGM10 is similar to GEMS seen in IDPs, many of which also show evidence of heating from atmospheric entry or nebular events (Brownlee et al. 1993; Dai and Bradley 2005). FGM10 is indistinguishable in its characteristics with GEMS in CP-IDPs. For example, see Fig. 5 in Keller and Messenger (2005) or Fig. 1 in Keller and Messenger (2011). FGM11 and FGM12 appear to be indistingishable from equilibrated aggregates in CP-IDPs, with no indicators of shock heating from the aerogel capture. Equilibrated ag- gregates are probably more robust to heating than GEMS and so it is possible that they were heated but not modified by the process. For comparison to EAs in IDPs, see Fig. 2 in Rietmeijer (2009), Fig. 2 in Keller and Messenger (2005) or Fig. 1 in Keller and Messenger (2009). Altered FGM We argue that the remaining nanophases studied in An- dromeda's FGM were thermally altered to a greater or lesser degree on account of the presence of one or more of the heating indicators we described in the previous sections. FGM20 was especially noteworthy because EDS imag- ing could clearly distinguish euhedral grains of distinct compositions but brightfield imaging and diffraction in- dicated the grains were amorphous. We interpret this as similar to "shadow grains" described previously by Stodolna et al. (2012) and Leroux et al. (2008a) when a crystalline phase is amorphized by rapid heating and cooling but not sufficiently heated for the atoms to diffuse away. The atoms then settle into metastable amorphous phases with shapes and compositions very close to their original crystalline counterparts. FGM20 was in contact with the primary impactor sulfide in a region containing many Fe-oxides, and did not show excess silica. Because it was found in a central part of the FGM, it would have been unlikely that it came into contact with hot, molten aerogel. Instead its morphology was consistent with the hot gas model we propose. If hot gas permeated the FGM, then it could have penetrated all open pores and come in contact with the sulfide where it would have rapidly cooled. Amorphous FeOx The amorphous Fe-oxides found near FGM16 (Fig. 7), FGM17 and FGM20 may be important to understanding the origin of the FGM. The oxide is embedded in epoxy so it is not possible to determine the exact oxygen abundance even after subtracting off a background epoxy spectrum. However, based on rough stoichiometry we expect a mini- mum oxidation of Fe2+, and in some cases more oxygen is present than would be expected for Fe3+ so even oxy- hydroxide is consistent with our error margin. Other rock forming elements are present in lower abundance. The Fe-oxide is closest in composition to magnetite rims in some mildly heated IDPs. In such IDPs it is pos- sible to find magnetite rims on the order of 100 nm which contain nanocrystalline and amorphous oxidized Fe. These often contain small abundances of other elements deriving primarily from the neighboring sulfides and silicates from which they formed. The oxides in Andromeda all contain significant abundances of other elements, but Si and S are present in every spectrum and furthermore are present in roughly equal amounts of a few At%. Experiments by Nuth et al. (2000) and Rietmeijer (2000) have shown that it is possible to synthesize Fe oxide amorphous nanocrystals directly from vapor phase in an ≈ 90 Torr H2-rich gas (0.12 bar). They were formed at 500 - 1500 K and on time scales "much less than a second." With the exception of the H2 atmosphere and the low pressures, these conditions are consistent with what we should expect from shock heating during aerogel capture. Fig. 9 shows nanocrystalline magnetite in an IDP alongside artificial smokes produced by Nuth et al. (2000). The inset shows an Fe-oxide from Andromeda for compar- ison (also cf. Fig. 7). Barth et al. (2017) show that at lower temperatures, e.g. 450 ◦C, Fe-oxides can form from a high fO2 and high fS2 gas. Specifically under high fugacity, low temperature conditions, oxides are stable unless log(fO2) < log(fS2) by about 20 log units. Our calculations above show that log(fO2) ≈ log(fS2) which means that hematite would have been the stable phase if aerogel + sulfide gas were condensing. Because the process would have been so rapid, the Fe-oxide would not have crystallized but instead solidified as an amorphous solid. 15 The prevalence of Fe-oxide at the interface between the large sulfide impactor and the FGM would have been expected. Because the sulfide did not have time to heat up more than a few hundred degrees during the capture process, it should have acted as a heat sink for nearby material. As hot gas permated through the FGM and came into contact with the sulfide, it would have cooled to a few hundred degrees which would have driven rapid condensation. Hot gas further out in the FGM would have diffused out of the FGM at a later point and would likely leave less residue. Wild 2 as an Aggregate Rock Immediately after the Stardust return, Brownlee et al. (2006) proposed that comet Wild 2 is, effectively, an ag- gregate of inner and outer Solar System materials which indicated substantial mixing throughout the solar nebula. The amount of presolar material present in Wild 2 is con- sistent with CP-IDPs, and other primitive bodies (Floss et al. 2013), so Wild 2 is certainly sampling some unal- tered solar nebula material. Work on large rocky objects (terminal particles) in Wild 2 has frequently identified chondrules, and refractory rocks indicating that inner So- lar System materials composed a significant fraction of the captured mass of Wild 2 (Joswiak et al. 2012; Nakamura et al. 2008; Brownlee 2013; Gainsforth et al. 2015a). Re- cently, Joswiak et al. (2017) found that CAI-like fragments account for ≈ 1 vol% in cometary samples including Wild 2 and a giant cluster IDP named U2-20. Joswiak et al. (2017) make a strong case that U2-20 has a cometary origin, and have noted (personal communication) that it has coarse grained rocks (CAI-like) in direct contact with FGM including GEMS and EAs. Wooden et al. (2017) recently tied together several lines of evidence pointing to late formation of Wild 2 materials, several Mya the onset of CAI formation. These include an abundance of objects with O isotopic signatures close to the terres- trial value, and an overabundance of late stage objects including Fe-rich type II chondrules, and late 26Al dates. Wild 2 olivines also show some evidence of having been metamorphosed. (Frank et al. 2014). Nevertheless, while Wild 2 FGM has occasionally been found near large terminal particles, the FGM has usually been heavily altered and as a consequence, comparison with FGM in CP-IDPs has been difficult. For example, the only enstatite whisker found to date in Wild 2 is encased in glass – a phonomenon never reported previ- ously (Stodolna et al. 2014). Febo has had to date the most promising FGM, but significant unresolved questions about the alteration of the FGM remain (Joswiak et al. 2012). A long-standing open question has therefore been whether Wild 2 has any "CP-IDP" FGM at all. The fortu- itous survival of the Andromeda FGM indicates that the answer is "yes". The presence of sulfides in the FGM with a signifi- cantly different composition than the impactor indicate that they are not simply ablation products from the im- pactor during capture. This means that the FGM does not share the same formation/alteration history that An- dromeda's large sulfide experienced. If the sulfide had been metamorphosed while aggregated to the FGM, the FGM would not have survived, so either the sulfide formed originally with Ni-rich bands, or aggregated with the FGM after metamorphosis/nebular heating. This observation supports the conclusion that Wild 2 is an aggregate rock with GEMS and equilibrated aggregates alongside larger rocks/minerals. K and P are volatile elements and are present in nearly all the objects within the FGM material that we studied, and may provide another window into the relationship between the FGM and the coarse material. Because of their volatility, the gas heating model would predict that K and P would disperse throughout the FGM with the gas and then recondense as the gas cooled. However, K is significantly enriched in the bulk composition with [K/Si](cid:12) = 0.32. Simple redistribution of K would predict that the bulk would remain unchanged or even depleted as K escaped into the nearby aerogel. High K has been noted previously within Stardust samples (Flynn et al. 2006), though there is some evidence for K as an occasional contaminant in aerogel alongside Cl and Ca (Rietmeijer 2015). K is especially intriguing as it is commonly enriched in more evolved materials. The carrier phase for the K in Andromeda has not been identified and may have been destroyed during capture. This suggests that an investigation of the abundance and carrier of K in CP- IDPs may yield interesting results. P, on the other hand, is not particularly enriched in the bulk FGM ([P/Si](cid:12) = 0.049 ≈ 0), but is significantly overabundant around the nanophases we investigated (typ- ical [P/Si](cid:12) = 0.15 with the highest value at FGM17 with [P/Si](cid:12))=0.55). The nanophases we investigated were preferentially those which looked like GEMS or EAs or other primitive material (i.e. not large crystals, sulfides, etc.) Therefore, while FGM21 shows that P was likely re- distributed to some degree during capture, it must also be more concentrated within the GEMS/EA-type objects or else it would have also been more concentrated in the bulk. The reason for this is not clear but warrants investigation in the future. CONCLUSIONS We summarize key conclusions: 1. Wild 2 contains objects that are indistinguishable from GEMS and equilibrated aggregates found in CP-IDPs. 2. Wild 2 is an aggregate containing both large crystals and FGM. 16 3. Ni-banding in the sulfide impactor may show evi- dence for heating between formation and incorpora- tion into comet Wild 2. 4. The distribution of K and P leads us to suggest that future research would benefit by identifying the origin and chemistry of K- and P- bearing phases in IDPs. We have also constrained the processes active during Andromeda's capture, which allows us to better resolve which processes can be attributed to cometary/Solar Sys- tem processes and which are artifacts of capture. We demonstrate that objects within the fine grained mate- rial (FGM) behind the sulfide impactor were heated by a transient hot gas on the order of ≈ 104 K but for 1 µs at most. This is sufficient to melt individual silicates but some objects appear to have survived in a relatively pristine condition. This is likely due to variations in the porosity and the thermal inertia of nearby objects (e.g. the impactor and large crystalline grains) which shielded some FGM from exposure to the hot gas. The FGM only has excess aerogel in select regions. We have identified the gas-phase condensation of Fe-oxide from volatilized Fe and O produced during aerogel capture. ACKNOWLEDGMENTS Work at the Molecular Foundry and Advanced Light Source was supported by the Office of Science, Office of Basic Energy Sciences, of the U.S. Department of Energy under Contract No. DE-AC02-05CH11231. This work was supported under the LARS program by NASA grant NNX16AK14G. MZ was supported by NASA's Emerging Worlds Program. We wish to thank the staff at Johnson Space Center for making IDPs and Stardust samples available to the community. We wish to thank Karen Bustillo and the rest of the staff at the National Center for Electron Microscopy for providing spectacular instrumentation without which this paper would not exist. We wish to thank Yufeng Liang and David Prendergast at the Molecular Foundry for tutelage with DFT and access to computing resources. We wish to thank Natasha Johnson and Joseph A. Nuth III for supplying the Fe smokes used for comparison with Fe-oxides. REFERENCES Arthur, D. and Vassilvitskii, S. 2007. k-means++: The ad- vantages of careful seeding. Proceedings of the eighteenth annual ACM-SIAM symposium on discrete algorithms pp. 1027–1035. Barth, M. I. F., Harries, D., Langenhorst, F., and Hoppe, P. 2017. Sulfide-oxide assemblages in Acfer 094-Clues to nebular metal-gas interactions. Meteoritics and Plan- etary Science 53:187–203. Black, D. and Long, G. 2004. X-Ray Topography, SP 960- 10, Washington DC: US Government Printing Office. Brownlee, D. 2013. The Stardust Mission: Analyzing Samples from the Edge of the Solar System. Annu. Rev. Earth Planet. Sci. 42:140205180347002. Brownlee, D., Tsou, P., Aléon, J., Alexander, C. M. O., Araki, T., Bajt, S., Baratta, G. A., Bastien, R. K., Bland, P., Bleuet, P., Borg, J., Bradley, J. P., Brearley, A., Brenker, F., Brennan, S., Bridges, J. C., Brown- ing, N. D., Brucato, J. R., Bullock, E., Burchell, M. J., Busemann, H., Butterworth, A. L., Chaussidon, M., Cheuvront, A., Chi, M., Cintala, M. J., Clark, B. C., Clemett, S. J., Cody, G., Colangeli, L., Cooper, G., Cordier, P., Daghlian, C., Dai, Z. R., D'Hendecourt, L., Djouadi, Z., Dominguez, G., Duxbury, T., Dworkin, J. P., Ebel, D. S., Economou, T. E., Fakra, S., Fairey, S. A. J., Fallon, S., Ferrini, G., Ferroir, T., Fleckenstein, H., Floss, C., Flynn, G. J., Franchi, I. A., Fries, M., Gainsforth, Z., Gallien, J.-P., Genge, M., Gilles, M. K., Gillet, P., Gilmour, J., Glavin, D. P., Gounelle, M., Grady, M. M., Graham, G. A., Grant, P. G., Green, S. F., Grossemy, F., Grossman, L., Grossman, J. N., Guan, Y., Hagiya, K., Harvey, R., Heck, P., Herzog, G. F., Hoppe, P., Hoerz, F., Huth, J., Hutcheon, I. D., Ignatyev, K., Ishii, H., Ito, M., Jacob, D., Jacobsen, C., Jacobsen, S., Jones, S., Joswiak, D., Jurewicz, A., Kearsley, A. T., Keller, L. P., Khodja, H., Kilcoyne, A. L. D., Kissel, J., Krot, A., Langenhorst, F., Lanzirotti, A., Le, L., Leshin, L. A., Leitner, J., Lemelle, L., Leroux, H., Liu, M.-C., Luening, K., Lyon, I., MacPherson, G., Marcus, M. A., Marhas, K., Marty, B., Matrajt, G., Mc- Keegan, K., Meibom, A., Mennella, V., Messenger, K., Messenger, S., Mikouchi, T., Mostefaoui, S., Nakamura, T., Nakano, T., Newville, M., Nittler, L. R., Ohnishi, I., Ohsumi, K., Okudaira, K., Papanastassiou, D. A., Palma, R., Palumbo, M. E., Pepin, R. O., Perkins, D., Perronnet, M., Pianetta, P., Rao, W., Rietmeijer, F. J. M., Robert, F., Rost, D., Rotundi, A., Ryan, R., Sandford, S. A., Schwandt, C. S., See, T. H., Schlut- ter, D., Sheffield-Parker, J., Simionovici, A. S., Simon, S., Sitnitsky, I., Snead, C. J., Spencer, M. K., Stader- mann, F. J., Steele, A., Stephan, T., Stroud, R., Susini, J., Sutton, S. R., Suzuki, Y., Taheri, M., Taylor, S., Teslich, N., Tomeoka, K., Tomioka, N., Toppani, A., Trigo-Rodríguez, J. M., Troadec, D., Tsuchiyama, A., Tuzzolino, A. J., Tyliszczak, T., Uesugi, K., Velbel, M., Vellenga, J., Vicenzi, E., Vincze, L., Warren, J., Weber, I., Weisberg, M., Westphal, A. J., Wirick, S., Wooden, D., Wopenka, B., Wozniakiewicz, P., Wright, I., Yabuta, H., Yano, H., Young, E. D., Zare, R. N., Zega, T., Ziegler, K., Zimmerman, L., Zinner, E., and Zolensky, 17 M. E. 2006. Comet 81P/Wild 2 Under a Microscope. Science 314:1711–1716. Brownlee, D. E., Joswiak, D. J., Bradley, J. P., Matrajt, G., and Wooden, D. H. 2005. Cooked GEMS - Insights into the Hot Origins of Crystalline Silicates in Circumstellar Disks and the Cold Origins of GEMS. Brownlee, D. E., Joswiak, D. J., Love, S. G., Nier, A. O., Schlutter, D. J., and Bradley, J. P. 1993. Identification of cometary and asteroidal particles in stratospheric IDP collections. Lunar and Planetary Sciences Conference 24. Burchell, M., Fairey, S., Wozniakiewicz, P., Brownlee, D. E., Hoerz, F., Kearsley, A. T., See, T. H., Tsou, P., Westphal, A. J., Green, S., Trigo-Rodriguez, J. M., and Dominguez, G. 2008. Characteristics of cometary dust tracks in Stardust aerogel and laboratory calibrations. Meteoritics and Planetary Science 43:23–40. Chase, M. W. J. 1998. NIST-JANAF Thermochemical Ta- bles Fourth Edition. Journal of Physical and Chemical Reference Data pp. 1–61. Cliff, G. and Lorimer, G. W. 1975. The Quantitative Analysis of Thin Specimens. Journal of Microscopy 103:203–207. Condit, R. H., Hobbins, R. R., and Birchenall, C. E. 1974. Self-diffusion of iron and sulfur in ferrous sulfide. Oxidation of Metals 8:409–455. Corso, A. D. 2014. Pseudopotentials periodic table: From H to Pu. Computational Materials Science 95:337–350. Dai, Z. R. and Bradley, J. P. 2005. Origin and properties of GEMS (Glass with Embedded Metal and Sulfides). Chondrites and the Protoplanetary Disk 341:668–674. Dominguez, G. 2009. Time evolution and temperatures of hypervelocity impact-generated tracks in aerogel. Mete- oritics and Planetary Science 44:1431–1443. Floss, C., Stadermann, F. J., Kearsley, A. T., Burchell, M. J., and Ong, W. J. 2013. The Abundance of Presolar Grains in Comet 81P/Wild 2. The Astrophysical Journal 763:140–11. Flynn, G. J., Bleuet, P., Borg, J., Bradley, J. P., Brenker, F. E., Brennan, S., Bridges, J., Brownlee, D. E., Bul- lock, E. S., Burghammer, M., Clark, B. C., Dai, Z. R., Daghlian, C. P., Djouadi, Z., Fakra, S., Ferroir, T., Floss, C., Franchi, I. A., Gainsforth, Z., Gallien, J.-P., Gillet, P., Grant, P. G., Graham, G. A., Green, S. F., Grossemy, F., Heck, P. R., Herzog, G. F., Hoppe, P., Hoerz, F., Huth, J., Ignatyev, K., Ishii, H. A., Janssens, K., Joswiak, D., Kearsley, A. T., Khodja, H., Lanzirotti, A., Leitner, J., Lemelle, L., Leroux, H., Luening, K., MacPherson, G. J., Marhas, K. K., Marcus, M. A., Matrajt, G., Nakamura, T., Nakamura-Messenger, K., Nakano, T., Newville, M., Papanastassiou, D. A., Pi- anetta, P., Rao, W., Riekel, C., Rietmeijer, F. J. M., Rost, D., Schwandt, C. S., See, T. H., Sheffield-Parker, J., Simionovici, A. S., Sitnitsky, I., Snead, C. J., Sta- dermann, F. J., Stephan, T., Stroud, R. M., Susini, J., Suzuki, Y., Sutton, S. R., Taylor, S., Teslich, N., Troadec, D., Tsou, P., Tsuchiyama, A., Uesugi, K., Veke- mans, B., Vicenzi, E. P., Vincze, L., Westphal, A. J., Wozniakiewicz, P., Zinner, E., and Zolensky, M. E. 2006. Elemental Compositions of Comet 81P/Wild 2 Samples Collected by Stardust. Science 314:1731–1735. Frank, D. R., Zolensky, M. E., and Le, L. 2014. Olivine in terminal particles of Stardust aerogel tracks and analogous grains in chondrite matrix. Geochimica et Cosmochimica Acta 142:240–259. Fraundorf, P., Hintz, C., Lowry, O., Mckeegan, K. D., and Sandford, S. A. 1982. Determination of the Mass, Surface Density, and Volume Density of Individual In- terplanetary Dust Particles. Workshop on Dust in Plan- etary Systems pp. 225–226. Gainsforth, Z. 2016. Stoichiometry Fitter, a GUI for Fitting Solid Solutions and Analyzing Mineral Phases. Microscopy and Microanalysis 22:1808–1809. Gainsforth, Z., Brenker, F. E., Simionovici, A. S., Schmitz, S., Burghammer, M., Butterworth, A. L., Cloetens, P., Lemelle, L., Tresserras, J.-A. S., Schoonjans, T., Sil- versmit, G., Solé, V. A., Vekemans, B., Vincze, L., Westphal, A. J., Allen, C., Anderson, D., Ansari, A., Bajt, S., Bastien, R. K., Bassim, N., Bechtel, H. A., Borg, J., Bridges, J., Brownlee, D. E., Burchell, M., Changela, H., Davis, A. M., Doll, R., Floss, C., Flynn, G., Fougeray, P., Frank, D. R., Grün, E., Heck, P. R., Hillier, J. K., Hoppe, P., Hudson, B., Huth, J., Hvide, B., Kearsley, A., King, A. J., Lai, B., Leitner, J., Leroux, H., Leonard, A., Lettieri, R., Marchant, W., Nittler, L. R., Ogliore, R., Ong, W. J., Postberg, F., Price, M. C., Sandford, S. A., Srama, R., Stephan, T., Sterken, V. J., Stodolna, J., Stroud, R. M., Sutton, S., Trieloff, M., Tsou, P., Tsuchiyama, A., Tyliszczak, T., Von Korff, J., Zevin, D., Zolensky, M. E., and >30,000 Stardust@home dusters 2013a. Stardust Interstellar Preliminary Exam- ination VIII: Identification of crystalline material in two interstellar candidates. Meteoritics and Planetary Science 49:1645–1665. Gainsforth, Z., Bustillo, K., Butterworth, A. L., Ogliore, R. C., and Westphal, A. J. 2014a. Trace Element Analy- sis in Geochemical Systems by STEM/EDS. Microscopy and Microanalysis 20:1682–1683. Gainsforth, Z., Butterworth, A. L., Stodolna, J., Westphal, A. J., Huss, G. R., Nagashima, K., Ogliore, R., Brown- 18 lee, D. E., Joswiak, D., Tyliszczak, T., and Simionovici, A. S. 2015a. Constraints on the formation environ- ment of two chondrule-like igneous particles from comet 81P/Wild 2. Meteoritics and Planetary Science 50:976– 1004. Gainsforth, Z., Butterworth, A. L., and Westphal, A. J. 2015b. Unequilibrated Spinels in Stardust Track C2062,2,162 (Cecil). Meteoritics and Planetary Sci- ence 46:2974. Gainsforth, Z., Jilly-Rehak, C. E., Butterworth, A. L., and Westphal, A. J. 2017a. Petrography of Four CP-IDPs. Lunar and Planetary Sciences Conference p. 1642. Gainsforth, Z., Lauretta, D. S., Tamura, N., Westphal, A. J., Jilly-Rehak, C. E., and Butterworth, A. L. 2017b. Insights into solar nebula formation of pyrrhotite from nanoscale disequilibrium phases produced by H2S sul- fidation of Fe metal. American Mineralogist 102:1881– 1893. Gainsforth, Z., McLeod, A. S., Butterworth, A. L., Dominguez, G., Basov, D., Keilmann, F., Thiemens, M., Tyliszczak, T., and Westphal, A. J. 2013b. Caligula, a Stardust sulfide-silicate assemblage viewed through SEM, nanoFTIR, and STXM. 45th Lunar and Planetary Science Conference 44:2332. Gainsforth, Z., Ogliore, R. C., Bustillo, K., Westphal, A. J., and Butterworth, A. L. 2014b. Ni Zoned Nano- Pyrrhotite from Stardust Track C2062,2,152 (Cecil). 45th Lunar and Planetary Science Conference p. 2637. Giannozzi, P., Baroni, S., Bonini, N., Calandra, M., Car, R., Cavazzoni, C., Ceresoli, D., Chiarotti, G. L., Co- coccioni, M., Dabo, I., Dal Corso, A., de Gironcoli, S., Fabris, S., Fratesi, G., Gebauer, R., Gerstmann, U., Gougoussis, C., Kokalj, A., Lazzeri, M., Martin-Samos, L., Marzari, N., Mauri, F., Mazzarello, R., Paolini, S., Pasquarello, A., Paulatto, L., Sbraccia, C., Scan- dolo, S., Sclauzero, G., Seitsonen, A. P., Smogunov, A., Umari, P., and Wentzcovitch, R. M. 2009. QUAN- TUM ESPRESSO: a modular and open-source software project for quantum simulations of materials. Journal of Physics: Condensed Matter 21:395502–20. Horz, F., Bastien, R. K., Borg, J., Bradley, J. P., Bridges, J. C., and et al 2006. Impact Features on Stardust: Implications for Comet 81P/Wild 2 Dust. Science p. 10. Ishii, H. A., Bradley, J. P., Dai, Z. R., Chi, M., Kearsley, A. T., Burchell, M. J., Browning, N. D., and Molster, F. 2008. Comparison of Comet 81P/Wild 2 Dust with Interplanetary Dust from Comets. Science 319:447. Jain, A., Ong, S. P., Hautier, G., Chen, W., Richards, W. D., Dacek, S., Cholia, S., Gunter, D., Skinner, D., Ceder, G., and Persson, K. A. 2013. Commentary: 19 The Materials Project: A materials genome approach to accelerating materials innovation. APL Materials 1:011002–12. Joswiak, D. J., Brownlee, D. E., and Bradley, J. P. 1996. Systematic analyses of major element distributions in GEMS from high speed IDPs. 40th Lunar and Planetary Sciences Conference . Joswiak, D. J., Brownlee, D. E., Matrajt, G., Westphal, A. J., and Snead, C. J. 2009. Kosmochloric Ca-rich pyroxenes and FeO-rich olivines (Kool grains) and asso- ciated phases in Stardust tracks and chondritic porous interplanetary dust particles: Possible precursors to FeO-rich type II chondrules in ordinary chondrites. Me- teoritics and Planetary Science 44:1561–1588. Joswiak, D. J., Brownlee, D. E., Matrajt, G., Westphal, A. J., Snead, C. J., and Gainsforth, Z. 2012. Comprehen- sive examination of large mineral and rock fragments in Stardust tracks: Mineralogy, analogous extraterrestrial materials, and source regions. Meteoritics and Planetary Science 47:471. Joswiak, D. J., Brownlee, D. E., Nguyen, A. N., and Messenger, S. 2017. Refractory materials in comet samples. Meteoritics and Planetary Science 52:1612– 1648. Joswiak, D. J., Brownlee, D. E., Pepin, R. O., and Schlut- ter, D. J. 2007. Densities and Mineralogy of Cometary and Asteroidal Interplanetary Dust Particles Collected in the Stratosphere. Workshop on Dust in Planetary Systems pp. 141–144. Keller, L. P. and Messenger, S. 2005. The Nature and Origin of Interplanetary Dust: High-Temperature Com- ponents, pp. 657–667. In A. N. Krot, E. R. D. Scott, and B. Reipurth (eds.), Chondrites and the Protoplanetary Disk. Chondrites and the Protoplanetary Disk. Keller, L. P. and Messenger, S. 2009. Equilibrated aggre- gates in cometary IDPs: insights into the crystallization process in protoplanetary disks. Lunar and Planetary Science Conference XXVII XXVII:2121. Keller, L. P. and Messenger, S. 2011. On the origins of GEMS grains. Geochimica et Cosmochimica Acta 75:5336–5365. Lauretta, D. Sulfidation iron–nickel–chromium–cobalt–phosphorus in 1400–1000◦C. Oxidation of Metals 64:1–22. S. 2005. of an alloy Lejaeghere, K., Bihlmayer, G., Bjorkman, T., Blaha, P., Blugel, S., Blum, V., Caliste, D., Castelli, I. E., Clark, S. J., Dal Corso, A., de Gironcoli, S., Deutsch, T., De- whurst, J. K., Di Marco, I., Draxl, C., Du ak, M., Eriks- son, O., Flores-Livas, J. A., Garrity, K. F., Genovese, L., Giannozzi, P., Giantomassi, M., Goedecker, S., Gonze, X., Granas, O., Gross, E. K. U., Gulans, A., Gygi, F., Hamann, D. R., Hasnip, P. J., Holzwarth, N. A. W., Iu an, D., Jochym, D. B., Jollet, F., Jones, D., Kresse, G., Koepernik, K., Kucukbenli, E., Kvashnin, Y. O., Locht, I. L. M., Lubeck, S., Marsman, M., Marzari, N., Nitzsche, U., Nordstrom, L., Ozaki, T., Paulatto, L., Pickard, C. J., Poelmans, W., Probert, M. I. J., Refson, K., Richter, M., Rignanese, G. M., Saha, S., Scheffler, M., Schlipf, M., Schwarz, K., Sharma, S., Tavazza, F., Thunstrom, P., Tkatchenko, A., Torrent, M., Vander- bilt, D., van Setten, M. J., Van Speybroeck, V., Wills, J. M., Yates, J. R., Zhang, G. X., and Cottenier, S. 2016. Reproducibility in density functional theory calculations of solids. Science 351:aad3000–aad3000. Leroux, H., Rietmeijer, F. J. M., Velbel, M. A., Brearley, A. J., Jacob, D., Langenhorst, F., Bridges, J. C., Zega, T. J., Stroud, R. M., Cordier, P., Harvey, R. P., Lee, M., Gounelle, M., and Zolensky, M. E. 2008a. A TEM study of thermally modified comet 81P/Wild 2 dust particles by interactions with the aerogel matrix during the Stardust capture process. Meteoritics and Planetary Science 43:97–120. Leroux, H., Stroud, R. M., Dai, Z. R., Graham, G. A., Troadec, D., Bradley, J. P., Teslich, N., Borg, J., Kears- ley, A. T., and Hörz, F. 2008b. Transmission electron mi- croscopy of cometary residues from micron-sized craters in the Stardust Al foils. Meteoritics and Planetary Science 43:143–160. Lodders, K. 2003. Solar System Abundances and Conden- sation Temperatures of the Elements. The Astrophysical Journal 591:1220–1247. Love, S. G. and Brownlee, D. E. 1991. Heating and thermal transformation of micrometeoroids entering the Earth's atmosphere. Icarus . Matrajt, G., Ito, M., Wirick, S., Messenger, S., Brown- lee, D. E., Joswiak, D., Flynn, G., Sandford, S. A., Snead, C., and Westphal, A. J. 2008. Carbon investi- gation of two Stardust particles: A TEM, NanoSIMS, and XANES study. Meteoritics and Planetary Science 43:315–334. Messenger, S., Nakamura-Messenger, K., Keller, L. P., and Clemett, S. J. 2015. Pristine stratospheric collection of interplanetary dust on an oil-free polyurethane foam substrate. Meteoritics and Planetary Science 50:1468– 1485. Nakamura, T., Noguchi, T., Tsuchiyama, A., Ushikubo, T., Kita, N. T., Valley, J. W., Zolensky, M. E., Kakazu, Y., Sakamoto, K., Mashio, E., Uesugi, K., and Nakano, T. 2008. Chondrulelike Objects in Short-Period Comet 81P/Wild 2. Science 321:1664–1667. Nuth, J. A., Hallenbeck, S. L., and Rietmeijer, F. J. M. 2000. Laboratory studies of silicate smokes: Analog studies of circumstellar materials. Journal of Geophysi- cal Research 105:10387–10396. Perdew, J. P., Burke, K., and Ernzerhof, M. 1996. Gener- alized Gradient Approximation Made Simple. Physical Review Letters 77:3865–3868. Prandini, G., Marrazzo, A., Castelli, I., Mounet, N., and Marzari, N. 2018. SSSP. p. In preparation. Rietmeijer, F. 2000. Metastable eutectic equilibrium brought down to Earth. Wiley Online Library 81:409– 420. Rietmeijer, F. 2009. A cometary aggregate interplanetary dust particle as an analog for comet Wild 2 grain chem- istry preserved in silica-rich Stardust glass. Meteoritics and Planetary Science 44:1589–1609. Rietmeijer, F. J. M. 2015. The smallest comet 81P/Wild 2 dust dances around the CI composition. Meteoritics and Planetary Science 50:1767–1789. Rotundi, A., Sierks, H., Della Corte, V., and Fulle, M. 2015. Dust measurements in the coma of comet 67P/Churyumov-Gerasimenko inbound to the Sun. Sci- ence 347:aaa3905–1–6. Shu, F. H. 1992. The Physics of Astrophysics: Gas dy- namics. University Science Books, Sausalito, California. Sierks, H., Barbieri, C., Lamy, P. L., and Rodrigo, R. 2015. On the nucleus structure and activity of comet 67P/Churyumov-Gerasimenko. Science 347. Stodolna, J., Gainsforth, Z., Butterworth, A. L., and Westphal, A. J. 2014. Characterization of preserved primitive fine-grained material from the Jupiter family comet 81P/Wild 2 – A new link between comets and CP-IDPs. Earth and Planetary Science Letters 388:367– 373. Stodolna, J., Jacob, D., and Leroux, H. 2012. Mineralogy and petrology of Stardust particles encased in the bulb of track 80: TEM investigation of the Wild 2 fine- grained material. Geochimica et Cosmochimica Acta 87:35–50. Tamura, N. 2014. XMAS: A versatile tool for analyzing synchrotron X-ray microdiffraction data , pp. 125–155. In R. Barabash and G. Ice (eds.), Strain and Dislocaton Gradients from Diffraction Spatially-Resolved Local Structure and Defects. ., London. Tamura, N., Kunz, M., Chen, K., Celestre, R. S., Mac- Dowell, A. A., and Warwick, T. 2009. A superbend X-ray microdiffraction beamline at the advanced light source. Materials Science and Engineering A 524:28–32. 20 Tsatis, D. and Theodossiou, A. 1982. Thermal diffusivity in pyrrhotite (Fe7S8). J. Phys. Chem. Solids 43:771–772. Vanderbilt, D. 1990. Soft self-consistent pseudopotentials in a generalized eigenvalue formalism. Physical Review B 41:7892–7895. Westphal, A. J., Snead, C., Butterworth, A. L., Graham, G. A., Bradley, J. P., Bajt, S., Grant, P. G., Bench, G., Brennan, S., and Pianetta, P. 2004. Aerogel keystones: Extraction of complete hypervelocity impact events from aerogel collectors. Meteoritics and Planetary Science 39:1375–1386. Wooden, D. H., Ishii, H. A., and Zolensky, M. E. 2017. Cometary dust: the diversity of primitive refractory grains. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences 375:20160260–58. 21 Table 6. Nanophase compositions of capture modified objects from TEM EDS Normalized to Mg and CI. Atomic % / Protosolar / (Mg or Si)1 O 0.22 0.41 0.34 0.26 0.74 0.33 0.37 0.49 0.42 0.72 1.17 0.50 1.64 Na 0.27 0.20 0.53 0.34 0.18 0.96 b.d. 0.74 0.31 0.76 0.69 0.77 0.50 Mg 0.08 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00 Ca 0.52 0.73 0.57 0.21 1.41 1.34 0.76 2.34 1.30 4.43 0.97 1.60 2.36 # FGM09 FGM10 FGM11 FGM12 FGM13 FGM14 FGM15 FGM16 FGM17 FGM18 FGM19 FGM20 FGM21 1 The FGM bulk is the only spectrum that has been normalized to Si. The FGM as a whole is subchondritic in Mg, but approximately chondritic in Si. See text. 2 χ2 3 p is the probability that this measurement is inconsistent with the distribution of GEMS compositions. 7 is the reduced χ2 with 7 degrees of freedom, showing the difference between this spectrum and an ideal GEMS composition. See text. K 2.08 4.06 3.83 2.70 5.84 0.97 1.45 2.88 3.17 5.92 2.54 5.94 39.82 (2) χ2 7 2.44 0.55 1.31 17.78 3.02 1.42 0.84 0.92 3.12 1.95 1.58 1.02 2.52 Ti b.d. 1.24 0.79 0.73 1.59 0.29 1.40 0.71 0.46 2.15 b.d. 1.07 8.76 Cr 0.16 1.05 0.92 5.30 1.46 1.35 1.32 1.43 3.12 1.56 0.85 1.29 9.73 Fe 1.03 0.53 0.33 0.23 1.27 0.21 0.84 1.05 1.00 1.48 3.34 0.99 5.60 Ni 0.17 0.45 0.24 0.04 5.80 0.27 0.49 0.51 0.05 1.16 0.89 0.14 5.01 Al 0.17 1.33 1.01 0.86 2.02 0.97 0.90 1.69 0.87 2.50 0.91 1.13 7.50 Si 1.00 1.61 1.10 0.99 3.37 0.89 1.07 1.83 1.38 2.94 5.84 1.44 8.49 P 1.12 1.06 1.30 1.41 b.d. 1.818 1.16 1.59 4.95 1.43 1.318 3.14 14.06 S 1.49 0.73 0.06 0.06 9.45 0.14 0.15 0.68 0.16 0.53 4.92 0.26 1.17 Mn 0.16 0.62 1.18 1.00 5.41 0.495 0.75 1.32 1.06 0.67 1.91 b.d. b.d. p(3) 0.98 0.20 0.76 1.00 1.00 0.81 0.45 0.51 1.00 0.94 0.87 0.58 0.99 2 2 2 3 Table 7. Sample ID numbers. Table 20160229 Andromeda G1,S1 thin blade 1 20151203 Andromeda G1,S1 - Stack 9 - Full Area Corrected 1 20151203 Andromeda G1,S1 - Stack 9 - Full Area Corrected 1 20151203 Andromeda G1,S1 - Stack 9 - Full Area Corrected 1 20151203 Andromeda G1,S1 - Stack 9 - Full Area Corrected 1 Calculated in discussion 1 20151203 Andromeda G1,S1 - Stack 5 - Sulfide 1 20131219 - Febo 1 20151203 Andromeda - Stack 2 - FGM Bulk 2 20160229 Andromeda -Mg-C 2 20151203 Andromeda - Stack 3 - EA1 2 20160229 Andromeda -Mg-D 2 20160229 Andromeda -Mg-B 2 20151203 Andromeda - Stack 4 - Daisy 2 20160130 Andromeda - Duck2 sum stacks 2 20160130 Andromeda - Duck1 Stack 3 2 20160229 Andromeda -Mg-E 2 20160229 Andromeda -Mg-F 2 20160229 Andromeda -Mg-A 2 20151203 Andromeda - Stack 6 - Robodog 2 20151203 Andromeda - Stack 7 2 20160130 Andromeda - Duck1 Stack 3 3 20160130 Andromeda - Duck1 Stack 3 3 20160130 Andromeda - Duck1 Stack 3 3 20160130 Andromeda - Duck1 Stack 3 3 20160130 Andromeda - Duck1 Stack 3 3 20160229 Andromeda - Mg-E - Fe Core 3 20160229 Andromeda - Mg-E - Fe Side 3 20170809 - Lambda B4 S7 - Stack 6 - Magnetite 3 4 20151203 Andromeda G1,S1 - Stack 5 - CPX aThe thickness correction has been chosen to optimize for troilite. Spectrum UC Berkeley ID Sulfide01 Sulfide02 Sulfide03 Sulfide04 Sulfide05 Sulfide06 Sulfide07 Sulfide08 FGM09 FGM10 FGM11 FGM12 FGM13 FGM14 FGM15 FGM16 FGM17 FGM18 FGM19 FGM20 FGM21 FeOx22 FeOx23 FeOx24 FeOx25 FeOx26 FeOx27 FeOx28 FeOx29 Px30 Σ Peak Counts 6.1 · 104 3.2 · 108 3.1 · 106 9.1 · 105 4.4 · 105 N/A 1.1 · 104 6.7 · 107 3.8·106 1.4·106 1.8·106 7.3·105 1.2·106 1.2·105 8.8·105 5·105 3.9·105 5.6·105 1.6·106 1.3·106 4.9·105 1.9·105 3.9·104 1·104 1.7·104 1.3·104 1.1·104 7·103 3.6·105 9.9 · 105
1507.03473
2
1507
2015-09-21T09:44:53
Modeling the variations of Dose Rate measured by RAD during the first MSL Martian year: 2012-2014
[ "astro-ph.EP", "astro-ph.SR", "hep-ex", "physics.space-ph" ]
The Radiation Assessment Detector (RAD), on board Mars Science Laboratory's (MSL) rover Curiosity, measures the {energy spectra} of both energetic charged and neutral particles along with the radiation dose rate at the surface of Mars. With these first-ever measurements on the Martian surface, RAD observed several effects influencing the galactic cosmic ray (GCR) induced surface radiation dose concurrently: [a] short-term diurnal variations of the Martian atmospheric pressure caused by daily thermal tides, [b] long-term seasonal pressure changes in the Martian atmosphere, and [c] the modulation of the primary GCR flux by the heliospheric magnetic field, which correlates with long-term solar activity and the rotation of the Sun. The RAD surface dose measurements, along with the surface pressure data and the solar modulation factor, are analysed and fitted to empirical models which quantitatively demonstrate} how the long-term influences ([b] and [c]) are related to the measured dose rates. {Correspondingly we can estimate dose rate and dose equivalents under different solar modulations and different atmospheric conditions, thus allowing empirical predictions of the Martian surface radiation environment.
astro-ph.EP
astro-ph
Modeling the variations of Dose Rate measured by RAD during the first MSL Martian year: 2012-2014 Jingnan Guo1, Cary Zeitlin2, Robert F. Wimmer-Schweingruber1, Scot Rafkin3, Donald M. Hassler3, Arik Posner4, Bernd Heber1, Jan Kohler1, Bent Ehresmann3, Jan K. Appel1, Eckart Bohm1, Stephan Bottcher1, Sonke Burmeister1, David E. Brinza5, Henning Lohf1, Cesar Martin1, H. Kahanpaa 6, Gunther Reitz 7 ABSTRACT The Radiation Assessment Detector (RAD), on board Mars Science Laboratory’s (MSL) rover Curiosity, measures the energy spectra of both energetic charged and neutral particles along with the radiation dose rate at the surface of Mars. With these first-ever measurements on the Martian surface, RAD observed several effects influ- encing the galactic cosmic ray (GCR) induced surface radiation dose concurrently: [a] short-term diurnal variations of the Martian atmospheric pressure caused by daily ther- mal tides, [b] long-term seasonal pressure changes in the Martian atmosphere, and [c] the modulation of the primary GCR flux by the heliospheric magnetic field, which cor- relates with long-term solar activity and the rotation of the Sun. The RAD surface dose measurements, along with the surface pressure data and the solar modulation factor, are analysed and fitted to empirical models which quantitatively demonstrate how the long-term influences ([b] and [c]) are related to the measured dose rates. Correspond- ingly we can estimate dose rate and dose equivalents under different solar modulations and different atmospheric conditions, thus allowing empirical predictions of the Mar- tian surface radiation environment. 1Institute of Experimental and Applied Physics, Christian-Albrechts-University, Kiel, Germany [email protected] 2Southwest Research Institute, Earth, Oceans & Space Department, Durham, NH, USA 3Southwest Research Institute, Space Science and Engineering Division, Boulder, USA 4NASA Headquarters, Science Mission Directorate, Washington DC, USA 5Jet Propulsion Laboratory, California Institute of Technology, Pasadena, CA, USA 6Finnish Meteorological Institute, Helsinki, Finland 7 Aerospace Medicine, Deutsches Zentrum fur Luft- und Raumfahrt, Koln, Germany – 2 – Subject headings: space vehicles: instruments – instrumentation: detectors – Sun: solar-terrestrial relations – GCR radiation – Manned mission to Mars – Predictions of dose rate 1. Introduction and Motivation The assessment of the radiation environment is fundamental for planning future human mis- sions to Mars and evaluating the impact of radiation on the preservation of organic bio-signatures. Contributions to the radiation environment on the Martian surface are very complex (e.g., Saganti et al. 2002; Dartnell et al. 2007; Ehresmann et al. 2011; Kohler et al. 2014; Ehresmann et al. 2014): en- ergetic particles entering the Martian atmosphere either pass through without any interactions with the ambient atomic nuclei, or undergo inelastic interactions with the atmospheric nuclei creating secondary particles (via spallation and fragmentation processes), which may further interact while propagating through the atmosphere. Finally all primary and secondary particles reaching the sur- face may also interact with the regolith and, amongst others, produce neutrons which could be backscattered and detected as albedo neutrons (e.g., Boynton et al. 2004). Therefore, the radiation environment measured at the surface of the planet is determined by the characteristics of the pri- mary radiation incident at the top of the atmosphere, the composition and mass of the atmosphere, and the composition and density of the surface soil. The above process can be described by a simplified mathematical equation: F j(z, t, E) = ∞ Z 0 Xi Mi j(z, E, E0)F0i(E0, t)dE0, (1) where E0 is the energy of a primary particle with species i (e.g., protons, alpha particles and heavy ions) reaching the top of the Martian atmosphere; F0i(E0, t) is the spectrum (in the unit of counts/MeV/sec/cm2) of primary particle type i at time t; Mi j(z, E, E0) is the yield matrix, repre- senting the interaction between particles and the atmosphere (and the regolith), of particle type i with energy E0 generating particle type j with energy E; Mi j therefore depends on the altitude z (or atmospheric pressure P); finally F j(z, t, E) (in the unit of counts/MeV/sec/cm2) is the result- ing particle spectra of type j at time t. Compared to Neutron Monitors on Earth which measure the count rates of secondary particles generated by primary fluxes going through the atmosphere (Clem & Dorman 2000), RAD measures a mix of primary and secondary particles. Further, there is no need to include the geomagnetic cutoff energy in the case of Mars due to the absence of a global magnetic field. – 3 – There are predominantly two types of primary particles (F0i(E0, t)) reaching Mars: galactic cosmic rays (GCRs) and solar energetic particles (SEPs). SEPs are sporadic and impulsive events and take place much more frequently during the active phase of the solar cycle. SEPs are mainly protons, electrons and α particles with energy typically ranging from 10 to several hundreds of MeV. GCRs generally originate from outside the Solar System, e.g. in supernova remnants, and their composition consists mainly of protons, ∼ 7-10% helium and ∼ 1 % heavier elements. Due to the scattering effect of the magnetic fields in interstellar space, charged GCRs are subject to continuous deflection and the observed spectra are mostly isotropic. The GCR flux in the Solar System is inversely modulated by the variations of the solar activity (e.g., Parker 1958). In the long term, during solar activity maximum the increased solar and heliospheric magnetic fields are more efficient at hindering low-energy GCRs from entering the inner heliosphere (e.g., Heber et al. 2007; Wibberenz et al. 2002) than at solar activity minimum when the interplanetary magnetic field strength are reduced (Goelzer et al. 2013; Smith et al. 2013; Connick et al. 2011). Conse- quently, the GCR population is most intense during solar minimum (e.g., Mewaldt et al. 2010; Schwadron et al. 2012). In the short term, the GCR spectrum can also be altered indirectly by solar events such as coronal mass ejections (CMEs) where the enhanced interplanetary magnetic field can sweep away a fraction of GCRs causing reductions in GCR doses in the form of Forbush decreases (Forbush 1938; Schwadron et al. 2012, 2014b). The Mars Science Laboratory (MSL) spacecraft (Grotzinger et al. 2012), carrying the Curios- ity rover, was launched on November 26, 2011 and the rover landed on Mars on August 6, 2012. The Radiation Assessment Detector (RAD, Hassler et al. 2012) onboard is an energetic particle detector and carried out radiation measurements during its cruise from Earth to Mars (Zeitlin et al. 2013) and now continues to do so on the surface of the planet (Hassler et al. 2014). The solar modulation of the dose rate measured by RAD during the MSL cruise phase has been discussed by Guo et al. (2015b). Typically, CME and SEP events are more common during solar maximum, but the latest solar maximum has been a very weak one compared to space-age averages (e.g., Schwadron et al. 2011; McComas et al. 2013; Komitov & Kaftan 2013). This may have been caused by the reduced solar wind pressure (e.g., McComas et al. 2008; Schwadron et al. 2014a) in the deep cycle 23- 24 minimum which has allowed the termination shock to move closer to the Sun and led to a weakened modulation of the heliosheath (Scherer et al. 2011). There were only three solar particle events detected by RAD over its first Martian year 1 on the surface of Mars. They can be seen in the middle panel of Figure 1 as three peaks of the measured dose rate (the dose rate measurement will be explained in more detail in Section 2) at sols 242, 420, and 504 respectively. The dose 11 Martian year ≈ 668 sols; 1 sol = 1 Mars day ≈ 1.03 earth day – 4 – rates during these SEP events were several times higher than the quiet-time dose rate. For a closer inspection of the variations of the GCR-driven dose rate during the solar quiet periods, we omit the peak values of the SEPs in this figure. A zoomed-out figure containing the surface dose rate measurements for the first 300 sols, as well as the peak doses of the first SEP can be found in Hassler et al. (2014). The bottom panel of Figure 1 shows the Mars surface atmospheric pressure as measured by REMS (G´omez-Elvira et al. 2012), while the middle one shows the dose rate measured by RAD. Both panels show high time resolution data as a shaded band and one-sol averages as solid dots. The variations in the shaded bands are real short-term (diurnal) oscillations in dose rate and are anti-correlated with corresponding diurnal pressure variations. A zoomed-in figure containing one- sol variations of the dose rate and the pressure can be found in Rafkin et al. (2014). However, some of the long-term evolutions, e.g., the drops in both dose rates and pressure in the time period pro- ceeding sol 200 to those around and after sol 300, can not be well explained by the anti-correlation between them. The aim of this paper is to investigate the relative influences of atmospheric pres- sure (bottom panel) and heliospheric modulation (top panel) on the measured dose rate (middle panel). During solar-quiet periods, several factors on different time scales may affect the variation of the GCR-induced dose rates measured by RAD on the surface of Mars: [a] the short-term diurnal variations of the Martian column mass (measured as surface pressure in a hydrostatic atmosphere) at Gale crater caused by daily thermal tides (Rafkin et al. 2014), [b] the long-term seasonal changes of the atmospheric pressure shown in the bottom panel of Figure 1, and [c] the modulation of the primary GCR flux by the solar magnetic field which correlates with the solar activity and the rotation of the heliosphere. A commonly used parameter of heliospheric modulation is the modulation potential Φ (Gleeson & Axford 1968; Usoskin et al. 2005) which corresponds to the mean electric potential that quantifies the energy loss of a cosmic ray particle experiences inside the heliosphere and is often used to parametrize the modulation of the GCR spectrum. The modulation potential allows specification of distributions across a range of GCR particle species with different nucleons A and charge state Z and has been often used in determining GCR spectra and flux based on analytic models (e.g., ONeill 2006). In other words, the GCR-driven primary flux in Equation 1 is a function of Φ, i.e., F0i(Φ) and the yield matrix, Mi j, is a function of atmospheric pressure, P; the resulting particle spectra, F j(E), along with the dose rate which is a good measure of the radiation environment, are consequently a function of Φ as well as pressure, i.e. F j(E) = F j(Φ, P, E). This study aims to derive an empirical expression for the dose rate as a function of Φ and pressure based on observational data and thus make it possible to predict the radiation environment on the surface of Mars under different solar modulations and pressure variations. From the neutron monitor count rates (CRNM, in the unit of counts per minute) recorded by the – 5 – 2 1 8/2 0 1 2 3 2 1/2 0 1 2 5 8/2 0 1 3 1 6 0/2 0 1 3 2 6 3/2 0 1 3 1/2 0 1 4 1 0 4/2 0 1 4 2 0 6/2 0 1 4 ] V M [ n o i t a u d o M l r a o S l ] y a d / y G u [ e t a r e s o d D A R ] a P [ e r u s s e r p S M E R 850 800 750 700 650 600 550 500 240 200 160 950 900 850 800 750 700 650 0 Earth Mars Silicon Plastic 100 200 Smoothed Daily Total Mean 500 600 700 300 400 Time since Landing [sol] Fig. 1.— Top: Solar modulation potential Φ at Earth derived from Oulu neutron monitor count rate is shown in black and the per-sol-averaged Φ corrected to Mars’ location is plotted as red dots. Middle: Dose rates recorded by RAD in the silicon detector B (gray curve) and plastic scintillator E (tissue-equivalent, black curve) with their per-sol-averaged values marked as magenta- and red- dots respectively. Bottom: The pressure data from REMS (gray curve) and the per-sol-averaged values (red dots). The overall average pressure of the first 700 sols is about 840 Pascal and is marked as a thick-green line. Note that we have given the unit of time in both sol (i.e., time since the landing of MSL) at the bottom of the figure and day of year/year format at the top of the figure. – 6 – Oulu neutron monitor 2, the modulation potential Φ (in the unit of MV) at Earth can be estimated (Usoskin et al. 2002; Guo et al. 2015b) and its results are plotted in black in the top panel of Figure 1. However, the potential Φ at Earth and at Mars may differ due to the longitudinal difference of the modulation across the Parker spirals resulted from three dimensional drifts (e.g., Jokipii et al. 2004; Potgieter & Le Roux 1992) as well as the small radial gradient between 1.0 and 1.5 AU (Schwadron et al. 2010; Gieseler et al. 2008). This radial effect can be corrected following the analytic function given by Schwadron et al. (2010) and the resulting modulation Φ at Mars (per- sol-average) is plotted in red dots. It is generally smaller than Φ at Earth with variant differences (shown as red-shaded areas) through time due to varying distances of the planets to the Sun. Transient effects that are localized to narrow ranges of heliospheric longitude, such as narrow CMEs, can also perturb the GCR fluxes differently at Earth and Mars. In order to reduce the spatial longitudinal discrepancy of Φ, we use binning techniques in our current study as presented in Section 4.2.2. 2. RAD Measurements RAD measures both the charged as well as the neutral radiation environment on Mars (Hassler et al. 2012). It uses the dE/dx vs total E or multiple dE/dx techniques (Ehresmann et al. 2014; Guo et al. 2015a) to identify charged particles in a telescope stack of three silicon semiconductor detectors, A, B, and C, followed by a high-density CsI scintillator calorimeter, D. The CsI crystal together with a plastic (BC-432m) scintillator (namely detector E) are enclosed in a highly efficient antico- incidence in order to measure the neutral radiation (Kohler et al. 2014). In addition, the dose rate is measured in the silicon detector B as well as in the plastic detector E. A detailed overview of the instrument is given in Hassler et al. (2012). In this work we will determine the influence of heliospheric modulation and atmospheric pres- sure on dose rate in silicon (detector B) and in plastic (detector E). The dose rate is defined as the energy deposited by radiation per unit mass and time and is measured in Gy/day (J/kg/day). During quiet times, the dose rate on the surface of Mars is - apart from a very small natural background - determined by the GCR and its interaction with the atmosphere and soil. It can be described by the following equation: ǫmax∞ D(Φ, P) = Xj Xarea " ǫmin0 λ j(E, ǫ)F j(Φ, P, E)dEdǫ/m, (2) 2The Oulu count rate data have been obtained from http://cosmicrays.oulu.fi/ and the pressure effect has been corrected. – 7 – where E is the energy of a particle with type j, F j(Φ, P, E) (in the unit of counts/MeV/sec/cm2) is the surface particle spectrum (equivalent to F j(z, t, E) in Equation 1) which is modulated by the heliospheric potential, Φ, and atmospheric pressure, P. m (kg) is the mass of the detector and ǫ is the energy deposited by the particle in the detector. This energy deposit can be estimated using either a simple Bethe-Bloch Ansatz (Bethe 1932) or with more sophisticated models such as GEANT4 (Agostinelli et al. 2003) and HZETRN (Wilson et al. 1995) and is included as a yield matrix, λ j(E, ǫ), in the above equation. ǫmin and ǫmax are the minimum and maximum energy over which the detector is sensitive and D is the corresponding dose rate integrated over the entire detection area (area) and all the detected particle species. Dose rate is in the unit of MeV/kg/sec and can be transfered to µGy/day. Correspondingly D depends on both heliospheric potential, Φ, and on atmospheric pressure, P. RAD measures radiation doses induced by both charged and neutral energetic particles in two detectors: the silicon detector B and the plastic scintillator E (Hassler et al. 2012). RAD is directly mounted on the ’shoulder’ of the rover body (Hassler et al. 2012) and the shielding of the rover from above can be ignored. It is indeed shielded by the rover from below. However since the upward flux (of albedo particles) is much smaller than the downward flux 3, the measured dose can be roughly assumed equivalent to the radiation dose of the Martian surface. RAD operates on adjustable ”observation” cycles, with typical durations of 32 minutes early in the surface mission and 16 minutes later in the mission to date. These cadences typically yield 44 or 88 measurements per day or sol for dose rates or charged/neutral particle count rates. Detector E has a composition similar to that of human tissue and is also more sensitive to neutrons than silicon detectors. For a given incident flux, the dose rate in detector B is generally less than the dose rate in E because of the comparatively larger ionization potential of silicon as shown in Figure 1. 3. Martian Atmospheric Pressure The Martian atmosphere is roughly 1% as thick as that of the Earth’s. The dominant compo- sition of the Martian atmosphere is about 95% CO2 (Owen et al. 1977), of which 25% condenses seasonally onto the winter pole. Nevertheless, the seasonal variation of the composition has a very little effect on the surface radiation field compared to the changes of the column mass (e.g., Rafkin et al. 2014). The oscillations in atmospheric column mass drive variation of energetic par- ticle radiation at the surface with both diurnal and seasonal periods. 3More detailed analysis shows that the upward-downward ratio of the proton flux is only about 13% (Appel et al., paper in preparation). – 8 – 3.1. Diurnal Pressure Variations With a constant value of gravitational acceleration g, the surface pressure is an exact measure of the column mass given a hydrostatic atmosphere. An increase in column mass corresponds to an increase of surface pressure (e.g., Rafkin et al. 2014). The Martian atmosphere exhibits a strong thermal tide excited by direct solar heating of the atmosphere on the dayside and strong infrared cooling on the nightside. Heating causes an inflation of the atmosphere with a simultaneous drop in surface pressure. In Gale Crater, the thermal tide produces a diurnal variation of column mass of about ±5% relative to the median, as measured by the Rover Environmental Monitoring Station (REMS) (G´omez-Elvira et al. 2012). The magnitude of the diurnal pressure cycle at Gale Crater is substantially greater than previous surface measure- ments. This is likely due to the topography of the crater environment, which yields hydrostatic adjustment flows that amplify the daily tides (Haberle et al. 2014). This daily pressure oscillation can be seen in the bottom panel of Figure 1 where the pressure (shown as gray lines) expands away from the daily averaged pressure (shown as red dots) within the range of ≈ ± 50 Pascals. 3.2. Seasonal Pressure Changes The seasonal Martian atmospheric pressure variation is controlled by a complex balance be- tween the cold and warm poles (e.g., Tillman 1988; Zurek 1988). As on Earth, when the south pole is in total darkness, the north pole is experiencing continuous sunlight; one might expect that the global pressure should stay roughly constant over the year, as CO2 vaporized at one pole would freeze at the other. However, the high eccentricity of Mars’ orbit causes the insolation to be signif- icantly different between poles. Mars is farther from the Sun during northern summer; the summer in the southern hemisphere is much warmer than summer in the northern hemisphere. As a result, the north and south poles have different impacts on the atmospheric pressure changes through con- densation of CO2 to the polar region in winter and the recession of CO2 polar cap during spring and summer. The seasonal CO2 condensation cycle results in the seasonal pressure variation: as more CO2 evaporates into the atmosphere in the summer, the measured surface pressure increases as shown in the bottom panel of Figure 1. Curiosity landed at the time when the northern hemisphere was in late summer and the global pressure was near its minimum since the southern CO2 ice cap had nearly reached its maximal extent during southern hemisphere winter. As shown in Figure 1, the atmospheric pressure then began to increase as the southern polar cap started shrinking during the southern spring (northern autumn). It then reached a peak (∼ sol 175) during early northern hemisphere winter, when the – 9 – southern cap was near its minimum level and before the northern cap had grown to its maximum size. A small minimum of the pressure is present during the late northern winter when the northern cap reaches its maximum extent (∼ sol 310). The diurnal pressure oscillation (as described in Section 3.1) superimposed on this long term seasonal pressure change can be averaged out by calculating the per-sol-averaged pressure as shown by the red dots in Figure 1. The peak to peak seasonal pressure difference reaches about 25% of the average pressure. This seasonal pressure variation, like the diurnal variation, causes column density changes of the atmosphere that affect the particle fluxes measured by RAD at a seasonal period. 4. Empirically Modeling the Dose Rate as a Function of Pressure and Solar Potential As discussed in Section 1, we aim to find out how RAD measured surface radiation environ- ment depends on atmospheric pressure P and solar modulation Φ. Both influences are blended and embedded in the long-term variations of the measured GCR dose rate as seen in Figure 1. While some previous studies have attempted to model the effects of atmospheric pressure on the Martian radiation environment and dose rate 4, this work has the advantage of using actual in-situ measured data for a quantitative and empirical study. Rafkin et al. (2014) have analyzed the diurnal pressure effect on the dose rate measurement by successfully isolating the diurnal variations in the RAD measurements from the longer-term influ- ences that include seasonal atmospheric shielding and variability of the heliosphere. This method will be explained in Section 4.1.1. The authors presented a robust fit with a linear correlation be- tween pressure changes and dose rate measurements 5 providing the reasonable assumption that the solar modulation Φ varies at a time scale longer than the diurnal period. Guo et al. (2015b) have studied the solar modulation of the GCR dose rate measured by RAD during the MSL cruise phase when the pressure-variation effect was not present. Two separate empirical models were employed to describe the anti-correlation between heliospheric modulation potential Φ and measured dose rate. Both a simple linear function and a non-linear regression model could equally well represent this anti-correlation given that the shielding of the spacecraft (’pressure’) was constant. Assuming that Φ and pressure P are two independent parameters influencing the surface dose rate (while the viability of this assumption will be discussed in Section 5), the pressure effect and the solar modu- 4Ehresmann et al. (2011) calculated the surface radiation exposure for much higher atmospheric pressures which might have been present during the Noachian epoch. 5This linear correlation has been obtained at the scale of pressure values measured at Gale Crater and should not be simply extrapolated over a much wider range of pressure, e.g., to the top ot the Martian atmosphere. – 10 – lation can be linearly combined. This results into the following two models of GCR-induced dose rate variations on the surface of Mars. 1. Both the pressure effect and the solar modulation drive the variation of dose rate linearly and independently, written as: D(Φ, p) = D01 + κP + βΦ, (3) where κ (in the unit of µGy/day/Pa) is the linear correlation factor between pressure and dose rate variations and can be fitted when Φ is constant; β (in the unit of µGy/day/MV) is the linear correlation factor between solar potential and dose rate changes and can also be fitted when pressure is approximately stable; D01 (in the unit of µGy/day) is some relevant dose rate when both pressure and Φ are at certain typical levels, e.g. P0 and Φ0. 2. The solar modulation effect can also be described by a non-linear empirical model (Guo et al. 2015b) as often used in the analysis of neutron count rates (e.g., Usoskin et al. 2011). The combination of the pressure and Φ changes results in: D(Φ, P) = D02 + κP + α1 Φ + α2 , (4) with the requirement of α2 ≥ 0 to assure a positive denominator. α1 (in the unit of MV · Gy/day) and α2 (in the unit of MV) can be fitted when pressure is constant. D02 (in the unit of µGy/day) is some relevant dose rate when both pressure and Φ are at certain typical levels. 4.1. Pressure Effect The day and night variations of the dose rate have been observed for the first time on the surface of Mars by RAD, as presented by Rafkin et al. (2014) where the anti-correlation between pressure and dose rate changes has been quantitatively investigated. Due to the day and night oscillations of the atmospheric column mass, the characteristics of the particle radiation at the surface also vary diurnally. The middle panel in Figure 1 shows the dose rate measured by RAD on the surface of Mars during the first 700 sols (06/Aug/2012 - 25/Jul/2014). The dose rate varies at a diurnal level seen as oscillations of the black curve for the plastic detector, E, and the gray curve for the silicon detector, B. The variation in the E dose rate over a diurnal cycle averages about 15 µGy/day (or ∼5 %) peak to peak, out of around 220 µGy/day. The oscillation in the silicon detector is more pronounced because it includes both the diurnal variation and larger statistical fluctuations due to its smaller geometric factor. The per-sol-averaged values of the dose rates are calculated and plotted over the curves as magenta-dots (for B measurements) and red-dots (for E measurements). – 11 – Three SEP events occurred on sol 243, 421 and 504 and they are excluded from the following study of the pressure effect since the particle fluxes and energies during SEPs are substantially different from those of GCRs and the pressure response of dose rate can be heavily modified. Quantitative analysis has shown a clear inverse relation between the variations in the atmospheric pressure and the RAD dose rates with a correlation coefficient for linear regression of 96%. 4.1.1. Fitting κ using Hourly Perturbation We use the method described in Rafkin et al. (2014) to produce the average diurnal perturba- tions of the data; this approach aims at isolating the diurnal variations in the RAD measurements from the longer-term influences that include seasonal atmospheric shielding and variability of the heliosphere, as well as solar event. Generally, the dose rates measured by RAD are distributed uniformly in time while REMS’ pressure data are recorded at 1 Hz over 5 minute periods at certain periods of the sol. In order to correlate these two data sets with different time frames, we obtain the hourly averaged measure- ments of both pressure and dose rates and their corresponding hourly perturbations. The hourly averaged dose rate and pressure are Dh,s and Ph,s respectively where h represents the hour (here defined to be 24 hours on each sol) and s corresponds to the sols. The measurements, especially the pressure data, were not always uniformly taken in time and there are gaps in the data lasting longer than one hour. Therefore we only consider sols where there are at least 20 hours with pres- sure measurements in each hour. Then we interpolate the pressure data using a spline interpolation method so that data are distributed uniformly and continuously over 24 hours through that sol and systematic errors of the hourly average and per-sol-average could be minimized. Finally the in- terpolated pressure data can be binned into 24 bins for each sol and the hourly binned pressure is the corresponding hourly average Ph,s. The hourly pressure perturbation δPh,s is defined as the difference between the hourly average, Ph,s, and the total average, Ps, of the corresponding sol: δPh,s = Ph,s − Ps = Ph,s − Ph,s 24 Ph=1 24 . (5) In this case, the hourly perturbation is isolated from the sol-to-sol changes; δPh,s can further be av- eraged through all sols to obtain the mean hourly perturbation ¯δPh and the corresponding standard error in that hour. The same binning technique has also been applied to RAD dose rate measure- ¯δDh, can be readily correlated with the hourly ment. The mean hourly perturbation of dose rate, ¯δPh, and their relationship follows a clear anti-correlation which can be fitted pressure perturbation, with a first-order polynomial function: ¯δDh = κd × ¯δPh. (6) – 12 – 4.1.2. Pressure Effect on Dose Rates Using the binning method described in Section 4.1.1, we fitted the diurnal dose rate variations (measured by detector E) as a response of pressure oscillations through the first MSL Martian year as shown in Figure 2. The correlation coefficient for linear regression is 98.8%. The fitted propor- tionality factor, κ, is -0.13 ± 0.02 µGy/day/Pa where the error bar is obtained through propagation of the uncertainties of the hourly perturbation data. The measured dose rate decreases when the pressure increases since κ is negative. This indicates that the atmosphere has a shielding effect on surface dose rates. The proportionality factor obtained here is consistent, within error bars, with what obtained by Rafkin et al. (2014) which was about 0.116 µGy/day/Pa. The slight difference is caused by the selection of different periods (Rafkin et al. (2014) used the first 350 sols of data), and by the method used here to interpolate data into a uniform distribution in time before the bin- ning process. We have applied the same method to fit the pressure effect on dose rates deposited in detector B and found the correlation coefficient to be 96.4% and the proportionality factor to be -0.10 ± 0.06 µGy/day/Pa which is smaller than that for the E detector since dose rates in B are generally smaller. Note that compositional changes in the atmosphere have only a vanishingly small effect on surface radiation compared to the pressure variations (Rafkin et al. 2014). Similarly, the occasional presence of dust in the atmosphere during dust storms is negligible6. 4.2. Solar Modulation The pressure variations and the modulation of the primary GCR radiation outside the atmo- sphere are the two factors which determine the longterm dose rate variations on the surface of Mars as explained in Section 2. In order to separate these two effects, we subtract the pressure effect using the proportionality of the variations found above. 4.2.1. Subtracting the Pressure Effect As we can already isolate the pressure effect and obtain the proportionality factor κ in Equa- tion 3 and 4, the effect of solar modulation can be investigated by first subtracting the pressure 6A PLANETOCOSMIC simulation has been carried out to derive the particle fluxes on the surface of Mars con- sidering the existence of dust storms. The result shows that the effect of dust storms is very small (Appel et al., paper in preparation). – 13 – SOL: 13-682 κ = -0.1306 ± 0.0176 −20 0 δ Pressure [Pa] 20 40 10 5 0 −5 ] y a d / y G u [ e t a r E e s o d δ −10 −40 Fig. 2.— Hourly perturbation of dose rate ¯δDh versus hourly perturbation of pressure ¯δPh through sol 13 to 682 (approximately one Martian year) as shown in blue. The error bars stand for the standard deviation of the averaged hourly perturbation. The fitted anti-correlation is shown as a red line with a slope of κd being -0.1306 ± 0.0176 µGy/day/Pa. – 14 – 3 2 1/2 0 1 2 5 8/2 0 1 3 1 6 0/2 0 1 3 2 6 3/2 0 1 3 1/2 0 1 4 1 0 4/2 0 1 4 Silicon: daily Silicon: pres. constant Plastic: daily Plastic: pres. constant Modulation Phi 900 800 700 ] V M 600 [ i h P n o i t a u d o M l 500 400 300 200 100 Time since landing [sol] 300 250 200 150 ] y a d / y G u [ e t a r e s o d D A R Fig. 3.— RAD dose rates (µGy/day, left axis) and solar modulation potential Φ at Mars’ radial distance (MV, think-black-dotted-line, right axis) through sol 13 to 682 (approximately one Mar- tian year). RAD dose rates from silicon detector B is in blue and from plastic detector E is in red. Thick-green and thick-cyan lines are the resulting dose rates (D′ in Equation 7) assuming pressure is constant at P0. Note that we have given the unit of time in both sol (i.e., time since the landing of MSL) at the bottom of the figure and day of year/year format at the top of the figure. – 15 – effect from dose rate and then correlating the ’constant-pressure’ dose rate with solar modulation potential Φ. The method can be described by restructuring, for instance Equation 3 as following: D′(Φ) = D′ 01 + βΦ, (7) where D′ = D − κ(P − P0) and D′ 01 = D01 + κP0. Employing all the per-sol-averaged dose rate data collected over one Martian year (sol 13 to 682), we can subtract the seasonal pressure effect assuming the surface pressure is constant at a particular value, i.e., P0 = 840 Pascals, which is the averaged pressure found by REMS during this time period as shown in Figure 1. By calculating the difference of the per-sol-averaged pressure P and P0 and using κ found in Section 4.1.2, we estimate the pressured-induced dose rate to be κ(P−P0) which is then subtracted from the long-term dose rate measurement with the ’constant-pressure-as-P0’ dose rate remaining, namely D′ being solely a function of solar modulation potential Φ. Figure 3 shows the RAD dose rates measured by both silicon and plastic detectors before (thin lines) and after (thick lines) the pressure correction. Also plotted is the solar modulation Φ which already shows an anti-correlation with the constant- pressure dose rates. 4.2.2. Correlation of dose rates and Φ The modulation potential, Φ, is derived from the Oulu neutron monitor on Earth (at 1AU) and corrected to the radial distance of Mars (∼ 1.5 AU) following Schwadron et al. (2010). However Mars and Earth are not always magnetically well connected and cross-field diffusion and drift can be extremely important. In other words, the modulation process is fundamentally 3-dimensional and Φ can not directly represent modulations at Mars7. In order to average out the cross-field discrepancy of the modulation between Earth and Mars and to smooth out the rotation of the heliospheric magnetic fields, we bin the per-sol-averaged data as shown in Figure 3 into 26-sol- averaged bins and correlate the binned Φ values and constant-pressure dose rates. The correlation coefficients are -0.66 and -0.56 for detectors B and E respectively, clearly indicating a negative correlation between solar modulation and GCR induced dose rates. We applied regression fittings of both the linear and non-linear models described by Equations 3 and 4 to data from both detectors as shown in Figure 4. In order to reliably propagate the uncertainties due to both measurements and binning pro- cesses, we have carried out a bootstrap Monte Carlo simulation (Efron 1981). Simulated data sets are generated using the uncertainty range of the binned data which are then fitted to the models; 7During the cruise phase MSL was mostly magnetically connected with the Earth (Posner et al. 2013; Guo et al. 2015b) and directly correlating Φ measured at Earth and the RAD dose rate was sensible. – 16 – ] y a d / y G u [ e t a r e s o d D A R 350 300 250 200 150 100 50 400 Silicon Plastic 1000 600 800 Solar modulation potential [MV] Fig. 4.— 26-sol binned data and fittings processed through bootstrap Monte Carlo simulations of the variations of the RAD dose rate (with pressure assumed to be constant P0 = 840 Pa) and the solar modulation Φ on the surface of Mars through sol 13 to 681.The red/magenta line and area represent the linear fits with standard errors (Eq. 7) of constant-pressure dose rate in detector E/B versus solar modulation potential. The green/cyan line and area show the results of the non-linear fits (Eq. 4). – 17 – dose rate results at a wider range of Φ are estimated simultaneously at each fit; 500 simulated fits were processed for each model; for every Φ value ranging from 250 to 1200 MV, 500 dose rate values were generated and their mean is taken to be the ’predicted’ dose rate while their standard deviation as the uncertainty. The fitting results of the two separate models applied to both detectors are presented below. • The fitted parameters β and D′ 01 for the linear model are obtained as the mean values of the 500 fitted parameters and their uncertainties are the standard deviation of the 500 Monte Carlo fits. For the silicon detector, B, we obtained the parameters of the linear model as 01 = 224.10 ±16.12 µGy/day and β = −0.12 ±0.03 µGy/day/MV. For plastic detector E, the D′ fitted parameters are D′ 01 = 267.17±16.29 µGy/day and β = −0.11±0.03 µGy/day/MV. Note 01 are not essential to our study and they depend on the choice that the absolute values of D′ of normalized pressure, P0. The parameter β however, as shown in Equation 7, directly represents the linear dependence of dose rate changes on the solar modulation potential, Φ. The results obtained from measurements of the two different detectors are consistent with each other within error bars. • The non-linear model has three fitting parameters: α1, α2 and D′ 02 = D02 + κP0. Their values and error bars are also obtained using the same Monte Carlo method. For silicon detector, 02 = 89.4 ± 14.0 MV, α1 = (3.7 ± 0.8) × 104 MV µGy/day and B, the fitted parameters are D′ α2 = (9.9 ± 0.1) × 10−4 MV. For plastic detector, E, the results are D′ 02 = 149.6 ± 15.3 MV, α1 = (3.3 ± 0.8) × 104 MV µGy/day and α2 = (1.3 ± 0.1) × 10−3 MV. The small absolute values of α2 indicates that it could be ignored in the model for the range of data in the current measurement. The ’predicted’ dose rates (with constant surface pressure P0) at given Φ values from 250 MV to 1200 MV were estimated for both models and are plotted in Figure 4. The uncertainty of the predicted dose rate increases when the extrapolation is further away from the actual measurements. The linear model often predicts a smaller dose rate than the nonlinear model and this difference is bigger for small solar potentials, i.e., during solar minimum. For instance, at Φ = 250 MV the discrepancy between the two models is as large as 80 µGy/day for the silicon detector, B. Because the predictions of the two models differ substantially at solar extreme conditions, choosing one or the other for predicting the radiation exposure of an astronaut would make a big difference. The data themselves do not rule out one or the other because the time period for which they are currently available does not cover a sufficiently large range of Φ. This underlines the importance of acquiring more data over an extended period of time to adequately cover the entire solar activity cycle. – 18 – 4.2.3. Estimates of the Surface Radiation Environment under different Φ and P The annually averaged modulation potential Φ reconstructed from Oulu neutron monitor data (Usoskin et al. 2011) from 1937 until 2014 can be used to calculate the corresponding ’expected’ dose rate predicted by our models. We can also use the sunspot number predicted by the U.S. Dept. of Commerce, NOAA, Space Weather Prediction Center (SWPC) for the coming years (2015 to 2019) to estimate the corresponding solar potential following a correlation study of monthly Φ and sunspot numbers (Guo et al. 2015b). The annual modulation potential, Φ, extrapolated to 1.5 AU radial distance from the Sun through year 1937 until 2019 is shown in the top panel of Figure 5. It shows a clear 11-year cycle and varies from less than 250 MV to more than 1200 MV. For evaluating the space radiation environment, dose equivalent is often derived and can be assumed to be proportional to the risk of lifetime cancer induction via population studies (ICRP60). Its approximated value, in Sieverts (Sv), is taken to be ,< Q > × D , where D is measured tissue- equivalent dose and < Q > is the average quality factor which is a conventional parameter for radiation risk estimation. A < Q > value of 3.05 ± 0.26 was found from RAD’s measurements on the surface of Mars over the first 350 sols (Hassler et al. 2014). Multiplying the tissue-equivalent dose rates (directly measured by the plastic detector E or modeled by e.g. HZETRN model) with the average quality factor yields the dose equivalent rate of GCR fluxes. For dose rates measured by detector B, a silicon to water conversion factor of 1.38 has to be applied first (Zeitlin et al. 2013) which approximately relates energy loss per unit of path length (dE/dx) in silicon to Linear Energy Transfer (LET) in water. RAD measured dose equivalent is thus most comparable to high-water content skin dose equivalent. The body effective dose can be further derived as the weighted sum of different organ dose equivalent (skin, eye, bone, brain, heart, etc.) and the weighting factor for each organ can be found in the National Council on Radiation Protection and Measurements (Linton & Mettler Jr 2003). The extrapolated dose rates at different modulation potential values have been estimated via two different models and two different detectors as shown in Figure 4. These four sets of modeled dose rates can be transfered into dose equivalent rate and the results mostly agree with each other within error bars 8. Because we can not currently determine which of the two models is the better approximation when the solar potential is outside the measured range, we let all the four modeled values serve as possible results and take their mean as predictions and the propagated errors as uncertainties. The thus final modeled dose equivalent rates and their standard deviations are shown in Figure 5. 8Figure 4 in Guo et al. (2015b) has shown the modeled dose equivalent rate during the cruise phase estimated by both linear and non-linear models based on dose rates from both detectors. The results are consistent within error bars with exceptions during extreme solar conditions for which the non-linear model predicts higher dose rates. – 19 – 1000 800 600 400 200 ] V M [ i h P n o i t a u d o M l r a o S l ] y a d / v S m 1.2 1.0 l [ e t a r t n e a v u q e e s o D i Years P0: 700 P0: 970 P0: 840 0.8 0.6 0.4 1940 1950 1960 1970 1980 1990 2000 2010 2020 Fig. 5.— Top Panel: Annual average of reconstructed modulation potential Φ at the radial distance of Mars (gray dotted line) since 1937 until 2014 (on the left side of the vertical dashed line) as well as the predicted values from year 2015 to 2019 (on the right side of the vertical dashed line). Bottom Panel: Modeled annual dose equivalent rate following the evolution of Φ assuming different seasonal surface pressures: blue for p0 = 700 Pa, green for p0 = 970 Pa and red for P0 = 840 Pa. – 20 – Note that the modeled dose equivalent rate has been derived based on the ’constant-pressure- at-840 Pa’ dose rate (see Section 4.2.1). However, the seasonal surface pressure at Gale crater, as shown in Figure 1, expands between 700 and nearly 1000 Pascals which would result in roughly 0.1 mSv/day of dose equivalent rate changes 9 and this is a considerable portion ∼ 16% of annual averaged surface dose equivalent rate ∼ 0.7 mSV/day. In order to show this seasonal pressure effect, we also estimated the dose equivalent rate (at various solar potentials) when pressure is 700 and 970 Pa respectively as shown Figure 5. The dose equivalent rates are inversely related to the surface pressure, although the seasonal pressure influence is much smaller than the longer term effects driven by solar modulation. The estimated surface dose equivalent rate ranges from about 0.35 mSv/day to about 1.15 mSv/day and has a clear anti-correlation with Φ as expected from both models. Stronger solar modulation leads to a decrease of dose equivalent rate and at Φ > 1000 MV the dose equivalent rate can be as low as ∼ 0.35 mSv/day within the uncertainties, considerably smaller than the cur- rent averaged measurement. At solar maximum and minimum when the modulation potential is further away from the measured range, the uncertainty of the estimations increases due to the large discrepancy between the models. Future measurements over solar minimum periods are essential for improving the predictions at low modulation potentials. Due to the shielding of the atmosphere, the current surface dose equivalent rate is only about 40% of the RAD cruise measurement ∼ 1.8 mSv/day (Zeitlin et al. 2013) and that from the Cosmic Ray Telescope for the Effects of Radiation (CRaTER) on the Lunar Reconnaissance Orbiter ∼ 1.6 mSv/day (Schwadron et al. 2014b). Consequently, our estimations of the surface dose equivalent rate over solar minimum and maximum periods are much smaller than the pre- dictions based on deep space measurements given by RAD (cruise phase, Guo et al. 2015b) and CRaTER (Schwadron et al. 2014b). The predictions of dose equivalent rates behind these space- craft shielding conditions are between about 1 mSv/day (solar maximum) and 5 mSv/day (solar minimum). 5. Discussion and Conclusions We have presented the dose rate data collected by MSL/RAD on the surface of Mars and analyzed its short-term and long-term variations which are driven by the atmospheric pressure changes and solar modulation. Following Rafkin et al. (2014), we first analyzed the dose rate dependence on diurnal pressure oscillations over the first MSL Martian year. This pressure-driven 9250 Pascals of pressure change would lead to 32.5 µGy/day of dose rate difference which is then transfered to dose equivalent rate via the quality factor. – 21 – effect on dose rate changes in the longterm is then subtracted to recover the constant-pressure dose rate which is then correlated with solar modulation potential Φ. A clear anti-correlation, with a correlation coefficient of about -0.6 between Φ and the recovered dose rate, suggests that, as expected for higher solar activity, GCR particles are more attenuated and the dose rate is decreased. We carried out a quantitative study of this anti-correlation and fitted two models to the mea- sured data using a bootstrap Monte-Carlo method to estimate the uncertainties. The predictions of the two models for solar activity minimum differ substantially and the data are insufficient to decide which of the two models should be used. Schwadron et al. (2014b) have estimated the deep space dose rate at different modulation potential derived from HZETRN model and the dose rate is indeed non-linearly dependent on Φ. However the shape (parameters) of this analytically de- rived model is different from that of our empirically modeled function. This highlights the need for extended measurements to cover solar activity minimum and an entire activity cycle and these observations can be used to constrain the analytic models. The extrapolated dose equivalent rate at various modulation potentials and different surface pressures are shown in Figure 5. The predicted dose equivalent rate during solar maximum years (e.g year 1991) when Φ ∼ 1100 MV was found to be as low as ∼ 0.35 mSv/day, which is consider- ably lower than the current surface measurement ∼ 0.7 mSv/day since the current solar maximum is atypically quiet. The modeled dose equivalent rate under solar minimum conditions can be as high as 1.15 mSv/day within the uncertainties. The seasonal pressure changes may affect the esti- mated dose equivalent rate at a level of about 0.1 mSv/day. Although this is less than the long-term solar modulation effect, it should not be ignored. Based on the solar modulation potential predicted for the next five years (Guo et al. 2015b), we estimate a trend of increasing dose equivalent rate (between 0.56 and 0.84 mSv/day) from 2015 until 2020. The correlation between heliospheric potential and RAD-measured dose rates during the cruise phase was investigated by Guo et al. (2015b) with the data modeled by both linear and non-linear functions. The linear dependence β of dose rate on Φ was -0.39 ± 0.07 µGy/day/MV for the silicon detector B and -0.44 ± 0.04 µGy/day/MV for the plastic detector E. In comparison, β derived here for the surface case are much smaller: -0.12 ± 0.03 and -0.11 ± 0.03 µGy/day/MV for B and E respectively. This is because [a] the magnetic connection between solar modulation potential Φ measured at Earth and dose rates evaluated during the cruise was very good, allowing the direct and thus stronger correlation of the daily values; and more importantly: [b] low energy particles which are more affected by solar modulation make a bigger contribution to the GCR doses detected during the cruise phase than to the doses measured at the surface of Mars. The process of a GCR spectrum penetrating through the Martian atmosphere can be simulated using for instance the PLANETOCOSMICS toolkit (Desorgher et al. 2006). Given typical Martian atmospheric con- ditions, we have found that protons with energies less than 170 MeV do not reach the bottom of – 22 – Gale Crater. The penetration energy is species dependent and increases with increasing ion charge. Instead, the spacecraft shielding was highly non-uniform while nearly 50% of incoming particle trajectories within the field of view of RAD from above were only lightly shielded (Zeitlin et al. 2013), thus allowing more low-energy particles to contribute to the dose rate. We will carry out more GEANT4 (Agostinelli et al. 2003) and PLANETOCOSMICS simulations as well HZETRN modelling in the future to study the interactions of GCR spectra with different shielding conditions (spacecraft and the atmosphere) in order to derive modeled function of the dose rate dependence on pressure and solar potential which can be compared with our empirical functions. Both models have assumed the independent effect of pressure and Φ on dose rates. However, this may be modified when pressure and Φ change over wider ranges than have been observed to date: • a much thinner atmosphere will allow more lower-energy particles to reach the surface which experience stronger modulation (e.g., bigger β in the linear model). Therefore, for signifi- cant pressure changes, β could be a function of P , i.e., β(P); • much stronger solar modulation (bigger Φ) would lead to a larger fraction of high-energy particles in the GCR flux and these energetic particles are less affected by the atmosphere (smaller κ); when pressure is much higher and the surface atmospheric depth is approach- ing the Pfotzer maximum (e.g., Richter & Rasch 2008), most primary particles are shielded while more secondary particles are generated and this may result in a decreased shielding effect; therefore the dependence of dose rate on pressure may be modified as P and Φ change substantially, i.e., κ = κ(P, Φ); • both the linear and non-linear models are empirical and derived from measurements; despite the robustness of the fitting of the actual data, the extrapolation is highly uncertain and a complete model requires measurements over the full range of solar conditions. To verify these non-linear and second-order effects of our fitted parameters in the empirical func- tion, Monte Carlo simulations as well as analytic HZETRN modeling will be carried out for con- straining the parameters and comparing the predictions. The quality factor used to derive equivalent dose from dose may also be sensitive to substantial changes in solar modulation because the estimation of the average quality factor depends on the spectra of particles depositing energies in the detector (Schwadron et al. 2014b). Hitherto, the measured < Q > has been quite stable since [a] the RAD measurements have undergone only small changes of the solar modulation and [b] the modulation affects more the low-energy ions which are more likely to be shielded by the atmosphere and contribute little to the surface spectra. The – 23 – change of the quality factor under much different solar conditions needs to be investigated with extended measurements. A total Mars mission GCR dose equivalent can be estimated based on our measurements and predictions of both the surface case and the cruise phase (Guo et al. 2015b). The fastest round trip with on-orbit staging and existing propulsion technologies has been estimated to be a 195-day trip (120 days out, 75 days back with an extra e.g., 14 days on the surface), as described by Folta et al. (2012). During solar maximum periods when Φ ∼ 1200MV, this would result in a GCR-induced cruise dose equivalent of 195 ± 98 mSv and surface dose equivalent of 4.9 ± 2.0 mSv, which adds to 200 ± 100 mSv during the total mission. Additional contributions to dose rate and dose equivalent rate by SEPs should not be ignored and they can differ significantly from the current measurements due to the high variability of their frequencies and intensities. Small, ”soft-spectrum” solar events where particle energies are modest will have little or no effect on the surface dose due to atmospheric shielding. For instance the total dose equivalent from the first SEP event observed on Mars (on sol 242) was only 0.025 mSv (Hassler et al. 2014) while the cruise SEP events had dose equivalent ranging from 1.2 mSv/event to 19.5 mSv/event (Zeitlin et al. 2013). Future measurements of much bigger SEP events on the surface of Mars are crucial for understanding extreme conditions of radiation environment on Mars for potential manned missions. RAD is supported by the National Aeronautics and Space Administration (NASA, HEOMD) under Jet Propulsion Laboratory (JPL) subcontract #1273039 to Southwest Research Institute and in Germany by DLR and DLR’s Space Administration grant numbers 50QM0501 and 50QM1201 to the Christian Albrechts University, Kiel. Part of this research was carried out at JPL, Cali- fornia Institute of Technology, under a contract with NASA. The sunspot data has been obtained from: WDC-SILSO, Royal Observatory of Belgium, Brussels. We are grateful to the Cosmic Ray Station of the University of Oulu and Sodankyla Geophysical Observatory for sharing their Neu- tron Monitor count rate data. The data used in this paper are archived in the NASA Planetary Data Systems Planetary Plasma Interactions Node at the University of California, Los Angeles. The archival volume includes the full binary raw data files, detailed descriptions of the struc- tures therein, and higher-level data products in human-readable form. The PPI node is hosted at http://ppi.pds.nasa.gov/. REFERENCES Agostinelli, S., Allison, J., Amako, K. a., et al. 2003, Nuclear instruments and methods in physics research section A: Accelerators, Spectrometers, Detectors and Associated Equipment, – 24 – 506, 250 Bethe, H. 1932, Zeitschrift fur Physik, 76, 293 Boynton, W., Feldman, W., Mitrofanov, I., et al. 2004, in 2001 Mars Odyssey (Springer), 37–83 Clem, J. M., & Dorman, L. I. 2000, Space Science Reviews, 93, 335 Connick, D. E., Smith, C. W., & Schwadron, N. A. 2011, The Astrophysical Journal, 727, 8 Dartnell, L., Desorgher, L., Ward, J., & Coates, A. 2007, Geophys. Res. Lett, 34, L02207 Desorgher, L., Fluckiger, E. O., & Gurtner, M. 2006, in 36th COSPAR Scientific Assembly, Vol. 36, 2361 Efron, B. 1981, Biometrika, 68, 589 Ehresmann, B., Burmeister, S., Wimmer-Schweingruber, R., & Reitz, G. 2011, Journal of Geo- physical Research (Space Physics), 116, 10106 Ehresmann, B., Zeitlin, C., Hassler, D. M., et al. 2014, Journal of Geophysical Research: Planets, 119, 468 Folta, D., Vaughn, F., Rawitscher, G., & Westmeyer, P. 2012, LPI Contributions, 1679, 4181 Forbush, S. 1938, Terrestrial Magnetism and Atmospheric Electricity, 43, 203 Gieseler, J., Heber, B., Dunzlaff, P., et al. 2008, in International Cosmic Ray Conference, Vol. 1, 571–574 Gleeson, L., & Axford, W. 1968, The Astrophysical Journal, 154, 1011 Goelzer, M. L., Smith, C. W., Schwadron, N. A., & McCracken, K. G. 2013, Journal of Geophysi- cal Research: Space Physics, 118, 7525 G´omez-Elvira, J., Armiens, C., Castaner, L., et al. 2012, Space science reviews, 170, 583 Grotzinger, J. P., Crisp, J., Vasavada, A. R., et al. 2012, Space science reviews, 170, 5 Guo, J., Zeitlin, C., Wimmer-Schweingruber, R. F., et al. 2015a, Radiation protection dosimetry, ncv297 —. 2015b, Astronomy & Astrophysics, 577, A58 Haberle, R., G´omez-Elvira, J., Torre Ju´arez, M., et al. 2014, Journal of Geophysical Research: Planets, 119, 440 – 25 – Hassler, D., Zeitlin, C., Wimmer-Schweingruber, R., et al. 2012, Space science reviews, 170, 503 Hassler, D. M., Zeitlin, C., Wimmer-Schweingruber, R. F., et al. 2014, science, 343, 1244797 Heber, B., Fichtner, H., & Scherer, K. 2007, in Solar Variability and Planetary Climates (Springer), 81–93 Jokipii, J., Giacalone, J., & Kota, J. 2004, The Astrophysical Journal Letters, 611, L141 Kohler, J., Zeitlin, C., Ehresmann, B., et al. 2014, Journal of Geophysical Research: Planets, 119, 594 Komitov, B., & Kaftan, V. 2013, Journal of Advanced Research, 4, 279 Linton, O. W., & Mettler Jr, F. A. 2003in , 321–329 McComas, D., Angold, N., Elliott, H., et al. 2013, The Astrophysical Journal, 779, 2 McComas, D., Ebert, R., Elliott, H., et al. 2008, Geophysical Research Letters, 35, 18103 Mewaldt, R., Davis, A., Lave, K., et al. 2010, The Astrophysical Journal Letters, 723, L1 Owen, T., Biemann, K., Rushneck, D., et al. 1977, Journal of Geophysical research, 82, 4635 ONeill, P. M. 2006, Advances in Space Research, 37, 1727 Parker, E. 1958, Physics of Fluids (1958-1988), 1, 171 Posner, A., Odstrcil, D., MacNeice, P., et al. 2013, Planetary and Space Science, 89, 127 Potgieter, M., & Le Roux, J. 1992, The Astrophysical Journal, 386, 336 Rafkin, S. C., Zeitlin, C., Ehresmann, B., et al. 2014, Journal of Geophysical Research (Planets), 119, 1345 Richter, J. H., & Rasch, P. J. 2008, Journal of Climate, 21, 1487 Saganti, P. B., Cucinotta, F. A., Wilson, J. W., & Schimmerling, W. 2002, Journal of radiation research, 43, S119 Scherer, K., Fichtner, H., Strauss, R., et al. 2011, The Astrophysical Journal, 735, 128 Schwadron, N., Townsend, L., Kozarev, K., et al. 2010, Space Weather, 8 Schwadron, N., Smith, C., Spence, H. E., et al. 2011, The Astrophysical Journal, 739, 9 – 26 – Schwadron, N., Goelzer, M., Smith, C., et al. 2014a, Journal of Geophysical Research: Space Physics, 119, 1486 Schwadron, N. A., Baker, T., Blake, B., et al. 2012, Journal of Geophysical Research: Planets, 117, E00H13 Schwadron, N. A., Blake, J. B., Case, A. W., et al. 2014b, Space Weather, 12, 622 Smith, C. W., Schwadron, N. A., & DeForest, C. E. 2013, The Astrophysical Journal, 775, 59 Tillman, J. E. 1988, Journal of Geophysical Research: Atmospheres (1984–2012), 93, 9433 Usoskin, I., Alanko, K., Mursula, K., & Kovaltsov, G. 2002, Solar Physics, 207, 389 Usoskin, I., Alanko-Huotari, K., Kovaltsov, G., & Mursula, K. 2005, J. Geophys. Res, 110, A12108 Usoskin, I., Bazilevskaya, G., & Kovaltsov, G. 2011, J. Geophys. Res, 116, A02104 Wibberenz, G., Richardson, I., & Cane, H. 2002, J. Geophys. Res, 107, 1353 Wilson, J. W., Badavi, F. F., Cucinotta, F. A., et al. 1995, HZETRN: description of a free-space ion and nucleon transport and shielding computer program (National Aeronautics and Space Administration, Langley Research Center) Zeitlin, C., Hassler, D., Cucinotta, F., et al. 2013, Science, 340, 1080 Zurek, R. W. 1988, Journal of Geophysical Research: Atmospheres (1984–2012), 93, 9452 This preprint was prepared with the AAS LATEX macros v5.2.
1711.06801
1
1711
2017-11-18T02:55:05
Whole planet coupling between climate, mantle, and core: Implications for the evolution of rocky planets
[ "astro-ph.EP" ]
Earth's climate, mantle, and core interact over geologic timescales. Climate influences whether plate tectonics can take place on a planet, with cool climates being favorable for plate tectonics because they enhance stresses in the lithosphere, suppress plate boundary annealing, and promote hydration and weakening of the lithosphere. Plate tectonics plays a vital role in the long-term carbon cycle, which helps to maintain a temperate climate. Plate tectonics provides long-term cooling of the core, which is vital for generating a magnetic field, and the magnetic field is capable of shielding atmospheric volatiles from the solar wind. Coupling between climate, mantle, and core can potentially explain the divergent evolution of Earth and Venus. As Venus lies too close to the sun for liquid water to exist, there is no long-term carbon cycle and thus an extremely hot climate. Therefore plate tectonics cannot operate and a long-lived core dynamo cannot be sustained due to insufficient core cooling. On planets within the habitable zone where liquid water is possible, a wide range of evolutionary scenarios can take place depending on initial atmospheric composition, bulk volatile content, or the timing of when plate tectonics initiates, among other factors. Many of these evolutionary trajectories would render the planet uninhabitable. However, there is still significant uncertainty over the nature of the coupling between climate, mantle, and core. Future work is needed to constrain potential evolutionary scenarios and the likelihood of an Earth-like evolution.
astro-ph.EP
astro-ph
Whole planet coupling between climate, mantle, and core: Implications for the evolution of rocky planets Bradford J. Foley, Peter E. Driscoll Department of Terrestrial Magnetism, Carnegie Institution for Science, Washington, DC Abstract Earth's climate, mantle, and core interact over geologic timescales. Cli- mate influences whether plate tectonics can take place on a planet, with cool climates being favorable for plate tectonics because they enhance stresses in the lithosphere, suppress plate boundary annealing, and promote hydration and weakening of the lithosphere. Plate tectonics plays a vital role in the long-term carbon cycle, which helps to maintain a temperate climate. Plate tectonics provides long-term cooling of the core, which is vital for generating a magnetic field, and the magnetic field is capable of shielding atmospheric volatiles from the solar wind. Coupling between climate, mantle, and core can potentially explain the divergent evolution of Earth and Venus. As Venus lies too close to the sun for liquid water to exist, there is no long-term carbon cycle and thus an extremely hot climate. Therefore plate tectonics cannot operate and a long-lived core dynamo cannot be sustained due to insufficient core cooling. On planets within the habitable zone where liquid water is possi- ble, a wide range of evolutionary scenarios can take place depending on initial atmospheric composition, bulk volatile content, or the timing of when plate tectonics initiates, among other factors. Many of these evolutionary trajecto- ries would render the planet uninhabitable. However, there is still significant uncertainty over the nature of the coupling between climate, mantle, and core. Future work is needed to constrain potential evolutionary scenarios and the likelihood of an Earth-like evolution. 1 1 Introduction 1.1 Overview Recent discoveries have revealed that rocky exoplanets are relatively common (Batalha, 2014). As a consequence, determining the factors necessary for a rocky planet to sup- port life, especially life that may be remotely observable, has become an increasingly important topic. A major requirement, that has been extensively studied, is that solar luminosity must be neither too high nor too low for liquid water to be stable on a planet's surface; this requirement leads to the concept of the "habitable zone," the range of orbital distances where liquid water is possible (Hart, 1978, 1979; Kasting et al., 1993b; Franck et al., 2000; Kopparapu et al., 2014). Inward of the habitable zone's inner edge, the critical solar flux that triggers a runaway greenhouse effect is exceeded. The critical flux is typically estimated at ≈ 300 W m−2, with variations of ∼ 10 − 100 W m−2 possible due to atmospheric composition, planet size, or surface water inventory (Ingersoll, 1969; Kasting, 1988; Nakajima et al., 1992; Abe et al., 2011; Goldblatt et al., 2013). In a runaway greenhouse state liquid water can not condense out of the atmosphere, so any water present would exist as steam. Fur- thermore a runaway greenhouse is thought to cause rapid water loss to space, and can thus leave a planet desiccated (Kasting, 1988; Hamano et al., 2013; Wordsworth & Pierrehumbert, 2013). Beyond the outer edge insolation levels are so low that no amount of CO2 can keep surface temperatures above freezing (Kasting et al., 1993b). However, lying within the habitable zone does not guarantee that surface condi- tions will be suitable for life. Variations in atmospheric CO2 content can lead to cold climates where a planet is globally glaciated, or hot climates where surface temper- atures are higher than any known life can tolerate (i.e. above ≈ 400 K (Takai et al., 2008)). A hot CO2 greenhouse can also cause rapid water loss to space (Kasting, 1988) (though Wordsworth & Pierrehumbert (2013) argues against this), or even a steam atmosphere if surface temperatures exceed water's critical temperature of 647 K. Moreover, the solar wind can strip the atmosphere of water and expose the surface to harmful radiation, unless a magnetic field is present to shield the planet (e.g. Kasting & Catling, 2003; Griessmeier et al., 2009; Brain et al., 2014). Atmospheric CO2 concentrations are regulated by the long-term carbon cycle on Earth, such that surface temperatures have remained temperate throughout geo- logic time (e.g. Walker et al., 1981; Berner, 2004). The long-term carbon cycle is facilitated by plate tectonics (e.g. Kasting & Catling, 2003). Furthermore the mag- netic field is maintained by convection in Earth's liquid iron outer core (i.e. the geodynamo). As a result, interior processes, namely the operation of plate tectonics 2 and the geodynamo, are vital for habitability. However, whether plate tectonics or a strong magnetic field is likely on rocky planets, especially those in the habitable zone where liquid water is possible, is unclear. The four rocky planets of our solar system have taken dramatically different evolutionary paths, with only Earth developing into a habitable planet possessing liquid water oceans, plate tectonics, and a strong, internally generated magnetic field. In particular the contrast between Earth and Venus, which is approximately the same size as Earth and has a similar composition yet lacks plate tectonics, a magnetic field, and a temperate climate, is striking. In this review we synthesize recent work to highlight that plate tectonics, climate, and the geodynamo are coupled, and that this "whole planet coupling" between surface and interior places new constraints on whether plate tectonics, temperate climates, and magnetic fields can develop on a rocky planet. We hypothesize that whole planet coupling can potentially explain the Earth-Venus dichotomy, as it allows two otherwise similar planets to undergo drastically different evolutions, due solely to one lying inward of the habitable zone's inner edge and the other within the habitable zone. We also hypothesize that whole planet coupling can lead to a number of different evolutionary scenarios for habitable zone planets, many of which would be detrimental for life, based on initial atmospheric composition, planetary volatile content, and other factors. We primarily focus on habitable zone planets, as these are most interesting in terms of astrobiology, and because the full series of surface- interior interactions we describe involves the long-term carbon cycle, which requires liquid water. Each process, the generation of plate tectonics from mantle convection, climate regulation due to the long-term carbon cycle, and dynamo action in the core, is still incompletely understood and the couplings between these processes are even more uncertain. As a result, significant future work will be needed to place more quantitative constraints on the evolutionary scenarios discussed in this review. 1.2 Whole planet coupling Several basic concepts illustrate the coupling between the surface and interior (Fig- ure 1). (1) Climate influences whether plate tectonics can take place on a planet. (2) Plate tectonics plays a vital role in the long-term carbon cycle, which helps to main- tain a temperate climate. (3) Plate tectonics effects the generation of the magnetic field via core cooling. (4) The magnetic field is capable of shielding the atmosphere from the solar wind. Cool climates are favorable for plate tectonics because they fa- cilitate the formation of weak lithospheric shear zones, which are necessary for plate tectonics to operate. Low surface temperatures suppress rock annealing, increase the negative buoyancy of the lithosphere, and allow for deep cracking and subsequent 3 water ingestion into the lithosphere, all of which promote the formation of weak shear zones. When liquid water is present on a planet's surface, silicate weathering, the primary sink of atmospheric CO2 and thus a key component of the long-term carbon cycle, is active. However, silicate weathering also requires a sufficient supply of fresh, weatherable rock at the surface, which plate tectonics helps to provide via orogeny and uplift. As a result, the coupling between plate tectonics and climate can behave as a negative feedback mechanism in some cases, where a cool climate promotes the operation of plate tectonics, and plate tectonics enhances silicate weathering such that the carbon cycle can sustain cool climate conditions. Figure 1: Flow chart representing the concept of whole planet coupling. Climate influences tectonics through the role of surface temperature in a planet's tectonic regime (i.e. stagnant lid versus plate tectonics), while the tectonic regime in turn affects climate through volatile cycling between the surface and interior. The tectonic regime also influences whether a magnetic field can be generated by dictating the core cooling rate. Finally, the strength of the magnetic field influences atmospheric escape, and therefore long-term climate evolution. An additional coupling comes into play via the core dynamo and the magnetic field. The magnetic field is generated by either thermal or chemical convection in the liquid iron core. Thermal convection requires a super-adiabatic heat flux out 4 Atmosphere*&*Climate*Tectonics*Magne5c*Field*Magne5cally*Limited*Escape*Cool*climate*&******vola5le****weakening*Vola5le*Recycling*Core*Cooling* of the core, which is controlled in part by the style of mantle convection, while chemical convection is driven by light element release during inner core nucleation, which also relies on cooling of the core. Plate tectonics cools the mantle efficiently by continuously subducting cold slabs into the deep interior, thus maintaining a high heat flow out of the core. The magnetic field can in turn limit atmospheric escape, helping retain liquid surface water. The coupling between plate tectonics and the core dynamo, and the magnetic field and the climate, completes the concept of whole planet coupling (Figure 1). Moreover the magnetic field and plate tectonics can also act as a negative feedback in cases where magnetic shielding is required to prevent rapid planetary water loss, and plate tectonics is needed to drive the dynamo. 1.3 Applications to the evolution of rocky exoplanets The concept of whole planet coupling discussed in this review is based on our knowl- edge of the Earth and the other rocky planets in the solar system, and is therefore limited to planets of a particular composition; i.e. planets made up mainly of silicate mantles and iron cores. It is unknown whether any of the processes discussed here are applicable to planets with more exotic compositions, such as planets primarily composed of carbides rather than silicates (e.g. Madhusudhan et al., 2012). Further- more we focus on H2O and CO2 as key volatiles; CO2 is an important greenhouse gas for stabilizing planetary climate, and water is crucial for driving the carbon cycle, may play a role in the operation of plate tectonics, and is thought to be a necessary ingredient for life. Thus planets must accrete a significant supply of both H2O and CO2 for whole planet coupling to be possible. However, volatile accretion is unlikely to be a problem as simulations indicate that planets commonly acquire large volatile inventories (i.e. Earth-like or larger) during late stage planet formation (Morbidelli et al., 2000; Raymond et al., 2004). Another important consideration is the redox state of the mantle, which deter- mines whether degassing via planetary magmatism releases H2O and CO2 to the atmosphere or reduced species such as H2 and CH4 (e.g. Kasting et al., 1993a). Car- bon dioxide is the only major greenhouse gas known to be regulated by negative feedbacks such that it has a stabilizing influence on climate. Thus having CO2 as a primary greenhouse gas is important for the whole planet coupling discussed here to operate. An oxidized upper mantle favors CO2 over CH4 and other reduced gases. Earth's mantle has been oxidized at present day levels since at least the early Archean (Delano, 2001), and possibly even since the Hadean (Trail et al., 2011). Oxidation of the mantle is thought to occur by disproportionation of FeO to Fe2O3-bearing per- ovskite and iron metal in the lower mantle during accretion and core formation. The 5 iron metal is then lost to the core leaving behind oxidized perovskite that mixes with the rest of the mantle (Frost et al., 2008; Frost & McCammon, 2008). Dispropor- tionation of FeO is expected to occur on rocky planets Earth sized or larger (Wade & Wood, 2005; Wood et al., 2006), so CO2 is likely to be an important greenhouse gas on exoplanets. Though other greenhouse gases can still be important, any planet where significant amounts of CO2 are degassed by mantle volcanism will need silicate weathering to act as a CO2 sink to avoid extremely hot climates. 1.4 Outline The paper is structured as follows. We review the basic physics behind the generation of plate tectonics from mantle convection, highlighting the role of climate, in §2. We then review the long-term carbon cycle and its influence on climate, and detail the specific ways plate tectonics is important for the operation of this cycle, in §3. Next we review magnetic field generation via a core dynamo and the physics of atmospheric shielding and volatile retention in §4. In §5 we integrate the discussion of the previous three sections to describe the range of different evolutionary scenarios that can result from differences in orbital distance, initial climate state, volatile inventory, and other factors as a consequence of whole planet coupling. We conclude by listing some major open questions that must be addressed in order to further our understanding of how terrestrial planets evolve and the geophysical factors that influence habitability (§6). 2 Generation of plate tectonics from mantle con- vection and the role of climate 2.1 General physics of plate generation Plate tectonics is the surface expression of convection in the Earth's mantle: the lithosphere, composed of the individual plates, is the cold upper thermal boundary layer of the convecting mantle, subducting slabs are convective downwellings, and plumes are convective upwellings (Davies, 1999; Bercovici et al., 2015). However, mantle convection does not always lead to plate tectonics, as evidenced by Mercury, Mars, and Venus, each of which is thought to have a convective mantle but lack plate tectonics (e.g. Breuer & Moore, 2007). On these planets, "stagnant lid convection," where the lithosphere acts as a rigid, immobile lid lying above the convecting mantle (e.g. Ogawa et al., 1991; Davaille & Jaupart, 1993; Solomatov, 1995), is thought to operate (Strom et al., 1975; Phillips et al., 1981; Solomon et al., 1992; Solomatov 6 & Moresi, 1996, 1997; Spohn et al., 2001; O'Neill et al., 2007), though Venus may experience occasional, short-lived episodes of subduction (Turcotte, 1993). Stagnant lid convection is a result of the temperature dependence of mantle viscosity, which increases by many orders of magnitude as temperature decreases from the hot con- ditions that prevail in the planetary interior to the much cooler temperatures that prevail at the surface (e.g. Karato & Wu, 1993; Hirth & Kohlstedt, 2003). When the viscosity at the surface is ∼ 103 − 104 times that of the interior, the top thermal boundary layer is so viscous that it can no longer sink under its own weight and form subduction zones, and stagnant lid convection ensues. (e.g. Richter et al., 1983; Christensen, 1984; Ogawa et al., 1991; Davaille & Jaupart, 1993; Moresi & Solo- matov, 1995; Solomatov, 1995). Laboratory experiments indicate that the viscosity ratio between the surface and mantle interior for typical terrestrial planets, includ- ing Earth, is ∼ 1010 − 1020, far larger than the threshold for stagnant lid convection. Stagnant lid convection is thus the "natural" state for rocky planets, and additional processes are necessary for plate tectonics to operate on Earth. In order to generate plate-tectonic style mantle convection, rheological complex- ities capable of forming weak, localized shear zones in the high viscosity lithosphere are necessary. These weak shear zones (essentially plate boundaries) remove the strong viscous resistance to surface motion that temperature-dependent viscosity causes and thus allow subduction and plate motions to occur. We discuss plate boundary formation in terms of viscous weakening in the lithosphere because the ma- jority of the lithosphere, including the region of peak strength in the mid-lithosphere, deforms via ductile or semibrittle/semiductile behavior; the lithosphere only "breaks" in a brittle fashion at shallow depths of < 10− 20 km (Bercovici et al., 2015). Many mechanisms have been proposed for producing the lithospheric weakening and shear localization necessary for plate-tectonic style mantle convection (see reviews by Tack- ley, 2000a; Bercovici, 2003; Bercovici & Karato, 2003; Regenauer-Lieb & Yuen, 2003; Bercovici et al., 2015); while we won't detail each of these mechanisms, they do all share some general principles. Namely, shear-thinning non-Newtonian rheologies, have been found to be successful at generating plate-like mantle convection. In a shear-thinning fluid, the strain-rate, ε, is a non-linear function of the stress, τ , (where ε and τ are the second invariants of the stress and strain-rate tensors, respectively): where n is a constant. As the effective viscosity, µef f , is proportional to τ ε−1, ε ∝ τ n µef f ∝ τ (1−n) ∝ ε (1−n) n . (1) (2) With n > 1, viscosity decreases as stress or strain-rate increases and high stress, 7 Figure 2: Illustration of plate generation using non-Newtonian (or effectively non- Newtonian) rheologies. Shown are a typical viscosity field (A) and temperature field (B) for stagnant lid convection and schematic strain-rate versus stress (C,E) and viscosity versus strain-rate curves (D,F) for viscoplastic (see §2.1.1) and n = 3 power law (equations (1) & (2)) rheologies. Weakening of the high viscosity litho- sphere present during stagnant lid convection is needed in order to allow subduction, and with non-Newtonian rheologies such weakening will occur at high stress regions (highlighted in grey in panels (A) and (B)). Panels (C) and (D) illustrate how non- Newtonian rheologies cause viscous weakening at high stresses or strain-rates, while panels (E) and (F) schematically demonstrate how outside conditions that influence the effectiveness of a given plate generation mechanism change the strain-rate versus stress curve (E) and the viscosity versus strain-rate curve (F). 8 0.20.40.60.80.20.40.60.80.51.01.52.02.53.03.50.51.01.52.02.53.03.5100101102103104105106µ0.20.40.60.80.20.40.60.80.51.01.52.02.53.03.50.51.01.52.02.53.03.50.20.40.60.80.20.40.60.80.51.01.52.02.53.03.50.51.01.52.02.53.03.50.000.250.500.751.00TTemperatureViscosityHigh stress regions of the lithosphere, above downwellings,undergo weakening with non-NewtonianrheologiesABLower yield stress ormore effective damage leads to larger degree of weakening0.000.250.500.751.001.251.50Stress0.00.51.01.52.0Strain−rateCNewtonian (n=1)Viscoplastic (τy = 0.75)Non−newtonian (n=3)0.10.20.512510Viscosity0.00.51.01.52.0Strain−rateD0.000.250.500.751.00Stress0.00.51.01.52.0Strain−rateE0.10.20.512510Viscosity0.00.51.01.52.0Strain−rateF high strain-rate regions of the lithosphere, such as regions of compression or tension caused by drips breaking off the base of the lithosphere, are weakened (see Figure 2). Thus lithospheric weakening occurs in exactly the regions necessary for facilitating subduction. Shear-thinning behavior can result from inherently non-Newtonian rhe- ologies with n > 1, or more complicated rheologies that involve "memory," where a time-evolving state variable, itself a function of stress and/or strain-rate, influences viscosity. In the case of memory, an effectively shear-thinning ε − τ relationship can be calculated by taking the state variable in steady-state (as we show for grainsize reduction in §2.1.2). Moreover, rheologies where stress decreases at high strain-rates, resulting in even stronger weakening, are possible but beyond the scope of this paper. 2.1.1 Viscoplasticity One popular plate generation mechanism is the viscoplastic rheology, where the litho- sphere "fails" and becomes mobilized when it's intrinsic strength, the yield stress, is reached (Fowler, 1993; Moresi & Solomatov, 1998). Typical implementations of the viscoplastic rheology assume that the mantle behaves like a Newtonian fluid (fol- lowing (1) & (2) with n = 1) until stress becomes equal to the yield stress, τy (see Figures 2C & D). At this point stress is independent of strain-rate, fixed at τy, and the viscosity follows µef f ∝ τy ε . (3) When convective stresses in the lithosphere hit the yield stress, low viscosity shear zones at convergent and divergent margins can form, and plate-like mantle convection can be obtained in two-dimensional (Moresi & Solomatov, 1998; Richards et al., 2001; Korenaga, 2010b), three-dimensional cartesian (Trompert & Hansen, 1998; Tackley, 2000b; Stein et al., 2004), and three-dimensional spherical convection models (van Heck & Tackley, 2008; Foley & Becker, 2009). Convective stresses lower than the yield stress do not form weak plate boundaries, and thus result in stagnant lid convection. The convective stress, τm, has typically been found to scale as τm ∝ µm εm (4) where µm and εm are the effective viscosity and strain-rate in the mantle, respectively (e.g. O'Neill et al., 2007). The strain-rate can be estimated as vm/d, where vm is convective velocity and d is the thickness of the mantle. Velocity scales as (e.g. Solomatov, 1995; Turcotte & Schubert, 2002) vm ∝ Ra (5) 2 3 , 9 where the Rayleigh number (Ra) is defined as Ra = ρgα∆T d3 κµm (6) and describes the convective vigor of the system (ρ is density, g gravity, α thermal expansivity, ∆T the non-adiabatic temperature drop across the mantle, and κ the thermal diffusivity). However, despite a large Rayleigh number on the order of 107 − 108, typical con- vective stresses are only ∼ 1−100 MPa (e.g. Tackley, 2000b; Solomatov, 2004; O'Neill & Lenardic, 2007), approximately an order of magnitude lower than the strength of the lithosphere based on laboratory experiments (e.g. Kohlstedt et al., 1995). Thus much work has focused on mechanisms for decreasing the lithospheric strength, usu- ally involving water (e.g. Tozer, 1985; Lenardic & Kaula, 1994; Regenauer-Lieb et al., 2001; van der Lee et al., 2008), though some recent models have looked at mecha- nisms for increasing convective stresses (Rolf & Tackley, 2011; Hoink et al., 2012). In §2.2 we describe how the possibility of water weakening links the operation of plate tectonics to climate. Moreover, climate also influences convective stresses, by altering ∆T , providing another important link between climate and tectonics. 2.1.2 Damage and grainsize reduction While viscoplasticity has shown some success at generating plate-like mantle con- vection, it fails to form localized strike-slip faults (Tackley, 2000b; van Heck & Tackley, 2008; Foley & Becker, 2009), a problem with all simple non-Newtonian rheologies (Bercovici, 1993). Furthermore viscoplasticity is an instantaneous rheol- ogy: when stresses fall below the yield stress weakened lithosphere instantly regains its strength. On Earth, however, dormant weak zones are known to persist over geo- logic timescales, and can serve as nucleation points for new subduction zones (Toth & Gurnis, 1998; Gurnis et al., 2000). An alternative mechanism, that allows for dormant weak zones (i.e. memory of past deformation) and is capable of generating strike-slip faults, is grainsize reduction (Bercovici & Karato, 2003; Bercovici & Ri- card, 2012, 2013, 2014). Grainsize reduction is commonly seen in field observations of exhumed lithospheric shear zones (White et al., 1980; Drury et al., 1991; Jin et al., 1998; Warren & Hirth, 2006; Skemer et al., 2010), and experiments show that vis- cosity is proportional to grainsize when deformation is dominated by diffusion creep or grain boundary sliding (e.g. Hirth & Kohlstedt, 2003). One major problem for a grainsize reduction plate generation mechanism, though, is that grainsize reduction occurs by dynamic recrystallization, which takes place 10 in the dislocation creep regime, while grainsize sensitive flow only occurs in the diffusion creep or grain boundary sliding regimes (e.g. Etheridge & Wilkie, 1979; De Bresser et al., 1998; Karato & Wu, 1993). Thus, a feedback mechanism, where deformation causes grainsize reduction, which weakens the material and leads to more deformation, is apparently not possible (e.g. De Bresser et al., 2001). Without such a feedback only modest weakening can occur. However, a new theory argues that when a secondary phase (e.g. pyroxene) is dispersed throughout the primary phase (e.g. olivine), a grainsize feedback is possible, because grainsize reduction can continue in the diffusion creep regime due the combined effects of damage to the interface between phases and Zener pinning (where the secondary phase blocks grain-growth of the primary phase) (Bercovici & Ricard, 2012). Using a simplified version of the theory in Bercovici & Ricard (2012) (see also Foley & Bercovici, 2014), the grainsize reduction mechanism can be formulated as: dA dt = fA γ Ψ − hAp (7) where A is the fineness (or inverse grainsize), γ is surface tension, h is the grain- growth (or healing) rate, p is a constant, Ψ is the deformational work, defined as Ψ = εijτij, where εij is the strain-rate tensor and τij is the stress tensor, and fA is the fraction of deformational work that partitions into surface energy, thus driving grainsize reduction. The grain-growth rate is strongly temperature dependent, and is given by where hn is a constant, Eh is the activation energy for grain-growth, Rg is the uni- versal gas constant, and T is temperature. The effective viscosity is µef f = µn exp = µ(T ) , (9) where µn and m are constants, Ev is the activation energy for diffusion creep, and A0 is the reference fineness. Taking (7) in steady-state, using the viscous constitutive law τij = 2µef f εij, and the definition εijεij = 2 ε, the fineness can be written as a function of strain-rate Combining (9) and (10), the viscosity is also a function of strain-rate, (cid:19) (cid:18)−Eh (cid:19)−m RgT h = hn exp (cid:18) Ev (cid:19)(cid:18) A RgT A0 (cid:18) A (cid:19)−m A0 (cid:18)4fAµ(T ) ε2Am (cid:32) (cid:18)4fA ε2 γh 0 µ(T )p p+m (cid:19) 1 (cid:19)−m(cid:33) 1 . p+m A = µef f = γhAp 0 11 (8) (10) (11) and using τ = 2µef f ε the effective τ − ε relationship is (cid:18) ε = 1 2 µ(T )−p (cid:18) fA (cid:19)m γhAp 0 (cid:19) 1 p−m τ p+m . (12) With typical parameters of m = 2 and p = 4 (Hirth & Kohlstedt, 2003; Bercovici & Ricard, 2012), the grainsize reduction mechanism results in an effectively non- Newtonian rheology with n = 3 (Figures 2C & D). Furthermore, as detailed in §2.2, climate can have a significant impact on plate generation with this mechanism, because grain-growth is temperature-dependent. At high surface temperatures, and thus higher temperatures in the mid-lithosphere, faster grain-growth (i.e. higher h) impedes grainsize reduction. 2.2 Influence of climate on tectonic regime Both the viscoplastic and grainsize reduction mechanisms for generating plate-tectonic style mantle convection are linked to climate. Although we treat these mechanisms separately, they are not mutually exclusive. The viscoplastic rheology is most rele- vant to brittle deformation in the upper lithosphere, while grainsize reduction pro- vides viscous weakening in the lower lithosphere. As a result both mechanisms can operate simultaneously, and both may be important for generating plate tectonics. Climate can influence wether plate-like convection occurs with a viscoplastic rhe- ology in three main ways: first, climate dictates whether water can exist in contact with ocean lithosphere at the surface, such that high pore pressure (e.g. Hubbert & Rubey, 1959; Sibson, 1977; Rice, 1992; Sleep & Blanpied, 1992) and weak, hydrous phases (Escart´ın et al., 2001; Hilairet et al., 2007) can lubricate faults and lower the lithosphere's yield strength; second, climate influences whether deep cracking of the lithosphere can occur, which is potentially important for hydrating the mid- to lower-lithosphere; and third, climate influences convective stresses. The first point, whether water can interact with rocks at the surface, provides only a weak influence of climate on tectonic regime, because hydrous phases can form over a wide range of surface temperatures. Geothermal heat flow can likely maintain a sub-ice ocean even at temperatures below the water freezing point (e.g. Warren et al., 2002), and hydrous silicates are stable at temperatures up to ≈ 500− 700◦ C (e.g. Hacker et al., 2003), meaning these phases could even form under a steam atmosphere. However, forming weak faults in the near surface environment is not sufficient for plate tectonics; weakening of the mid- and lower-lithosphere, where strength is maximum, is also necessary. The deep lithosphere is initially dry due to dehydration during mid-ocean ridge melting (Hirth & Kohlstedt, 1996; Evans et al., 2005), and is 12 deeper than hydrothermal circulation can reach (Gregory & Taylor, 1981). Thus, a mechanism capable of hydrating the mid-lithosphere is necessary for water weakening to be viable. One possible mechanism is the ingestion of water along deep thermal cracks (Korenaga, 2007). Deep cracking is a result of thermal stresses that arise in the cooling lithosphere. The thermal stresses are proportional to the temperature difference between the cold lithosphere and the mantle interior, so climate can di- rectly influence hydration of the mid-lithosphere. Higher surface temperatures could potentially lead to shallow cracks that leave the mid-lithosphere dry and strong. The surface temperature range where thermal cracking is effective has not been explored, but temperatures would likely have to be on the order of hundreds of degrees hotter than the present day Earth to impede plate tectonics. An even stronger role for climate stems from the influence of surface temperature on convective stresses. High surface temperatures decrease stresses by lowering the negative thermal buoyancy of the lithosphere (e.g. ∆T decreases, resulting in lower τm from equations (4)-(6)). If the convective stress drops below the yield stress, then stagnant lid convection ensues. Lenardic et al. (2008) found that increasing Earth's present day surface temperature by ∼ 100 K or more is sufficient to induce stagnant lid convection, even with a weak, hydrated lithosphere (see also Weller et al., 2015). When a grainsize reduction mechanism is considered, climate controls whether plate-like convection can occur through the influence of temperature on lithospheric grain-growth rates (Figure 3). Grain-growth is faster at higher temperatures (see (8)), and higher grain-growth rates impede the formation of weak shear zones by acting to increase grainsize (see (10)). Thus high surface temperatures, which lead to higher temperatures in the mid-lithosphere, promote stagnant lid convection (Lan- duyt & Bercovici, 2009; Foley et al., 2012). Foley et al. (2014) found that surface temperatures need to reach ≈ 500−600 K for an Earth-like planet to enter a stagnant lid regime due to enhanced grain-growth. Increasing surface temperature even fur- ther would eventually lead to a state where the lithosphere has a low enough viscosity to "subduct" even without any of the rheological weakening mechanisms discussed in this section (i.e. mantle convection would resemble constant viscosity convection). However, for this style of "subduction" to occur, the viscosity ratio between the sur- face and mantle interior must be less than ≈ 104 (Solomatov, 1995), which requires a surface temperature of > 1000 K with an Earth-like mantle temperature of ≈ 1600 K. Summarizing the results from both mechanisms, cold surface temperatures are generally favorable for plate tectonics because they promote grainsize reduction and boost convective stresses. Moreover, cold surface temperatures are unlikely to pre- vent water weakening because geothermal heat flow can maintain liquid water be- 13 Figure 3: Effective stress versus strain-rate relationship (A) and viscosity versus strain-rate relationship (B) resulting from grain-damage (equations (8)-(12) with Eh = 500 kJ mol−1, Ev = 300 kJ mol−1, m = 2, p = 4, and fA = 10−4). A lower temperature in the mid-lithosphere, which can result from a lower surface temperature, enhances viscous weakening, i.e. lower viscosities result from the same strain-rate. As a result, the plate tectonic regime is found at low surface temperatures while the stagnant lid regime is found at high temperatures, as shown by the regime diagram from Foley et al. (2012) (C, reprinted with permission). Note that the boundary between the plate-tectonic and stagnant lid regimes in (C) is normalized to the Earth's present day state. The exact position of the boundary between Earth and Venus could not be constrained by the simple models used in Foley et al. (2012). 14 200 300 400 500 600 700 800 0.5 1 1.5 2 2.5Surface Temperature [K]Planet Radius [Earth Radii]Stagnant LidRegimePlate TectonicRegimeVenusMarsEarthSuper-Earth?C0246810Stress [MPa]02e−154e−156e−158e−15Strain−rate [s−1]A10191020102110221023Viscosity [Pa s]02e−154e−156e−158e−15Strain−rate [s−1]Lithosphere Temp = 1050 KLithosphere Temp = 850 KB neath an ice covered ocean. However, high temperatures, in the range of 400-600 K, can cause stagnant lid convection by dropping convective stresses, increasing grain- growth rates, and possibly suppressing thermal cracking. The influence of climate on tectonic regime is also a leading a hypothesis for the lack of plate tectonics on Venus, as the Venusian surface temperature, 750 K, is easily hot enough to shut down plate tectonics based on the geodynamical models discussed in this section (Lenardic et al., 2008; Landuyt & Bercovici, 2009). 2.3 Importance of mantle temperature and other factors Naturally climate is not the only important factor governing whether terrestrial plan- ets will have plate tectonics. Mantle temperature, in particular, has been shown to have a major influence on a planet's tectonic regime. Mantle temperature modu- lates mantle convective stresses and lithospheric grain-growth rates, similar to the influence of surface temperature discussed in §2.2, and mantle temperature controls the thickness of the oceanic crust formed at ridges, which affects the lithosphere's buoyancy with respect to the underlying mantle. All of these effects act to inhibit plate tectonics when interior temperatures are high, implying that planets with high internal heating rates or young planets, which are hotter because they have had less time to lose their primordial heat (e.g. Abbott et al., 1994; Korenaga, 2006; Labrosse & Jaupart, 2007; Herzberg et al., 2010) (see also §4.2), will be less likely to have plate tectonics. However, there is disagreement over just how important mantle temperature is in dictating a planet's mode of surface tectonics. Increasing mantle temperature drops the mantle viscosity, µm. From equations (4)-(6), mantle stress scales as µ1/3 m . Thus higher mantle temperatures lead to lower convective stresses, and, if stresses drop below the yield strength, stagnant lid con- vection in viscoplastic models (O'Neill et al., 2007; Moore & Webb, 2013; Stein et al., 2013); this effect may have even caused stagnant lid convection, possibly with episodic subduction events, on the early Earth (O'Neill et al., 2007; Moore & Webb, 2013). The influence of mantle temperature on stress can also lead to hystere- sis in viscoplastic mantle convection models (Crowley & O'Connell, 2012; Weller & Lenardic, 2012; Weller et al., 2015). Stagnant lid convection results in higher interior temperatures, and thus lower stresses, than mobile lid convection, due to low heat flow across the thick, immobile lithosphere. It is therefore harder to initiate plate tectonics starting from a stagnant lid initial condition than it is to sustain plate tectonics on a planet where it is already in operation. Such hysteresis would also mean that a planet that loses plate tectonics will be unlikely to restart it at a later time, even before the couplings between climate and the magnetic field are taken 15 into account. High mantle temperatures also play a role in plate generation via grainsize reduc- tion, because they result in a warmer mid-lithosphere, and thus higher grain-growth rates. However, the influence of mantle temperature on lithospheric grain-growth rate is weaker than that of surface temperature, so elevated mantle temperatures alone are not capable of shutting down plate tectonics (Foley et al., 2014), in con- trast to viscoplasticity. Furthermore, the weak role of mantle temperature in plate generation means that the hysteresis loops seen in viscoplastic models may be less prominent with a grainsize reduction mechanism, though future study is needed to confirm this. Mantle temperature also determines the thickness of oceanic crust generated at spreading centers, with hotter temperatures generally leading to a thicker crust (e.g. White & McKenzie, 1989). As crust is less dense than the underlying mantle, cool- ing of the lithospheric mantle is necessary for the lithosphere as a whole to become convectively unstable (e.g. Oxburgh & Parmentier, 1977). A very thick crust then requires a long cooling time, such that a thick lithospheric mantle root can form, in order to create a negatively buoyant lithospheric column. Forming a thick litho- spheric mantle root is problematic, especially under high mantle temperature con- ditions, because the lithospheric mantle can delaminate before the lithosphere as a whole reaches convective instability, preventing subduction and plate tectonics (e.g. Davies, 1992). The ability of crustal buoyancy to preclude plate tectonics on both the early Earth (Davies, 1992; Vlaar, 1985; van Thienen et al., 2004b), and on exoplan- ets (Kite et al., 2009), has been discussed by many authors. However, the transition from basalt to compositionally dense eclogite, which occurs at shallow depths of ≈ 50 km on Earth (e.g. Hacker, 1996), allows episodic subduction events to occur despite a thick, buoyant crust in models of the early Earth (van Thienen et al., 2004a) and super-Earth exoplanets (O'Rourke & Korenaga, 2012). Moreover, even stable, mod- ern day plate tectonics can potentially still operate because the same melting process that creates the crust also dehydrates and stiffens the sub-crustal mantle; this stiff lithospheric mantle then resists delamination and allows thick, negatively buoyant lithosphere to form through conductive cooling (Korenaga, 2006). It is also important to point out that other factors, in particularly composition and size, can potentially have a major impact on whether plate tectonics can take place on a rocky planet. However, both the influence of size and bulk composition on a planet's tectonic regime are not well understood. There is still significant dis- agreement over whether increasing planet size makes plate tectonics more or less likely, owing mainly to uncertainties in lithospheric rheology, the pressure depen- dence of mantle viscosity and thermal conductivity, and the expected radiogenic 16 heating budget of exoplanets (see also §5.2). Likewise, little is known about the key material properties for planetary mantles with a significantly different composition than Earth. Understanding how these factors influence a planet's propensity for plate tectonics and overall evolution is a vital area of future research. 3 Climate regulation and the long-term carbon cycle A primary factor determining whether a planet's climate will be conducive to plate tectonics is the degree of greenhouse warming. Venus has ≈ 90 bar of CO2 in its atmosphere resulting in a surface temperature of ≈ 750 K that is unfavorable for plate tectonics, while Earth only has ≈ 4 × 10−4 bar of atmospheric CO2, and thus has a temperate climate that allows for plate tectonics. However, on Earth there is enough CO2, anywhere from ≈ 60− 200 bar (e.g. Sleep & Zahnle, 2001), locked in carbonate rocks at the surface and in the mantle to cause very hot surface temperatures of ≈ 400 − 500 K (Kasting & Ackerman, 1986). Thus a key for maintaining plate tectonics on a planet is the ability to prevent large quantities of CO2 from building up in the atmosphere. On planets possessing liquid water, like Earth, this can be accomplished by silicate weathering at the surface and on the seafloor. Crucially, the weathering rate is sensitive to climate, with higher temperatures leading to larger weathering rates (and thus higher CO2 drawdown rates), meaning that weathering acts as a negative feedback mechanism that works to maintain temperate climate conditions (Walker et al., 1981; Berner et al., 1983; Tajika & Matsui, 1990; Brady, 1991; Tajika & Matsui, 1992; Berner & Caldeira, 1997; Berner, 2004). Planets that lack liquid water, like Venus, have no mechanism for regulating atmospheric CO2. However, as we show in this section, simply possessing water does not guarantee that weathering can maintain climates within a range favorable for plate tectonics. A sufficient supply of fresh, weatherable rock at the surface is also needed. We further argue that plate tectonics enhances the supply of fresh rock to the surface, opening the possibility that plate tectonics and the long-term carbon cycle act as a self-sustaining feedback mechanism in some cases. 3.1 Modeling the long-term carbon cycle Figure 4 gives a simple illustration of the global carbon cycle where four main reser- voirs of carbon are considered (see Berner (2004) and Ridgwell & Zeebe (2005) for detailed reviews): the seafloor, the mantle, the atmosphere, and the ocean (the atmo- 17 Figure 4: Schematic diagram of the global carbon cycle after Foley (2015). sphere and ocean reservoirs are often combined because equilibration between these two reservoirs is essentially instantaneous on geologic timescales (e.g. Sleep & Zahnle, 2001)). Carbon dioxide in the atmosphere is consumed by weathering reactions with silicate rocks, producing calcium, magnesium, and bicarbonate ions that flow from groundwater into rivers, eventually draining into the ocean (see Gaillardet et al. (1999) for a global compilation of CO2 consumption via chemical weathering). Once in the ocean, these ions recombine to precipitate carbonates, leading to an overall net transfer of CO2 from the atmosphere to carbonate rocks on the seafloor. Hydrother- mal alteration of basalt can also act as a sink for CO2 dissolved in the oceans, and given rapid equilibration between the atmosphere and ocean, a sink for atmospheric CO2 as well (Staudigel et al., 1989; Alt & Teagle, 1999; Gillis & Coogan, 2011). Carbonates on the seafloor, both in the form of sediments and altered basalt, are subducted into the mantle at trenches. Here, a fraction of the carbon devolatilizes and returns to the atmosphere through arc volcanoes, with the remaining carbon being recycled to the deep mantle. To close the cycle, mantle carbon is degassed through mid-ocean ridge and plume volcanism back to the atmosphere and ocean reservoirs. The balance between weathering and volcanic outgassing dictates atmospheric 18 Mantle reservoirAtmosphere reservoirOcean reservoirPlate reservoirFvol :Arc volcanism & ridge volcanismSubduction of carbon into mantleSilicate weathering, FweatherHydrothermal alteration (seafloor weathering) CO2 content. This balance is formulated as Fweather + Fsf w = Farc + Fridge, (13) where Fweather is the silicate weathering flux on land, Fsf w is the seafloor weathering flux, Farc is the flux of CO2 degassing from volcanic arcs, and Fridge is the degassing flux at ridges. Balance between silicate weathering and degassing typically occurs rapidly, on a timescale of ∼ 1 Myr or less (e.g. Sundquist, 1991; Berner & Caldeira, 1997; Driscoll & Bercovici, 2013; Foley, 2015), so assuming that weathering always balances degassing is reasonable when studying long-term climate evolution. The arc degassing flux is typically written as Farc = f vpLRp Ap , (14) where f is the fraction of subducted carbon that degasses, vp is the plate speed, L is the length of subduction zone trenches, Rp is the amount of carbon on the seafloor, and Ap is the area of the seafloor. The ridge degassing flux is given by Fridge = fdvpLdmeltRman Vman , (15) where fd is the fraction of upwelling mantle that degasses, L is the length of ridges (assumed equal to the length of trenches), dmelt is the depth where mid-ocean ridge melting begins, Rman is the amount of carbon in the mantle, and Vman is the volume of the mantle (e.g. Tajika & Matsui, 1990, 1992; Sleep & Zahnle, 2001; Driscoll & Bercovici, 2013; Foley, 2015). Silicate weathering can exert a negative feedback on climate because mineral dis- solution rates, and hence CO2 drawdown rates, increase with increasing temperature and precipitation (e.g. Walker et al., 1981; Berner et al., 1983) (see Kump et al. (2000) and Brantley & Olsen (2014) for reviews on mineral dissolution kinetics and atmospheric CO2). A similar link between climate and weathering rate can be seen in many field studies (e.g. Velbel, 1993; White & Blum, 1995; Dessert et al., 2003; Gislason et al., 2009; White & Buss, 2014; Viers et al., 2014). However, many other locations show no link between weathering and climate; instead the weathering flux is linearly related to the physical erosion rate (Stallard & Edmond, 1983; Edmond et al., 1995; Gaillardet et al., 1999; Oliva et al., 1999; Millot et al., 2002; Dupr´e et al., 2003; Riebe et al., 2004; West et al., 2005; Viers et al., 2014). These conflicting field observations are likely due to the distinction between "kinetically limited" (or "re- action limited") weathering and "supply limited" weathering. When weathering is 19 kinetically limited, the kinetics of mineral dissolution controls the weathering rate. However, when weathering is supply limited, the supply of fresh rock brought to the weathering zone by erosion controls the weathering rate. Supply limited weathering occurs when the reaction between silicate minerals and CO2 runs to completion in the regolith, requiring physical erosion to expose fresh bedrock for weathering to continue. The supply limited weathering flux (Fws) can be written as (Riebe et al., 2004; West et al., 2005; Mills et al., 2011; Foley, 2015) Fws = AlandEχρcc ¯m , (16) where Aland is the surface area of all exposed land, E is the physical erosion rate, χ is the fraction of reactable elements in the crust, ρcc is the density of the crust, and ¯m is the molar mass of reactable elements. The kinetically limited weathering flux, Fwk, is sensitive to climate and can be written as (e.g. Walker et al., 1981; Berner et al., 1983; Tajika & Matsui, 1990, 1992; Berner, 1994; Sleep & Zahnle, 2001; Berner, 2004; Driscoll & Bercovici, 2013) (cid:20) Ea Rg (cid:18) 1 T ∗ − 1 T (cid:19)(cid:21)(cid:18)PCO2 (cid:19)β(cid:18) R (cid:19)α(cid:18) fland (cid:19) P ∗ CO2 R∗ f∗ land , (17) Fwk = F ∗ w exp where F ∗ w is the present day rate of atmospheric CO2 drawdown by silicate weathering w ≈ 6× 1012 mol yr−1, or half the estimate of Gaillardet et al. (1999) because half (F ∗ of the CO2 removed by weathering is re-released to the atmosphere when carbonates form (Berner, 2004)), Ea is the activation energy for mineral dissolution, Rg is the universal gas constant, T is temperature, PCO2 is the partial pressure of atmospheric CO2, R is the runoff, fland is the land fraction (subaerial land area divided by Earth's surface area), and β and α are constants. Stars represent present day values. The exponential term in (17) describes the temperature sensitivity of mineral dissolution rates, where typical values of Ea range from ≈ 40 − 50 kJ mol−1 (e.g. Brady, 1991). The PCO2 term describes the direct dependence of silicate weath- ering rates on atmospheric CO2 concentration, which results from the role of pH in mineral dissolution. Specifically, more acidic pH values cause faster dissolution rates. However, when dissolution is caused by organic acids produced by biology (mainly plants on the modern Earth), soil pH is fixed, and the direct dependence of weathering rate on PCO2 is weak (i.e. β in (17) is approximately 0.1 or less (Volk, 1987; Berner, 1994; Sleep & Zahnle, 2001)). Without biologically produced acids, carbonic acid formed when atmospheric CO2 dissolves in rainwater sets the soil pH, making the direct dependence of weathering rate on PCO2 stronger (β ≈ 0.5 (Berner, 1992)). The runoff term adds an additional climate feedback, as precipitation rates 20 increase with temperature (α ≈ 0.6 − 0.8 (Berner, 1994; Bluth & Kump, 1994; West et al., 2005); see Berner (1994) for a detailed discussion of how silicate weathering scales with runoff). Finally, the weathering rate scales with land fraction, because decreasing the total area of land undergoing weathering lowers the total weathering flux. The supply limited and kinetically limited weathering fluxes can be combined into a total weathering flux following Gabet & Mudd (2009), Hilley et al. (2010), and West (2012) (see Foley (2015) for a full derivation) as (cid:20) (cid:18) (cid:19)(cid:21) Fweather = Fws 1 − exp −Fwk Fws . (18) When Fwk is below the supply limit, Fws, the overall weathering rate is approximately equal to the kinetically limited weathering rate, i.e. Fweather ≈ Fwk. However, when Fwk reaches or exceeds the supply limit, the overall weathering rate is fixed at Fws; i.e. Fweather can no longer increase with increasing surface temperature or atmospheric CO2 content once the supply limit to weathering has been reached. The present day Earth consists of both regions undergoing supply limited weathering and those undergoing kinetically limited weathering. However, kinetically limited weathering appears to be the dominant component of the global silicate weathering flux (West et al., 2005). Finally, CO2 drawdown also occurs via seafloor weathering on mid-ocean ridge flanks (Alt & Teagle, 1999). Seafloor weathering is thought to be a smaller present day CO2 sink than weathering on land, and whether it exerts a strong feedback on climate is debated (Caldeira, 1995; Brady & G´ıslason, 1997; Berner, 2004; Coogan & Gillis, 2013; Coogan & Dosso, 2015). Many studies have assumed a weak dependence on PCO2, based on the experiments of Brady & G´ıslason (1997), and formulate the seafloor weathering flux as (e.g. Sleep & Zahnle, 2001; Mills et al., 2014; Foley, 2015) Fsf w = F ∗ sf w p , CO2 v∗ (19) P ∗ sf w ≈ 1.75 × 1012 mol yr−1 (Mills where the present day seafloor weathering flux F ∗ et al., 2014) and ξ ≈ 0.25. However, both a stronger PCO2 dependence or a direct temperature feedback are possible. Seafloor weathering can also become "supply limited," because only a finite amount of CO2 can be stored in the ocean crust. Sleep et al. (2001) and Sleep et al. (2014) estimate that only the top ≈ 500 m of ocean crust can be completely reacted with CO2, meaning no more than ≈ 4 × 1021 mol (or ≈ 30 bar), ≈ 1/10-1/2 Earth's total CO2 budget, can be locked in the seafloor at any one time. (cid:18) vp (cid:19)(cid:18)PCO2 (cid:19)ξ 21 Figure 5: Response of surface temperature (A) and partial pressure of atmospheric CO2, normalized by the present day value (B), to increases in the CO2 degassing rate (where Fvol/F ∗ vol is the volcanic outgassing rate normalized by the present day value). Different values of Ea and β are used, as described in the legend, to show how stronger climate feedbacks cause climate to be less sensitive to degassing rate. Typical values for the Earth are Ea = 42 kJ mol−1 and β = 0.5 (red line). 3.2 The silicate weathering climate feedback and plate tec- tonics Kinetically limited weathering is able to prevent massive quantities of CO2 from building up in the atmosphere, because the weathering rate increases with increasing atmospheric CO2 content. Considering a balance between a given degassing flux, Fvol = Farc + Fridge, and weathering, the atmospheric CO2 content as a function of degassing rate can be calculated (Figure 5). When the degassing rate is increased, only a small increase in atmospheric CO2 concentration is needed to bring weathering back into balance with degassing, because the weathering rate is a strong function of PCO2. The stronger the dependence of weathering on atmospheric CO2, the less sensitive PCO2 is to variations in the degassing rate. It is this ability to balance the volcanic outgassing flux with minimal changes in PCO2 that allows weathering to prevent extremely hot climates, and thus maintain conditions favorable for plate tectonics. The negative feedback between climate and weathering is also important for hab- itability, because it acts to stabilize climate in response to variations in solar lumi- nosity. Stars increase in luminosity as they age; for example the sun's luminosity 22 280290300310320330340350360370Surface Temperature [K]0102030405060708090100Fvol/F*volA050010001500200025003000350040004500PCO2/P*CO20102030405060708090100Fvol/F*volEa = 42 kJ/mol, β = 0Ea = 0 kJ/mol, β = 0.5Ea = 42 kJ/mol, β = 0.5Ea = 100 kJ/mol, β = 0.5B was 70 % its present day level during the Archean (Gough, 1981). The weathering feedback acts to partially counteract this change in luminosity by allowing higher CO2 levels when luminosity is low and lower CO2 levels when luminosity is high. However, this topic has been extensively studied (e.g. Walker et al., 1981; Franck et al., 1999; Sleep & Zahnle, 2001; Abbot et al., 2012), and going into more detail is beyond our scope. On the other hand, when weathering becomes supply limited, it can no longer increase with atmospheric CO2 level and is therefore unable to balance the degassing flux (globally supply limited weathering requires that Fvol ≥ Fws, otherwise weath- ering would not be supply limited). As a result, atmospheric CO2 accumulation from volcanic outgassing continues unabated until the mantle and plate reservoirs are depleted in carbon, and extremely hot climates, that are unfavorable for plate tectonics, prevail (Figure 6). Small land areas or low erosion rates lead to supply limited weathering, because both factors limit the supply of fresh rock to the surface, and hence lower Fws (see Foley, 2015, for details). Alternatively, factors that increase the degassing rate can also drive a planet into the supply limited weathering regime, even with a large land area or high erosion rates. Planets with large total CO2 inven- tories are more susceptible to supply limited weathering because degassing rates are higher. Another important factor is the fraction of subducted carbon that reaches the deep mantle, instead of devolatilizing and returning to the atmosphere at arcs (f from equation (14)). When more carbon can be subducted and stored in the mantle, Farc is lower and kinetically limited weathering is easier to maintain (Figure 6D). The fraction of subducted carbon that degasses at arcs is not well constrained, and the physical and chemical processes controlling this number are poorly understood (e.g. Kerrick & Connolly, 2001; Dasgupta & Hirschmann, 2010; Ague & Nicolescu, 2014; Kelemen & Manning, 2015). Moreover, f is likely a function of mantle tem- perature, as more slab CO2 will devolatilize during subduction into a hotter mantle (e.g. Dasgupta & Hirschmann, 2010). Thus planets with hot mantles, as expected for young planets or those with high radiogenic heating budgets (see §4.2), may be more susceptible to supply limited weathering. Better constraints on carbon subduction and devolatilization are clearly needed for understanding global climate feedbacks. Figure 6 assumes that seafloor weathering follows (19), as in Foley (2015), which provides only a modest climate feedback. However, including a stronger climate feedback does not prevent extremely hot climates from forming when weathering on land is supply limited, because seafloor weathering will also become supply limited when all of the accessible basalt is completely altered (Sleep et al., 2001, 2014). With the ocean crust unable to take in any more carbon, a CO2 rich atmosphere still forms (Foley, 2015). Furthermore, the high surface temperatures predicted for the supply 23 Figure 6: Partial pressure of atmospheric CO2 (A) and surface temperature (B) as a function of total planetary CO2 budget and land fraction (fland) after Foley (2015). Kinetically limited weathering occurs at large land fractions and low planetary CO2 budgets, resulting in temperate surface temperatures, while supply limited weather- ing results in extremely hot climates when land fraction is low or total CO2 budget is high. The estimated location of the Earth in land fraction-planetary CO2 budget space is shown in panel (B). The boundary between the kinetically limited and sup- ply limited weathering regimes is shown in panel (C) as a function of erosion rate, E, and in panel (D) as a function of the fraction of subducted carbon that degasses at arcs, f (where the label KL refers to the kinetically limited regime and SL refers to the supply limited regime). 24 -5-4-3-2-1log10(fland)10-210-1100101102103104105106107PCO2 [Pa]KineticallyLimitedSupply LimitedA-5-4-3-2-1log10(fland)21222324log10(Ctot [mol])250300350400450500550600650Surface Temperature [K]EarthBCE = 10 mm/yr E = 1 mm/yrE = 0.1 mm/yr21222324log10(Ctot [mol])DKLSLf = 0.95 f = 0.5f = 0.25 limited weathering regime could potentially lead to rapid water loss to space or even trigger a runaway greenhouse. Such scenarios are not explicitly modeled here, but would reinforce the fact that supply limited weathering causes a climate state that is unfavorable for plate tectonics; losing water would prevent volatile weakening of the lithosphere, and a runaway greenhouse climate results in temperatures even higher than those shown in Figure 6. 3.3 Extent of coupling between climate and surface tectonics Plate tectonics helps to sustain kinetically limited weathering, and therefore prevent extremely hot climates from forming, by acting to maintain high erosion rates and large subaerial land areas through orogeny and volcanism. Erosion rates are highest in rapidly uplifting, tectonically active areas (e.g. Portenga & Bierman, 2011), and orogenic processes are the primary cause of such uplift. Without tectonic uplift, erosion rates would decay to very low values because the surface can only erode as quickly as uplift creates topography. Orogeny results from plate tectonic processes, such as continent-continent collisions, island arc formation, and accretion of arcs to cratons, so plate tectonics is essential for keeping erosion rates on Earth high. Tectonic uplift and physical erosion have long been thought to be important for weathering and climate (e.g. Raymo & Ruddiman, 1992; Maher & Chamberlain, 2014), and the transition between kinetically limited and supply limited weathering provides an exciting new mechanism for explaining this influence (e.g. Kump & Arthur, 1997; Froelich & Misra, 2014). Volcanism also creates topography and supplies fresh rock to the surface, so the fact that plate tectonics causes widespread subaerial volcanism (e.g. through arc, plume, and other hot spot melting), is important for sustaining high erosion rates as well. Perhaps the more important role for volcanism, however, is in creating exposed land, as large land areas also increase the supply of weatherable rock. In particular, continental crust, which makes up the large majority of Earth's exposed land, is thought to be predominantly created by subduction zone volcanism (e.g. Rudnick, 1995; Cawood et al., 2013), a process that is unique to plate tectonics. Though hotspot or plume volcanism, the dominant mode of volcanism for a stagnant lid planet, can still create land and continental crust (e.g. Stein & Goldstein, 1996; Smithies et al., 2005), plate tectonics clearly enhances subaerial land. In fact, on a planet with a large area of exposed land, like Earth, the high erosion rates provided by plate tectonics may not be necessary for maintaining kinetically limited weathering. On Earth, typical erosion rates for flat-lying, tectonically inactive regions are on the order of 0.01 mm yr−1 (Portenga & Bierman, 2011; Willenbring 25 et al., 2013), which gives a supply-limited weathering flux on the order of 1013 mol yr−1 when combined with Earth's large land area of ≈ 1.5 × 1014 m2. However, as half of the CO2 removed by silicate weathering is re-released to the atmosphere when carbonates form (e.g. Berner et al., 1983), the net rate of CO2 drawdown is ∼ 5×1012 mol yr−1. The present day degassing flux is ≈ (6−10)×1012 mol yr−1 (e.g. Marty & Tolstikhin, 1998; Sleep & Zahnle, 2001; Burton et al., 2013), which would be significantly lower if Earth were in a stagnant lid regime. Marty & Tolstikhin (1998) estimate ≈ 3 × 1012 mol yr−1 of CO2 degassing from plumes, which is a reasonable approximation for Earth's degassing flux were it in a stagnant lid regime. As the stagnant lid degassing flux is lower than the supply-limited weathering flux, kinetically limited weathering is possible on a hypothetically stagnant lid Earth. Therefore plate tectonics and the long-term carbon cycle probably only act as a self-sustaining feedback on planets with small land areas or large planetary CO2 inventories, where high erosion rates are needed to prevent supply limited weather- ing (i.e. planets that would lie near the boundary between kinetically limited and supply limited weathering in Figure 6). Unfortunately a more quantitative estimate is not possible with our current level of understanding on how tectonics modulates erosion and weathering. Nevertheless, given that volatile acquisition during terres- trial planet formation can vary significantly (e.g. Raymond et al., 2004), volatile rich planets where plate tectonics and the carbon cycle are so tightly coupled may be common. Furthermore, even if a planet's exposed land area is large enough to sustain kinetically limited weathering at low erosion rates, plate tectonics may still be important. Plate tectonics may be responsible for creating most of the exposed land through continental crust formation, without which much higher erosion rates would be needed for kinetically limited weathering. Earth may even be an example of such a planet. Another aspect of the coupling between plate tectonics and climate is that plate tectonics leads to long-lived CO2 degassing by recycling carbon into the mantle at subduction zones. This process is important for habitability, because when CO2 degassing rates are low, or cease entirely, snowball climates can form (Kadoya & Tajika, 2014). However, snowball climates are unlikely to also shut down plate tectonics, so the role of plate tectonics in sustaining CO2 degassing probably does not represent a self-sustaining feedback between tectonics and climate. The different evolutionary scenarios that result from coupling between climate and surface tectonics are described in §5. In particular planets at different orbital distances can undergo different evolutions if one planet lies inward of the habitable zone's inner edge and thus lacks silicate weathering. Likewise different initial cli- mate conditions can cause divergent evolutions if an initially hot climate prevents 26 plate tectonics and kinetically limited weathering, or even weathering at all, from transpiring. 4 Generation of the magnetic field and its role in atmospheric escape We have described the interactions between the surface environment (atmosphere plus ocean) and the mantle in sections 2 & 3. In this section we demonstrate that additional interactions exists between the mantle and core, and the geomagnetic field (which is generated by the core dynamo) and atmosphere. The mantle controls the rate at which the core cools, thereby playing a crucial role in maintaining the energy flow necessary to drive convection and dynamo action in the core. The geomagnetic field provides a shield that holds the solar wind far above the surface (presently at about 9 Earth-radii) so that most high energy particles are diverted and prevented from disrupting the near surface environment. As a result, magnetic fields may limit the atmospheric escape rate under certain conditions. The magnetic field's influence on escape rate can then have an important control on long-term climate evolution, opening the possibility for an indirect influence of the core dynamo on mantle convection, in addition to the direct role mantle convection plays in driving the dynamo. 4.1 Core dynamo Earth has maintained an internally generated planetary magnetic field through con- vective dynamo action in its core over much of its history (Tarduno et al., 2015). The presence of a long-lived magnetic field is another unique feature of our planet, as only Mercury and Ganymede also have active magnetic fields today among the terrestrial planets and moons of the solar system (Schubert & Soderlund, 2011). Internally generated magnetic fields are created by convective dynamo action in a large rotating volume of electrically conductive liquid. In the terrestrial planets this typically occurs in an iron core. The most commonly cited process for driv- ing dynamo action, and thus the focus of this section, is thermal or compositional convection. However, other driving mechanisms, such as gravitational tides (e.g. C´ebron & Hollerbach, 2014) or precipitation of Mg initially dissolved in the liquid iron (O'Rourke & Stevenson, 2016) are possible. Thermal convection occurs when the core heat flow, Qcmb, exceeds the heat con- ducted adiabatically through the core. Recent estimates place the core's conductive 27 heat flow at 13-15 TW (Driscoll & Bercovici, 2014; Pozzo et al., 2014). A high core heat flow is also inferred from revisions to the dynamics of mantle plumes, which are thought to be generated at the core-mantle boundary (CMB) and be indicative of the CMB heat flow (Bunge, 2005; Zhong, 2006). Even if the total core heat flow is less than that needed to keep the core adiabatically well mixed, compositional buoyancy may still be able to drive convection. Compositional convection, which can be driven a number of ways, usually involves a phase change in the liquid where a density gradient develops or by the dissolution of an incompatible element due to a change in the solubility. For example, as Earth's core cools the inner core crystallizes from below, releasing buoyant light element rich fluid into the surrounding iron-rich fluid and driving compositional convection. In this example core heat flow is still important for driving compositional convection, because core cooling is necessary for inner core growth. The heat transferred across the core-mantle boundary, Qcmb, is controlled by the temperature gradient in the viscous mantle above: Qcmb = AcmbkLM ∆TLM δLM (20) where Acmb is CMB area, and kLM , ∆TLM , and δLM are the thermal conductivity, temperature drop, and thickness of the lower mantle thermal boundary layer. The temperature drop, ∆TLM = Tcmb− Tm, measures the thermal disequilibrium between the mantle and core. If the mantle cools efficiently then ∆TLM is large, resulting in a high Qcmb. The thermal boundary layer thickness, δLM , can be derived by assuming the boundary layer Rayleigh number (i.e. ρgα∆TLM δ3 LM /(κµLM ), where µLM is the effective viscosity of the lower mantle boundary layer) is at the critical value for convection. Compositional differences between the lower mantle and the bulk mantle (e.g. Sleep, 1988) can influence the thickness of δLM ; this influence can be incorporated by modifying µLM (Driscoll & Bercovici, 2014) or the critical Rayleigh number (Stevenson et al., 1983). Calculating δLM from the boundary layer Rayleigh number implies that a hotter mantle produces a thinner boundary layer, increasing Qcmb. Therefore, the ratio ∆TLM /δLM in (20) implies that the core and mantle will evolve towards thermal equilibrium faster in a hotter mantle and that efficient mantle cooling leads to efficient core cooling. However, cooling the core too fast can be detrimental to dynamo action in two ways: (1) rapid cooling can lead to complete solidification of the core, preventing fluid motion and dynamo action, and (2) rapid cooling can bring the core into thermal equilibrium with the mantle, which will eventually decrease the cooling rate below the threshold for driving convection. The later is the eventual fate of every rocky 28 body, but the longer a planet can avoid thermal equilibrium and maintain moderate cooling the longer it will generate a magnetic field. 4.2 Coupling between core dynamo and tectonic mode As the core cooling rate is determined by the mantle, the surface tectonic mode, which determines the mantle cooling rate, dramatically influences the dynamo. A mobile lid can accommodate a larger surface heat flow by exposing the mantle's top thermal boundary layer to the surface, while a stagnant lid inhibits cooling by insulating the mantle with a thick conductive layer. To model the cooling and convective power available to drive dynamo action over time, the energy balance of the core and mantle are solved simultaneously. Volume averaged temperature evolution can be derived using the secular cooling equation Qi = −ciMi Ti, where c is specific heat and i refers to either mantle (i = m) or core (i = c) in the mantle and core energy balances (e.g. Driscoll & Bercovici, 2014). Solving for Tm and Tc in terms of sources and sinks gives the mantle and core thermal evolution equations, Tm = Tc = − (Qcmb + Qrad − Qconv − Qmelt) Mmcm (Qcmb − Qrad,c) Mccc − Aicρicc dRic dTcmb (LF e + EG) (21) (22) where Qconv is heat conducted through the lithospheric thermal boundary layer by mantle convection (in W), Qmelt is heat loss by mantle melt eruption, and Qrad and Qrad,c are radiogenic heat production in the mantle and core, respectively. Crustal heat sources are excluded because they do not contribute to the mantle heat budget. The denominator of (22) is the sum of core specific heat and heat released by inner core growth, where Aic is inner core surface area, ρic is inner core density, c is a constant that relates average core temperature to CMB temperature, dRic/dTcmb is the rate of inner core growth as a function of CMB temperature, and LF e and EG are the latent and gravitational energy released at the ICB per unit mass. Detailed expressions for heat flows and temperature profiles as functions of mantle and core properties are given in Driscoll & Bercovici (2014) and Driscoll & Barnes (2015). Heat is lost from the mantle via conduction through the upper mantle thermal boundary layer (Qconv) and by melt eruption (Qmelt). The rate at which this heat is lost is a complex function of mantle temperature, material properties, and style of convection. A mobile lid implies that the upper mantle thermal boundary layer 29 reaches the planetary surface, so that the convective heat loss is controlled by con- duction through this thin boundary layer. Assuming the boundary layer is at the critical Rayleigh number for convection, the convective heat flow is Qconv = amν−β m T β+1 m , (23) where νm = µm/ρ is temperature-dependent mantle kinematic viscosity and am = 8.4×1010 W(m2s−1K−4)1/3 and β = 1/3 are constants (see Driscoll & Bercovici, 2014, equation 43). In steady state, a stagnant lid has a thicker conductive boundary layer, which can be modeled by decreasing am by a factor of ∼ 25 (Solomatov, 1995). Mobile lid convection therefore favors dynamo action because it efficiently cools the mantle and thus boosts the core heat flow. Stagnant lid convection, on the other hand, can impede core cooling; stagnant lid convection on Venus is a leading hy- pothesis for why Venus lacks a magnetic field (Nimmo, 2002). However, stagnant lid convection does not always prevent a core dynamo, as thermal history models commonly show a relatively short period of rapid cooling, regardless of the surface tectonic mode, during which core cooling rates are high and dynamo action is possi- ble. This period of rapid cooling, or thermal adjustment period, results from initially hot interior temperatures that are a consequence of planetary accretion. The thermal adjustment period typically lasts for the first 1− 2 Gyr as the internal temperatures and heat flows adjust to boundary conditions and heat sources. Similar early thermal dynamos have been proposed for the Moon (Stegman et al., 2003) and Mars (Nimmo & Stevenson, 2000). Figure 7 shows example thermal histories for Earth and Venus, where the Earth model uses a mobile lid heat flow scaling while the Venus model uses a stagnant lid scaling (see Driscoll & Bercovici, 2014). The initial mid-mantle and CMB tem- peratures are assumed to be at the silicate liquidus, which is a reasonable starting temperature following the last large accretion event. Both models experience an initial thermal adjustment for 1− 2 Gyr, but diverge soon after. After the initial ad- justment the Earth model cools monotonically, maintaining a thermal dynamo and later a thermo-chemical dynamo after inner core nucleation around 4 Gyr, which can be seen as a jump in the predicted magnetic moment (Figure 7B). The core of the Venus model cools slightly during the thermal adjustment period, driving a transient thermal dynamo, but is slowly heated as the mantle and core heat up due to radiogenic heat trapped beneath the stagnant lid. In this case the core is too hot to solidify, precluding a compositional dynamo. These models predict Venus main- tained a magnetic field for about 1.5-4 Gyr (Driscoll & Bercovici, 2014). During the thermal adjustment period the core and mantle are still strongly coupled, but the cooling rate is less sensitive to the style of mantle convection. Thus the primary role 30 of plate tectonics is to extend the life of the dynamo beyond this thermal adjustment period. In addition to normal convective cooling, volcanic heat loss can potentially extend core cooling beyond the thermal adjustment period on a stagnant lid planet, provided most of the melt can reach the surface of the planet and cool (Morgan & Phillips, 1983; Nakagawa & Tackley, 2012; Armann & Tackley, 2012; Driscoll & Bercovici, 2014). Obviously, magmatic heat loss was not efficient enough for Venus, Mars, or the Moon to sustain dynamos, so we are left with the tentative conclusion that a stagnant lid reduces both the long-term convective and volcanic mantle heat loss. Figure 7: Comparison of the predicted thermal (A) and magnetic (B) histories of Earth (solid) and Venus (dashed). The only difference between the models is the Earth model uses a mobile lid heat flow scaling while the Venus model uses a stagnant lid scaling (see (20)-(23); Driscoll & Bercovici (2014) contains additional details). The role of mantle dynamics in dictating magnetic field strength also indirectly links the core to climate. Climate influences the tectonic regime of a planet, and the tectonic regime dictates the core cooling rate through time and thus whether a long- lived magnetic field can be maintained. Therefore a cool climate, being favorable for plate tectonics, will also be favorable for long-term magnetic field generation, and hot climates, to the extent that they lead to stagnant lid convection, will be unfavorable for the dynamo. An important question is then whether the core can 31 exert any influence on climate through the role the magnetic field plays in limiting atmospheric escape. We address this topic next. 4.3 Magnetic limited escape In describing the escape of a planetary atmosphere it is helpful to think of the lim- iting escape mechanism rather than the specific physical escape process, as several escape processes typically occur concurrently with a single limiting bottleneck. The escape limit is typically characterized as being in either the diffusion or energy lim- ited regime, with no connection to the presence of a magnetic field. However, it is commonly argued that magnetic field strength should play an important role in atmospheric escape (Michel, 1971; Chassefiere, 1997; Lundin et al., 2007; Dehant et al., 2007; Lammer et al., 2012; Owen & Adams, 2014; Brain et al., 2014). Below we speculate about how a magnetic field could influence escape. The hydrodynamic and diffusion limits to escape have been heavily studied (e.g. Hunten, 1973; Watson et al., 1981; Shizgal & Arkos, 1996; Lammer et al., 2008; Luger & Barnes, 2015). Both of these escape limits depend on hydrogen number density. In the diffusion limit escape is linearly proportional to hydrogen mixing ratio at the tropopause (Hunten, 1973), while hydrodynamic (or energy) limited escape has a weaker dependence on hydrogen mixing ratio but is expected to occur at much higher H concentrations (Watson et al., 1981). Hydrodynamic escape is fundamentally driven by incident energy absorbed at the base of the escaping region, but a bottleneck occurs as the escape rate grows and the absorbing species are lost, thereby limiting absorption and the energy available to drive escape. Regardless of their detailed dependences, diffusion and energy limited escape are expected to intersect at some H mixing ratio (therefore denoting a transition from one mechanism to the other) as depicted in Figure 8. If a strong planetary magnetic field is present that balances the solar wind far from the top of the atmosphere, then the density of species exposed to the flow will be limited. In other words, the rate at which ionized planetary species are swept away by the stellar wind will decrease with increasing magnetopause distance and magnetic field strength. This effect adds a third limiting escape regime between the diffusion and hydrodynamic regimes. Figure 8 illustrates this scenario, where increas- ing hydrogen concentration leads to a transition from diffusion to magnetic limited, and then from magnetic to hydrodynamic limited escape. Driscoll & Bercovici (2013) propose that the limiting physical mechanism in the magnetic regime could be the removal of planetary ions from the magnetopause via Kelvin-Helmholtz instabilities, the occurrence of which are well documented (e.g. Wolff et al., 1980; Barabash et al., 32 Figure 8: Schematic diagram of atmospheric limiting escape regimes. Hydrogen escape rate versus mixing ratio with the diffusive (black), magnetic (blue), and hy- drodynamic (black) regimes. The transition from diffusive to hydrodynamic escape may be interrupted by magnetically limited escape if the planetary magnetic field is strong. See Driscoll & Bercovici (2013) for more details. 2007; Taylor & Lavraud, 2008). Mathematically the magnetically limiting escape rate is predicted to be a function of hydrogen ion concentration, magnetopause sur- face area, and instability time scale, such that stronger magnetic fields expose a lower density of planetary species to the solar wind because their number density decreases faster than the magnetopause surface area (Driscoll & Bercovici, 2013). Weaker (or non-existent) planetary magnetic fields will allow the solar wind to inter- act with the "top" of the atmosphere (where escape occurs), potentially pushing the magnetic limit above the diffusion and hydrodynamic limits, rendering the magnetic limit irrelevant (Figure 8). 33 4.4 Extent of coupling between mantle, core, and climate The magnetic field is therefore most important for atmospheric evolution during the transition from hydrodynamic to diffusion limited escape. A strong magnetic field could reduce the escape rate during this transition, thereby helping to preserve a planet's water budget. Without magnetic shielding significant water loss could occur, potentially preventing the silicate weathering feedback from acting to stabilize climate and, in turn, plate tectonics from operating. Whether plate tectonics is itself necessary for magnetic shielding during the transition to diffusion limited escape depends on the timing of this transition (Figure 9). If the transition to the diffusion regime occurs quickly, i.e. fH decreases rapidly, then it would occur during a planet's thermal adjustment period, when dynamo action is possible without plate tectonics (see §4.2). On the contrary, if the magnetic limited escape regime is occupied longer than the mantle thermal adjustment period (e.g. fH decreases slowly) then plate tectonics is probably important for preventing significant water loss through magnetic shielding. Unfortunately, the timing of the transition to diffusion limited escape is not well known, as fH depends on the details of a planet's atmospheric structure. Future work in this area is needed to determine how long rocky planets occupy the magnetic limited escape regime, and the conditions under which plate tectonics is necessary for water retention. The timing of the transition to diffusion limited escape is also affected by stellar wind strength. In particular, magnetic shielding may be important over a wider range of stratospheric hydrogen mixing ratios, and thus over a longer period of time, for planets exposed to stellar winds much stronger than those at Earth today. Strong stellar winds are likely for planets orbiting active solar mass stars or orbiting close to moderately active small mass stars. In these cases maintaining a magnetic field may be crucial for preserving liquid surface water over billion year timescales, and such a long lived dynamo likely requires plate tectonics. In fact the magnetic field, climate, and plate tectonics can act as a self-sustaining feedback in this case, where the magnetic field is required to prevent water loss, water is necessary for silicate weathering to keep the climate cool enough for plate tectonics, and plate tectonics is required to drive the dynamo. Our knowledge of how planetary magnetic fields influence atmospheric escape is still in its infancy with many unanswered questions. In fact a strong magnetic field may even enhance escape in some instances, by producing a larger interaction cross section with the solar wind, which can concentrate incident energy flux by a factor of 10 − 100 (Brain et al., 2014). Clearly a more thorough understanding of how the properties of the atmosphere, stellar wind, and magnetic field influence escape rates is needed to constrain the coupling between the core dynamo and climate. 34 Initially the dynamo Figure 9: Schematic of magnetic field intensity over time. is maintained by cooling during the thermal adjustment period (solid black) and tectonic mode is not important. This continues until core cooling begins to slow and the dynamo will either die without efficient mantle cooling (dashed black), or stay strong with efficient interior cooling associated with plate tectonics (solid black). If the timing of this event occurs after the transition from magnetic to diffusion limited escape (Scenario #1, vertical dashed red) then the shielding of the atmosphere is not dependent on the tectonic mode. If the dynamo divergence event occurs before the escape transition (Scenario #2, vertical dashed red) then the shielding of the atmosphere is coupled to the tectonic mode. 35 5 Whole planet coupling and planetary evolution Sections 2, 3, and 4 together outline whole planet coupling between climate, plate tec- tonics, and the magnetic field. In this section we describe how whole planet coupling can potentially explain the Earth-Venus dichotomy, and lead to a number different evolutionary scenarios for rocky exoplanets, many of which would be unfavorable for life. In particular we focus on habitable zone planets, showing how events early in a planet's history, such as the initial atmospheric composition and the timing of the initiation of plate tectonics, play a major role in determining whether long-term habitable conditions can develop. A planet's volatile inventory is likely important as well. Given our incomplete understanding of plate tectonics, the carbon cycle, the geodynamo, and how they interact, our discussion is mostly qualitative, and only represents some initial hypotheses for the factors governing planetary evolution. 5.1 The Earth-Venus dichotomy Figures 10 and 11 illustrate how we propose that climate, mantle, and the core in- teract on Earth and Venus. Venus, being inward of the inner edge of the habitable zone, can not have liquid water at its surface. As a result, silicate weathering can not draw CO2 out of the atmosphere, so a hot, CO2 rich climate forms via volcanic degassing and/or primordial degassing during accretion and magma ocean solidifica- tion. The hot climate prevents plate tectonics, and in turn the long-lived operation of a core dynamo due to a low core heat flux (Figure 11). For the Earth, being in the habitable zone allows liquid water, and in turn silicate weathering. Thanks to a temperate climate plate tectonics can operate, and the resulting high core heat flow powers the geodynamo (Figure 10). Jellinek & Jackson (2015) invoke a similar series of couplings to explain the Earth-Venus dichotomy, though they argue that the lack of plate tectonics on Venus is caused by high rates of radiogenic heating in the mantle, and that Venus' hot climate stems from the lack of plate tectonics. Their hypothesis contrasts with our interpretation, that Venus' orbital position is the key factor explaining its evolution. If surface-interior coupling is responsible for the divergent evolution of Earth and Venus, the runaway greenhouse climate, loss of plate tectonics (if it ever existed), and death of the magnetic field should all be correlated in time. Thus, determining the climate, magnetic, and tectonic history of Venus is a vital test. Unfortunately our current knowledge of Venusian history is poor. The trace amount of water in the atmosphere requires a minimum of about 500 Myr for the H to escape to space (Donahue, 1999), so the runaway greenhouse must have occurred by at least 36 Figure 10: Flow chart of climate-tectonic-magnetic coupling for an Earth-like planet. Arrows between reservoirs indicate volatile (blue) and heat (red) fluxes, their width roughly in proportion to magnitude. Black lines indicate conceptual couplings, such as the influence of magnetic field strength on escape rate or chemical weathering on climate. ≈ 0.5 − 1 Ga, but could have taken place much earlier. In fact, Venus may have entered a runaway greenhouse and lost its water during formation (Hamano et al., 2013). The Venusian tectonic and magnetic histories are similarly uncertain. There is evidence for a massive resurfacing event at 0.5-1 Ga. Surface features related to this event are most consistent with volcanism and lava flows rather than subduction of old crust and creation of new crust at ridges (Smrekar et al., 2013; Ivanov & Head, 2013). The planet may still be volcanically "active" today (Smrekar et al., 2010), but 37 Atmosphere*Ocean*Oceanic*Crust*Mantle*Con4nental*Crust*Mobile*Lid*Core*Sub9Arc*Lithosphere*Cool*Climate*Magne4c*Field*Escape'Chemical'Weathering'Dissolu5on'Equilibrium'Runoff'Carbonate'Forma5on'Subduc5on'Arc'Degassing'Mid?Ocean'Ridge'Hot'Spot'Deep'Subduc5on' Figure 11: Flow chart of climate-tectonic-magnetic coupling on a Venus-like planet. eruptions are likely sporadic. The style of tectonics before the resurfacing event is unknown, with stagnant lid convection, episodic overturns, or even Earth-like plate tectonics all possibilities. Unraveling the Venusian magnetic history is challenging because the preservation and in-situ measurement of any remanent magnetization on Venus' hot surface is unlikely. However, another possible line of evidence that may imply the presence of a paleo-Venusian magnetic field would be the loss of H+, He+, and O+ along polar magnetic field lines to the solar wind, a process known on Earth as the polar wind (Moore & Horwitz, 2007). This ion escape mechanism relies on the presence of an internally generated magnetic field, and could potentially leave a chemical fingerprint in the Venusian atmosphere (Brain et al., 2014). However our present day knowledge provides no evidence for a strong, internally generated magnetic field on Venus. Future exploration of Venus is needed to place tighter constraints on its evolution. 5.2 Evolution of rocky exoplanets Venus demonstrates one likely evolutionary scenario for planets lying inward of the habitable zone's inner edge. However, cooler climates, that still allow plate tectonics and a magnetic field, are also possible for such planets if they have a significantly smaller CO2 inventory than Venus. In this case a temperate climate can still exist, even without liquid water and silicate weathering to regulate atmospheric CO2 levels, 38 Atmosphere*Basal.c*Crust*Mantle*Core*Hot$Climate$Magne.c$Field$Escape$Hot$Spot$Stagnant*Lid* simply because there isn't enough CO2 to cause extreme greenhouse warming. In some cases a planet that experiences a runaway greenhouse could even retain some water at the poles, and potentially remain habitable (Kodama et al., 2015). Planets lying beyond the habitable zone's outer edge will likely be cold, and thus can plausibly sustain plate tectonics and a magnetic field as well. However, whether complex life can develop on any of these planets is unclear. Likewise, an Earth-like evolution is one likely scenario for planets lying within the habitable zone, but a number of factors could lead to a different evolution that is unsuitable for life. Hot, CO2 rich climates are expected after planet formation due to degassing during planetesimal accretion and magma ocean solidification (Abe & Matsui, 1985; Zahnle et al., 2007; Elkins-Tanton, 2008). If a planet's initial climate is so hot that liquid water is not stable (i.e. temperature and pressure conditions are beyond the critical point for water), then developing a temperate climate is probably not possible. Initial surface temperature and pressure conditions in excess of the critical point imply that silicate weathering would not occur, with or without plate tectonics (though some limited reaction between atmospheric CO2 and the crust is possible). As a result the climate would remain extremely hot, preventing both plate tectonics and a core dynamo. However, a climate hot enough to exceed the water critical point is an extreme case, requiring ≈ 500 bar of CO2 for a planet with an atmosphere composed of CO2 and H2O and receiving the same insolation as the Hadean Earth (Lebrun et al., 2013). Earth's total planetary CO2 budget is estimated at ≈ 100 − 200 bar, so even with complete degassing during accretion or magma ocean solidification liquid water was still stable (e.g. Sleep & Zahnle, 2001; Zahnle et al., 2007). Thus a planet would need a significantly larger total CO2 budget than Earth for initial atmospheric makeup to exceed the liquid water critical point. Another possibility is a climate where liquid water is stable, but is still too hot for plate tectonics. In this case low land fractions or erosion rates can potentially prevent plate tectonics, and a cool climate, from ever developing. High rates of volcanism would be needed to avoid this fate, by creating a sufficient supply of fresh rock at the surface for silicate weathering to cool the climate. Finally, even when climate conditions are amenable to plate tectonics, an uninhabitable state can be reached if plate tectonics does not initiate before increasing insolation warms the climate to the point where it is no longer possible (Figure 12). Before plate tectonics initiates on a planet, weathering may be supply limited and the atmosphere CO2 rich. Thus surface temperature will increase with increasing luminosity, and can potentially become hot enough to preclude plate tectonics from ever starting. If plate tectonics does not initiate before this divergence point is reached, a planet could become permanently stuck with a hot, uninhabitable climate, stagnant lid 39 Figure 12: Schematic diagram of the divergence point in planetary evolution involv- ing the initiation of plate tectonics. Before plate tectonics a planet's weathering rate may be supply limited, such that surface temperature climbs over time as a result of increasing luminosity. Once plate tectonics initiates, silicate weathering is enhanced by higher erosion rates and continent formation, and climate cools. convection in the mantle, and no protective magnetic field. Conversely on a planet where plate tectonics does initiate before reaching this divergence point, continental growth and orogeny will increase both land area and erosion rates, enhancing the ability of silicate weathering to establish and maintain a temperate, habitable climate. If a large enough land area forms, weathering may even be capable of maintaining a temperate climate without plate tectonics to ele- vate erosion rates. Likewise plate tectonics means that long-lived dynamo action, and therefore volatile shielding from the solar wind, is possible. The divergence point in- volving the initiation of plate tectonics could be particularly important if high mantle temperatures are a significant impediment to plate tectonics through low convective stresses or the creation of thick buoyant crust (see §2.3). Mantle temperature may need to cool before plate tectonics can begin, even if surface temperatures are not hot enough to preclude plate motions. However, the increase in luminosity during a star's main sequence evolution is relatively gradual (the sun's luminosity was only ≈ 30 % lower 4.5 Gyrs ago (Gough, 1981)), so there is a large time window, on the order of 1 Gyr, for sufficient mantle cooling to occur before climate becomes too hot for plate tectonics. Another divergence point involves magnetic shielding and planetary water loss. 40 TimeSurface TemperatureToo hot for plate tectonicsPlate-tectonic evolutionStagnant lidevolutionInitiation of plate tectonics The magnetic field can limit H escape, and thus help preserve surface water, when atmospheric escape transitions from being hydrodynamically limited to diffusion lim- ited (see §4.3 and §4.4.) If no magnetic field is available to shield the solar wind then massive amounts of H may be lost, leaving the planet desiccated and unable to regu- late atmospheric CO2 levels. As discussed in §4.4, if the transition to diffusion limited escape is early in a planet's history, then magnetic shielding is possible regardless of tectonic regime, and many planets will likely be able to keep their volatiles and possibly develop habitable climates. However, if the transition occurs after the man- tle's thermal adjustment period, then plate tectonics is likely necessary for magnetic shielding, and fewer planets will be able to retain surface water. Additional divergence points are possible later in planetary evolution if climate, tectonics, and the magnetic field are tightly coupled. For example, if the carbon cycle and plate tectonics act as a self-sustaining feedback, where high erosion rates, supplied by plate tectonics, are required to maintain kinetically limited weathering and thus keep surface temperatures cool enough for plate tectonics (see §3.3), then the loss of plate tectonics would lead directly to a hot climate state that likely precludes plate tectonics from re-initiating. Without plate tectonics to enhance erosion, the inhospitably hot climate that results from supply limited weathering would likely be permanent. Such tight coupling between plate tectonics and the carbon cycle is most likely on planets with small exposed land areas or high total CO2 budgets. Another divergence point is possible for planets orbiting very active solar mass stars or very close to active small mass stars where stronger stellar winds can more efficiently strip atmospheric volatiles. For these planets cessation of the core dynamo could cause significant water loss, in turn halting silicate weathering and, due to the ensuing hot climate, likely shutting down plate tectonics as well (see Figure 13 for a schematic illustration of the divergence points discussed in this paragraph). Size is also a major factor in terrestrial planet evolution. Large planets are expected to have wider habitable zones due to the influence of higher gravity on atmospheric scale height and the greenhouse effect (Kopparapu et al., 2014), but also shallower ocean basins and thus less exposed land, unless feedbacks can regulate ocean volume such that continents are always exposed (Cowan & Abbot, 2014, see also §6). The influence of size on plate tectonics and magnetic field generation is also unclear. Previous studies have found that plate tectonics is more likely on larger planets (Valencia et al., 2007; Valencia & O'Connell, 2009; van Heck & Tackley, 2011; Foley et al., 2012), less likely on larger planets (O'Neill & Lenardic, 2007; Kite et al., 2009; Stein et al., 2013; Noack & Breuer, 2014; Stamenkovi´c & Breuer, 2014; Miyagoshi et al., 2014; Tachinami et al., 2014), or that size is relatively unimportant (Korenaga, 2010a). Different studies reach very different conclusions because of 41 Figure 13: Schematic diagram of possible scenarios where failure of the feedbacks between climate, plate tectonics, and the magnetic field leads to divergent evolution of terrestrial planets within the habitable zone. the large uncertainties in the rheological mechanism necessary for generating plate tectonics, how key features such as internal heating rate scale with size, and how mantle properties are affected by pressure and temperature. The influence of size on magnetic field strength is likewise debatable. Several studies have found a weak dependence of field strength and lifetime on planet and core size, and possibly a peak in strength for Earth-sized planets (Gaidos et al., 2010; Tachinami et al., 2011; Driscoll & Olson, 2011; Van Summeren et al., 2013). Generally, larger dynamo regions are expected to produce stronger magnetic fields (e.g. Christensen et al., 2009), but variations in more subtle properties, like mantle and core composition, likely play a fundamental role. The coupling between plate tectonics, climate, and the magnetic field is expected to apply to planets of different size, but future work is needed to place tighter constraints on the influence of size on this coupling and on planetary evolution. The issue of size highlights the many uncertainties remaining in our knowledge of planetary dynamics and evolution. Each aspect of planetary dynamics that is important for magnetic, tectonic, and climate evolution needs to be better under- stood before more rigorous predictions or interpretations can be made. In particular the interactions between different components of the planetary system, specifically interactions between surface tectonics, mantle convection, and the long-term carbon cycle, between mantle convection and the core dynamo, and between the magnetic field, atmospheric escape, and climate, deserve significant attention. With a large number of rocky exoplanets already discovered, many of which are in their respective 42 Loss of magnetic fieldHIncreased escape rateif escape is magnetically-limitedIf water is lost, CO2 builds up in atmosphere until degassing ceases, plate tectonicsceases, and Venus-type planet developsLoss of plate tectonicsWeathering becomes supplylimitedCarbon cycle can still stabilize climate, but loss of magnetic field increases escape rateMagnetic field still maintainedor water loss insignificantHot, CO2 rich climateforms habitable zones (Batalha, 2014), and more certain to follow, improving our knowledge of planetary evolution is an important goal. 6 Future Directions In addition to furthering our understanding of plate tectonics, magnetic field gener- ation, climate evolution, and the interactions between these processes, there are a number of new questions and research topics that must be addressed to advance our knowledge of planetary evolution. We summarize a few of these important questions below. 1) What are the material properties of Earth-like and non-Earth-like terrestrial planets at high temperature, high pressure conditions? Without tighter constraints on basic properties like density, viscosity, and thermal conductiv- ity models of exoplanet evolution will continue to be highly uncertain. Conducting both laboratory and numerical experiments at the conditions relevant for Earth and super-Earths is challenging, but is also vital for determining how terrestrial plan- ets behave. Moreover, the composition of exoplanets could differ significantly from Earth, so constraints on the properties of non-Earth-like materials are necessary for modeling the evolution of these planets as well. 2) What controls the volume of water at a planet's surface (i.e. the size of the oceans and the amount of subaerial land)? Earth has maintained a relatively constant freeboard, or water level relative to the continents, throughout much of its history (Wise, 1974), despite the fact that there could be multiple oceans worth of water stored in the mantle (e.g. Ohtani, 2005). Are there feedbacks in Earth's deep water cycle acting to keep a large continental surface area exposed, as proposed by some (Kasting & Holm, 1992; Cowan & Abbot, 2014)? Given the impor- tance of exposed land to the long-term carbon cycle, this question is of fundamental importance to planetary evolution. 3) What are typical volatile abundances, in particular water and carbon dioxide, for rocky planets, and how much of this volatile inventory resides in the atmosphere just after accretion and magma ocean solidification? Both volatile inventory and initial atmospheric composition are important for plane- tary evolution. Thus, a better understanding of how and when volatiles are delivered during accretion, how much degassing occurs during accretion and solidification of a possible magma ocean, and how much atmospheric loss occurs during this early phase of planetary history, are all necessary for constraining planetary evolution. 4) Can strong gravitational tides render a planet uninhabitable? Planets experiencing strong gravitational tides (caused by nearby stars, planets, or satellites) 43 can generate significant internal heating via tidal dissipation, which can cause ex- treme surface volcanism and hinder dynamo action (Driscoll & Barnes, 2015). Effi- cient cooling of the interior could allow the orbits of such planets to circularize faster, minimizing the length of time spent in a tidally heated regime, and could move the planet in (or out) of the habitable zone. Future work should explore the details of tidal dissipation in the mantle and core, and how tides can influence long-term evolution. 5) How and when does plate tectonics initiate? Factors other than climate, like mantle temperature, can have an important control over whether plate tectonics can operate. Thus even with favorable climate conditions, other process such as mantle cooling may be necessary for Earth-like plate tectonics. Understanding the type of tectonics that might take place on a planet before plate tectonics, how much land and weatherable rock can be created through non-plate-tectonic volcanism, and the factors that then allow for Earth-like plate tectonics to develop, will all be crucial for determining how likely planets are to follow an evolutionary trajectory similar to Earth's. 7 Acknowledgements BF acknowledges funding from the NASA Astrobiology Institute under cooperative agreement NNA09DA81A. We thank Norm Sleep for a thorough and constructive review that helped to significantly improve the manuscript. No new data was used in producing this paper; all information shown is either obtained by solving the equations presented in the paper or is included in papers cited and listed in the references. 44 References Abbot, D. S., Cowan, N. B., & Ciesla, F. J. (2012). Indication of Insensitivity of Planetary Weathering Behavior and Habitable Zone to Surface Land Fraction. Astrophys. J., 756 , 178. Abbott, D., Burgess, L., Longhi, J., & Smith, W. H. F. (1994). An empirical thermal history of the Earh's upper mantle. J. Geophys. Res., 99 , 13835–13850. Abe, Y., Abe-Ouchi, A., Sleep, N. H., & Zahnle, K. J. (2011). Habitable Zone Limits for Dry Planets. Astrobiology, 11 , 443–460. Abe, Y., & Matsui, T. (1985). The formation of an impact-generated H2O atmo- sphere and its implications for the early thermal history of the earth. Proc. 15th Lunar and Planetary Sci. Conf. J. Geophys. Res., 90 , C545–C559. Ague, J. J., & Nicolescu, S. (2014). Carbon dioxide released from subduction zones by fluid-mediated reactions. Nature Geoscience, 7 , 355–360. Alt, J. C., & Teagle, D. A. H. (1999). The uptake of carbon during alteration of ocean crust. Geochim. Cosmochim. Acta, 63 , 1527–1535. Armann, M., & Tackley, P. J. (2012). Simulating the thermochemical magmatic and tectonic evolution of venus's mantle and lithosphere: Two-dimensional models. Journal of Geophysical Research: Planets (1991–2012), 117 . Barabash, S., Fedorov, A., Sauvaud, J., Lundin, R., Russell, C., Futaana, Y., Zhang, T., Andersson, H., Brinkfeldt, K., Grigoriev, A. et al. (2007). The loss of ions from venus through the plasma wake. Nature, 450 , 650–653. Batalha, N. M. (2014). Exploring exoplanet populations with NASA's Kepler Mis- sion. Proc. Natl. Acad. Sci., 111 , 12647–12654. Bercovici, D. (1993). A simple model of plate generation from mantle flow. Geophys. J. Int., 114 , 635–650. Bercovici, D. (2003). The generation of plate tectonics from mantle convection. Earth Planet. Sci. Lett., 205 , 107–121. Bercovici, D., & Karato, S. (2003). Theoretical analysis of shear localization in the lithosphere. In S. Karato, & H. Wenk (Eds.), Reviews in Mineralogy and Geochemistry: Plastic Deformation of Minerals and Rocks chapter 13. (pp. 387– 420). Washington, DC: Min. Soc. Am. volume 51. 45 Bercovici, D., & Ricard, Y. (2012). Mechanisms for the generation of plate tectonics by two-phase grain-damage and pinning. Phys. Earth Planet. Inter., 202 , 27–55. Bercovici, D., & Ricard, Y. (2013). Generation of plate tectonics with two-phase grain-damage and pinning: Source-sink model and toroidal flow. Earth Planet. Sci. Lett., 365 , 275–288. Bercovici, D., & Ricard, Y. (2014). Plate tectonics, damage, and inheritance. Nature, 508 , 513–516. Bercovici, D., Tackley, P., & Ricard, Y. (2015). The generation of plate tectonics from mantle dynamics. In D. Bercovici, & G. Schubert (chief editor) (Eds.), Treatise on Geophysics (pp. 271–318). New York: Elsevier volume 7, Mantle Dynamics. Berner, R. (2004). The Phanerozoic carbon cycle. Oxford University Press. Berner, R. A. (1992). Weathering, plants, and the long-term carbon cycle. Geochim. Cosmochim. Acta, 56 , 3225–3231. Berner, R. A. (1994). GEOCARB II; a revised model of atmospheric CO2 over Phanerozoic time. Am. J. Sci., 294 , 56–91. Berner, R. A., & Caldeira, K. (1997). The need for mass balance and feedback in the geochemical carbon cycle. Geology, 25 , 955. Berner, R. A., Lasaga, A. C., & Garrels, R. M. (1983). The carbonate-silicate geochemical cycle and its effect on atmospheric carbon dioxide over the past 100 million years. Am. J. Sci, 283 , 641–683. Bluth, G. J. S., & Kump, L. R. (1994). Lithologic and climatologic controls of river chemistry. Geochim. Cosmochim. Acta, 58 , 2341–2359. Brady, P. V. (1991). The effect of silicate weathering on global temperature and atmospheric CO2. J. Geophys. Res., 96 , 18101. Brady, P. V., & G´ıslason, S. R. (1997). Seafloor weathering controls on atmospheric CO2 and global climate. Geochim. Cosmochim. Acta, 61 , 965–973. Brain, D., Leblanc, F., Luhmann, J., Moore, T. E., & Tian, F. (2014). Planetary magnetic fields and climate evolution. In Comparative Climatology of Terrestrial Planets (pp. 487–501). University of Arizona Press volume 1. 46 Brantley, S., & Olsen, A. (2014). 7.3 - Reaction Kinetics of Primary Rock-Forming Minerals under Ambient Condition. In J. I. Drever, K. K. Turekian, & H. D. Holland (Eds.), Treatise on Geochemistry (Second Edition) (pp. 69 – 113). Oxford: Elsevier volume 7. Breuer, D., & Moore, W. (2007). Dynamics and thermal history of the terrestrial planets, the Moon, and Io. In T. Spohn; G. Schubert (chief editor) (Ed.), Treatise on Geophysics (pp. 299–348). New York: Elsevier volume 10, Planets and Moons. Bunge, H.-P. (2005). Low plume excess temperature and high core heat flux in- ferred from non-adiabatic geotherms in internally heated mantle circulation mod- els. Phys. Earth Planet. Inter., 153 , 3–10. Burton, M. R., Sawyer, G. M., & Granieri, D. (2013). Deep carbon emissions from volcanoes. Rev. Mineral. Geochem, 75 , 323–354. Caldeira, K. (1995). low- temperature seafloor alteration or terrestrial silicate-rock weathering? Am. J. Sci., 295 , 1077–1114. Long-term control of atmospheric carbon dioxide; Cawood, P. A., Hawkesworth, C. J., & Dhuime, B. (2013). The continental record and the generation of continental crust. Geological Society of America Bulletin, 125 , 14–32. C´ebron, D., & Hollerbach, R. (2014). Tidally driven dynamos in a rotating sphere. The Astrophysical Journal Letters, 789 , L25. Chassefiere, E. (1997). Loss of water on the young venus: the effect of a strong primitive solar wind. Icarus, 126 , 229–232. Christensen, U. (1984). Convection with pressure- and temperature-dependent non- newtonian rheology. Geophys. J. R. Astron. Soc, 77 , 343–384. Christensen, U., Holzwarth, V., & Reiners, A. (2009). Energy flux determines mag- netic field strength of planets and stars. Nature, 457 , 167–169. Coogan, L. A., & Dosso, S. E. (2015). Alteration of ocean crust provides a strong temperature dependent feedback on the geological carbon cycle and is a primary driver of the sr-isotopic composition of seawater. Earth Planet. Sci. Lett., 415 , 38–46. 47 Coogan, L. A., & Gillis, K. M. (2013). Evidence that low-temperature oceanic hydrothermal systems play an important role in the silicate-carbonate weathering cycle and long-term climate regulation. Geochem. Geophys. Geosyst., 14 , 1771– 1786. Cowan, N. B., & Abbot, D. S. (2014). Water Cycling between Ocean and Mantle: Super-Earths Need Not Be Waterworlds. Astrophys. J., 781 , 27. Crowley, J. W., & O'Connell, R. J. (2012). An analytic model of convection in a system with layered viscosity and plates. Geophys. J. Int., 188 , 61–78. Dasgupta, R., & Hirschmann, M. M. (2010). The deep carbon cycle and melting in Earth's interior. Earth Planet. Sci. Lett., 298 , 1–13. Davaille, A., & Jaupart, C. (1993). Transient high-Rayleigh-number thermal convec- tion with large viscosity variations. J. Fluid Mech., 253 , 141–166. Davies, G. (1992). On the emergence of plate tectonics. Geology, 20 , 963–966. Davies, G. (1999). Dynamic Earth. Cambridge Univ. Press. De Bresser, J., ter Heege, J., & Spiers, C. (2001). Grain size reduction by dynamic recrystallization: can it result in major rheological weakening? Intl. J. Earth Sci., 90 , 28–45. De Bresser, J., Peach, C., Reijs, J., & Spiers, C. (1998). On dynamic recrystallization during solid state flow: effects of stress and temperature. Geophys. Res. Lett., 25 , 3457–3460. Dehant, V., Lammer, H., Kulikov, Y., Griessmeier, J., Breuer, D., Verhoeven, O., Karatekin, O., Van Hoolst, T., Korablev, O., & Lognonn´e, P. (2007). Planetary magnetic dynamo effect on atmospheric protection of early earth and mars. Space Science Reviews, 129 , 279–300. Delano, J. W. (2001). Redox History of the Earth's Interior since 3900 Ma: Impli- cations for Prebiotic Molecules. Origins Life Evol. B., 31 , 311–341. Dessert, C., Dupr´e, B., Gaillardet, J., Fran¸cois, L. M., & Allegre, C. J. (2003). Basalt weathering laws and the impact of basalt weathering on the global carbon cycle. Chem. Geol., 202 , 257–273. Donahue, T. (1999). New analysis of hydrogen and deuterium escape from venus. Icarus, 141 , 226–235. 48 Driscoll, P., & Bercovici, D. (2013). Divergent evolution of earth and venus: Influence of degassing, tectonics, and magnetic fields. Icarus, 226 , 1447–1464. Driscoll, P., & Bercovici, D. (2014). On the thermal and magnetic histories of Earth and Venus: Influences of melting, radioactivity, and conductivity. Phys. Earth Planet. Inter., 236 , 36–51. Driscoll, P., & Olson, P. (2011). Optimal dynamos in the cores of terrestrial exo- planets: Magnetic field generation and detectability. Icarus, 213 , 12–23. Driscoll, P. E., & Barnes, R. (2015). Tidal heating of earth-like exoplanets around m stars: Thermal, magnetic, and orbital evolutions. Astrobiology, 15 , 739–760. Drury, M. R., Vissers, R. L. M., van der Wal, D., & Hoogerduijn Strating, E. H. (1991). Shear localisation in upper mantle peridotites. Pure Appl. Geophys., 137 , 439–460. Dupr´e, B., Dessert, C., Oliva, P., Godd´eris, Y., Viers, J., Fran¸cois, L., Millot, R., & Gaillardet, J. (2003). Rivers, chemical weathering and Earth's climate. C. R. Geosci., 335 , 1141–1160. Edmond, J. M., Palmer, M. R., Measures, C. I., Grant, B., & Stallard, R. F. (1995). The fluvial geochemistry and denudation rate of the Guayana Shield in Venezuela, Colombia, and Brazil. Geochim. Cosmochim. Acta, 59 , 3301–3325. Elkins-Tanton, L. T. (2008). Linked magma ocean solidification and atmospheric growth for Earth and Mars. Earth Planet. Sci. Lett., 271 , 181–191. Escart´ın, J., Hirth, G., & Evans, B. (2001). Strength of slightly serpentinized peri- dotites: Implications for the tectonics of oceanic lithosphere. Geology, 29 , 1023. Etheridge, M. A., & Wilkie, J. C. (1979). Grainsize reduction, grain boundary sliding and the flow strength of mylonites. Tectonophysics, 58 , 159–178. Evans, R. L., Hirth, G., Baba, K., Forsyth, D., Chave, A., & Mackie, R. (2005). Geophysical evidence from the MELT area for compositional controls on oceanic plates. Nature, 437 , 249–252. Foley, B., & Becker, T. (2009). Generation of plate-like behavior and mantle het- erogeneity from a spherical, visco-plastic convection model. Geochem., Geophys., Geosyst., 10 . Q08001. 49 Foley, B. J. (2015). The role of plate tectonic-climate coupling and exposed land area in the development of habitable climates on rocky planets. Astrophys. J., 812 , 36–59. Foley, B. J., & Bercovici, D. (2014). Scaling laws for convection with temperature- dependent viscosity and grain-damage. Geophys. J. Int., 199 , 580–603. Foley, B. J., Bercovici, D., & Elkins-Tanton, L. T. (2014). Initiation of plate tectonics from post-magma ocean thermochemical convection. J. Geophys. Res. Solid Earth, 119 , 8538–8561. Foley, B. J., Bercovici, D., & Landuyt, W. (2012). The conditions for plate tectonics on super-earths: Inferences from convection models with damage. Earth Planet. Sci. Lett., 331?332 , 281 – 290. Fowler, A. C. (1993). Boundary layer theory and subduction. J. Geophys. Res., 98 , 21997–22005. Franck, S., Block, A., von Bloh, W., Bounama, C., Schellnhuber, H.-J., & Svirezhev, Y. (2000). Habitable zone for Earth-like planets in the solar system. Planet. Space Sci., 48 , 1099–1105. Franck, S., Kossacki, K., & Bounama, C. (1999). Modelling the global carbon cycle for the past and future evolution of the earth system. Chem. Geol., 159 , 305–317. Froelich, F., & Misra, S. (2014). Was the late paleocene-early eocene hot because earth was flat? An ocean lithium isotope view of mountain building, continental weathering, carbon dioxide, and earth's cenozoic climate. Oceanography, 27 , 36– 49. Frost, D. J., Mann, U., Asahara, Y., & Rubie, D. C. (2008). The redox state of the mantle during and just after core formation. Phil. Trans. R. Soc. A, 366 , 4315–4337. Frost, D. J., & McCammon, C. A. (2008). The Redox State of Earth's Mantle. Annu. Rev. Earth Planet. Sci., 36 , 389–420. Gabet, E. J., & Mudd, S. M. (2009). A theoretical model coupling chemical weath- ering rates with denudation rates. Geology, 37 , 151–154. Gaidos, E., Conrad, C., Manga, M., & Hernlund, J. (2010). Thermodynamic limits on magnetodynamos in rocky exoplanets. The Astrophysical Journal , 718 , 596– 609. 50 Gaillardet, J., Dupr´e, B., Louvat, P., & Allegre, C. (1999). Global silicate weathering and co 2 consumption rates deduced from the chemistry of large rivers. Chem. Geol., 159 , 3–30. Gillis, K. M., & Coogan, L. A. (2011). Secular variation in carbon uptake into the ocean crust. Earth Planet. Sci. Lett., 302 , 385–392. Gislason, S. R., Oelkers, E. H., Eiriksdottir, E. S., Kardjilov, M. I., Gisladottir, G., Sigfusson, B., Snorrason, A., Elefsen, S., Hardardottir, J., Torssander, P., & Oskarsson, N. (2009). Direct evidence of the feedback between climate and weathering. Earth Planet. Sci. Lett., 277 , 213–222. Goldblatt, C., Robinson, T. D., Zahnle, K. J., & Crisp, D. (2013). Low simulated radiation limit for runaway greenhouse climates. Nature Geosci., 6 , 661–667. Gough, D. O. (1981). Solar interior structure and luminosity variations. Solar Phys., 74 , 21–34. Gregory, R. T., & Taylor, H. P. (1981). An oxygen isotope profile in a section of Cretaceous oceanic crust, Samail Ophiolite, Oman: Evidence for δ18O buffering of the oceans by deep (> 5 km) seawater-hydrothermal circulation at mid-ocean ridges. J. Geophys. Res., 86 , 2737–2755. Griessmeier, J., Stadelmann, A., & et al. (2009). On the protection of extrasolar Icarus, 199 , earth-like planets around k/m stars against galactic cosmic rays. 526–535. Gurnis, M., Zhong, S., & Toth, J. (2000). On the competing roles of fault reactiva- tion and brittle failure in generating plate tectonics from mantle convection. In M. A. Richards, R. Gordon, & R. van der Hilst (Eds.), History and Dynamics of Global Plate Motions, Geophys. Monogr. Ser. (pp. 73–94). Washington, DC: Am. Geophys. Union volume 121. Hacker, B. R. (1996). Eclogite formation and the rheology, buoyancy, seismicity, and H2O content of oceanic crust. In G. Bebout, D. Scholl, S. Kirby, & J. P. Platt (Eds.), Subduction Top to Bottom (pp. 337–346). AGU Geophysical Monograph Series volume 96. Hacker, B. R., Abers, G. A., & Peacock, S. M. (2003). Subduction factory 1. Theo- retical mineralogy, densities, seismic wave speeds, and H2O contents. J. Geophys. Res., 108 , 2029. 51 Hamano, K., Abe, Y., & Genda, H. (2013). Emergence of two types of terrestrial planet on solidification of magma ocean. Nature, 497 , 607–610. Hart, M. H. (1978). The evolution of the atmosphere of the earth. Icarus, 33 , 23–39. Hart, M. H. (1979). Habitable zones about main sequence stars. Icarus, 37 , 351–357. Herzberg, C., Condie, K., & Korenaga, J. (2010). Thermal history of the Earth and its petrological expression. Earth and Planetary Science Letters, 292 , 79–88. Hilairet, N., Reynard, B., Wang, Y., Daniel, I., Merkel, S., Nishiyama, N., & Petit- girard, S. (2007). High-Pressure Creep of Serpentine, Interseismic Deformation, and Initiation of Subduction. Science, 318 , 1910–. Hilley, G. E., Chamberlain, C. P., Moon, S., Porder, S., & Willett, S. D. (2010). Com- petition between erosion and reaction kinetics in controlling silicate-weathering rates. Earth Planet. Sci. Lett., 293 , 191–199. Hirth, G., & Kohlstedt, D. (1996). Water in the oceanic upper mantle: implications for rheology, melt extraction and the evolution of the lithosphere. Earth Planet. Sci. Lett., 144 , 93–108. Hirth, G., & Kohlstedt, D. (2003). Rheology of the upper mantle and the mantle wedge: a view from the experimentalists. In J. Eiler (Ed.), Subduction Factory Mongraph (pp. 83–105). Washington, DC: Am. Geophys. Union volume 138. Hoink, T., Lenardic, A., & Richards, M. (2012). Depth-dependent viscosity and mantle stress amplification: implications for the role of the asthenosphere in main- taining plate tectonics. Geophys. J. Int., 191 , 30–41. Hubbert, M. K., & Rubey, W. W. (1959). Role of fluid pressure in mechanics of overthrust faulting i. mechanics of fluid-filled porous solids and its application to overthrust faulting. Geological Society of America Bulletin, 70 , 115–166. Hunten, D. M. (1973). The Escape of Light Gases from Planetary Atmospheres. J. Atmos. Sci., 30 , 1481–1494. Ingersoll, A. P. (1969). The runaway greenhouse: A history of water on Venus. J. Atm. Sci., 26 , 1191–1198. Ivanov, M. A., & Head, J. W. (2013). The history of volcanism on venus. Planetary and Space Science, 84 , 66–92. 52 Jellinek, A. M., & Jackson, M. G. (2015). Connections between the bulk composition, geodynamics and habitability of Earth. Nature Geosci., 8 , 587–593. Jin, D., Karato, S., & Obata, M. (1998). Mechanisms of shear localization in the con- tinental lithosphere: Inference from the deformation microstructures of peridotites from the Ivrea zone, northwestern Italy. J. Struct. Geol., 20 , 195–209. Kadoya, S., & Tajika, E. (2014). Conditions for Oceans on Earth-like Planets Orbit- ing within the Habitable Zone: Importance of Volcanic CO2 Degassing. Astrophys. J., 790 , 107. Karato, S., & Wu, P. (1993). Rheology of the Upper Mantle: A Synthesis. Science, 260 , 771–778. Kasting, J. F. (1988). Runaway and moist greenhouse atmospheres and the evolution of earth and Venus. Icarus, 74 , 472–494. Kasting, J. F., & Ackerman, T. P. (1986). Climatic consequences of very high carbon dioxide levels in the earth's early atmosphere. Science, 234 , 1383–1385. Kasting, J. F., & Catling, D. (2003). Evolution of a Habitable Planet. Annu. Rev. Astron. Astrophys., 41 , 429–463. Kasting, J. F., Eggler, D. H., & Raeburn, S. P. (1993a). Mantle Redox Evolution and the Oxidation State of the Archean Atmosphere. Journal of Geology, 101 , 245–257. Kasting, J. F., & Holm, N. G. (1992). What determines the volume of the oceans? Earth Planet. Sci. Lett., 109 , 507–515. Kasting, J. F., Whitmire, D. P., & Reynolds, R. T. (1993b). Habitable Zones around Main Sequence Stars. Icarus, 101 , 108–128. Kelemen, P. B., & Manning, C. E. (2015). Reevaluating carbon fluxes in subduction zones, what goes down, mostly comes up. Proc. Natl. Acad. Sci., 112 , E3997– E4006. Kerrick, D. M., & Connolly, J. A. D. (2001). Metamorphic devolatilization of sub- ducted marine sediments and the transport of volatiles into the Earth's mantle. Nature, 411 , 293–296. Kite, E. S., Manga, M., & Gaidos, E. (2009). Geodynamics and Rate of Volcanism on Massive Earth-like Planets. Astrophys. J., 700 , 1732–1749. 53 Kodama, T., Genda, H., Abe, Y., & Zahnle, K. J. (2015). Rapid Water Loss Can Extend the Lifetime of Planetary Habitability. Astrophys. J., 812 , 165. Kohlstedt, D., Evans, B., & Mackwell, S. (1995). Strength of the lithosphere: Con- straints imposed by laboratory experiments. J. Geophys. Res., 100 , 17,587–17,602. Kopparapu, R. K., Ramirez, R. M., SchottelKotte, J., Kasting, J. F., Domagal- Goldman, S., & Eymet, V. (2014). Habitable Zones around Main-sequence Stars: Dependence on Planetary Mass. Astrophys. J., 787 , L29. Korenaga, J. (2006). Archean geodynamics and the thermal evolution of earth. In K. Benn, J.-C. Mareschal, & K. Condie (Eds.), Archean Geodynamics and Environments (pp. 7–32). AGU Geophysical Monograph Series volume 164. Korenaga, J. (2007). Thermal cracking and the deep hydration of oceanic litho- J. Geophys. Res., 112 . sphere: A key to the generation of plate tectonics? Doi:10.1029/2006JB004502. Korenaga, J. (2010a). On the likelihood of plate tectonics on super-earths: Does size matter? Astrophys. J., 725 , L43. Korenaga, J. (2010b). Scaling of plate tectonic convection with pseudoplastic rheol- ogy. J. Geophys. Res., 115 , B11405. Kump, L. R., & Arthur, M. A. (1997). Global chemical erosion during the cenozoic: Weatherability balances the budgets. In W. F. Ruddiman (Ed.), Tectonic Uplift and Climate Change (pp. 399–426). Springer. Kump, L. R., Brantley, S. L., & Arthur, M. A. (2000). Chemical Weathering, Atmo- spheric CO2, and Climate. Annual Review of Earth and Planetary Sciences, 28 , 611–667. Labrosse, S., & Jaupart, C. (2007). Thermal evolution of the Earth: Secular changes and fluctuations of plate characteristics. Earth Planet. Sci. Lett., 260 , 465–481. Lammer, H., Gudel, M., Kulikov, Y., Ribas, I., Zaqarashvili, T. V., Khodachenko, M. L., Kislyakova, K. G., Groller, H., Odert, P., Leitzinger, M. et al. (2012). Variability of solar/stellar activity and magnetic field and its influence on planetary atmosphere evolution. Earth, planets and space, 64 , 179–199. Lammer, H., Kasting, J., Chassefi`ere, E., Johnson, R., Kulikov, Y., & Tian, F. (2008). Atmospheric escape and evolution of terrestrial planets and satellites. Space Science Reviews, 139 , 399–436. 54 Landuyt, W., & Bercovici, D. (2009). Variations in planetary convection via the effect of climate on damage. Earth Planet. Sci. Lett., 277 , 29–37. Lebrun, T., Massol, H., Chassefi`eRe, E., Davaille, A., Marcq, E., Sarda, P., Leblanc, F., & Brandeis, G. (2013). Thermal evolution of an early magma ocean in inter- action with the atmosphere. J. Geophys. Res., 118 , 1155–1176. Lenardic, A., Jellinek, M., & Moresi, L.-N. (2008). A climate change induced tran- sition in the tectonic style of a terrestrial planet. Earth Planet. Sci. Lett., 271 , 34?42. Lenardic, A., & Kaula, W. (1994). Self-lubricated mantle convection: Two- dimensional models. Geophys. Res. Lett., 21 , 1707–1710. Luger, R., & Barnes, R. (2015). Extreme water loss and abiotic o2 buildup on planets throughout the habitable zones of m dwarfs. Astrobiology, 15 , 119–143. Lundin, R., Lammer, H., & Ribas, I. (2007). Planetary magnetic fields and solar forcing: Implications for atmospheric evolution. Space Science Reviews, 129 , 245– 278. Madhusudhan, N., Lee, K. K. M., & Mousis, O. (2012). A Possible Carbon-rich Interior in Super-Earth 55 Cancri e. Astrophys. J. Lett., 759 , L40. Maher, K., & Chamberlain, C. (2014). Hydrologic regulation of chemical weathering and the geologic carbon cycle. Science, 343 , 1502–1504. Marty, B., & Tolstikhin, I. N. (1998). Co 2 fluxes from mid-ocean ridges, arcs and plumes. Chem. Geol., 145 , 233–248. Michel, F. C. (1971). Solar-wind-induced mass loss from magnetic field-free planets. Planetary and Space Science, 19 , 1580–1583. Millot, R., Gaillardet, J., Dupr´e, B., & All`egre, C. J. (2002). The global control of silicate weathering rates and the coupling with physical erosion: new insights from rivers of the Canadian Shield. Earth Planet. Sci. Lett., 196 , 83–98. Mills, B., Lenton, T. M., & Watson, A. J. (2014). Proterozoic oxygen rise linked to shifting balance between seafloor and terrestrial weathering. Proc. Natl. Acad. Sci., 111 , 9073–9078. 55 Mills, B., Watson, A. J., Goldblatt, C., Boyle, R., & Lenton, T. M. (2011). Timing of Neoproterozoic glaciations linked to transport-limited global weathering. Nature Geoscience, 4 , 861–864. Miyagoshi, T., Tachinami, C., Kameyama, M., & Ogawa, M. (2014). On the vigor of mantle convection in super-earths. The Astrophysical Journal Letters, 780 , L8. Moore, T., & Horwitz, J. (2007). Stellar ablation of planetary atmospheres. Rev. Geophys, 45 , 1–34. Moore, W. B., & Webb, A. A. G. (2013). Heat-pipe earth. Nature, 501 , 501–505. Morbidelli, A., Chambers, J., Lunine, J. I., Petit, J. M., Robert, F., Valsecchi, G. B., & Cyr, K. E. (2000). Source regions and time scales for the delivery of water to Earth. Meteorit. Planet. Sci., 35 , 1309–1320. Moresi, L., & Solomatov, V. (1998). Mantle convection with a brittle lithosphere: Thoughts on the global tectonic style of the earth and venus. Geophys. J. Int., 133 , 669–682. Moresi, L.-N., & Solomatov, V. S. (1995). Numerical investigation of 2d convection with extremely large viscosity variations. Phys. Fluids, 7 , 2154–2162. Morgan, P., & Phillips, R. J. (1983). Hot spot heat transfer: Its application to venus and implications to venus and earth. Journal of Geophysical Research: Solid Earth (1978–2012), 88 , 8305–8317. Nakagawa, T., & Tackley, P. J. (2012). Influence of magmatism on mantle cooling, surface heat flow and Urey ratio. Earth Planet. Sci. Lett., 329 , 1–10. Nakajima, S., Hayashi, Y.-Y., & Abe, Y. (1992). A study on the 'runaway greenhouse effect' with a one-dimensional radiative-convective equilibrium model. J. Atmos. Sci., 49 , 2256–2266. Nimmo, F. (2002). Why does venus lack a magnetic field? Geology, 30 , 987–990. Nimmo, F., & Stevenson, D. J. (2000). Influence of early plate tectonics on the thermal evolution and magnetic field of Mars. J. Geophys. Res., 105 , 11969–11980. Noack, L., & Breuer, D. (2014). Plate tectonics on rocky exoplanets: Influence of initial conditions and mantle rheology. Planetary and Space Science, 98 , 41–49. 56 Ogawa, M., Schubert, G., & Zebib, A. (1991). Numerical simulations of three- dimensional thermal convection in a fluid with strongly temperature-dependent viscosity. J. Fluid Mech., 233 , 299–328. Ohtani, E. (2005). Water in the mantle. Elements, 1 , 25–30. Oliva, P., Viers, J., Dupr´e, B., Fortun´e, J. P., Martin, F., Braun, J. J., Nahon, D., & Robain, H. (1999). The effect of organic matter on chemical weathering: study of a small tropical watershed: Nsimi-Zo´et´el´e site, Cameroon. Geochim. Cosmochim. Acta, 63 , 4013–4035. O'Neill, C., Jellinek, A. M., & Lenardic, A. (2007). Conditions for the onset of plate tectonics on terrestrial planets and moons. Earth Planet. Sci. Lett., 261 , 20–32. O'Neill, C., & Lenardic, A. (2007). Geological consequences of super-sized Earths. Geophys. Res. Lett., 34 , 19204–19208. O'Neill, C., Lenardic, A., Moresi, L., Torsvik, T., & Lee, C.-T. (2007). Episodic precambrian subduction. Earth Planet. Sci. Lett., 262 , 552 – 562. O'Rourke, J. G., & Korenaga, J. (2012). Terrestrial planet evolution in the stagnant- lid regime: Size effects and the formation of self-destabilizing crust. Icarus, 221 , 1043–1060. O'Rourke, J. G., & Stevenson, D. J. (2016). Powering earth?s dynamo with magne- sium precipitation from the core. Nature, 529 , 387–389. Owen, J. E., & Adams, F. C. (2014). Magnetically controlled mass-loss from extra- solar planets in close orbits. Monthly Notices of the Royal Astronomical Society, 444 , 3761–3779. Oxburgh, E., & Parmentier, E. (1977). Compositional and density stratification in oceanic lithosphere-causes and consequences. J. Geol. Soc., 133 , 343–355. Phillips, R. J., Kaula, W. M., McGill, G. E., & Malin, M. C. (1981). Tectonics and evolution of Venus. Science, 212 , 879–887. Portenga, E. W., & Bierman, P. R. (2011). Understanding earth's eroding surface with 10 Be. GSA Today, 21 , 4–10. Pozzo, M., Davies, C., Gubbins, D., & Alf`e, D. (2014). Thermal and electrical conductivity of solid iron and iron–silicon mixtures at earth's core conditions. Earth and Planetary Science Letters, 393 , 159–164. 57 Raymo, M. E., & Ruddiman, W. F. (1992). Tectonic forcing of late Cenozoic climate. Nature, 359 , 117–122. Raymond, S. N., Quinn, T., & Lunine, J. I. (2004). Making other earths: dynamical simulations of terrestrial planet formation and water delivery. Icarus, 168 , 1–17. Regenauer-Lieb, K., & Yuen, D. (2003). Modeling shear zones in geological and planetary sciences: solid- and fluid- thermal-mechanical approaches. Earth Science Reviews, 63 , 295–349. Regenauer-Lieb, K., Yuen, D. A., & Branlund, J. (2001). The Initiation of Subduc- tion: Criticality by Addition of Water? Science, 294 , 578–581. Rice, J. R. (1992). Fault stress states, pore pressure distributions, and the weakness of the san andreas fault. In B. Evans, & T.-F. Wong (Eds.), Fault Mechanics and Transport Propoerties in Rocks (pp. 475–503). San Diego CA: Academic Press Ltd. volume 51. Richards, M., Yang, W.-S., Baumgardner, J., & Bunge, H.-P. (2001). Role of a low- viscosity zone in stabilizing plate tectonics: Implications for comparative terrestrial planetology. Geochem. Geophys. Geosystems (G3), 2 , 2000GC000115. Richter, F. M., Nataf, H.-C., & Daly, S. F. (1983). Heat transfer and horizontally averaged temperature of convection with large viscosity variations. J. Fluid Mech., 129 , 173–192. Ridgwell, A., & Zeebe, R. E. (2005). The role of the global carbonate cycle in the regulation and evolution of the Earth system. Earth Planet. Sci. Lett., 234 , 299–315. Riebe, C. S., Kirchner, J. W., & Finkel, R. C. (2004). Erosional and climatic effects on long-term chemical weathering rates in granitic landscapes spanning diverse climate regimes. Earth Planet. Sci. Lett., 224 , 547–562. Rolf, T., & Tackley, P. J. (2011). Focussing of stress by continents in 3D spherical mantle convection with self-consistent plate tectonics. Geophys. Res. Lett., 38 , 18301. Rudnick, R. L. (1995). Making continental crust. Nature, 378 , 571–578. Schubert, G., & Soderlund, K. (2011). Planetary magnetic fields: Observations and models. Physics of the Earth and Planetary Interiors, 187 , 92–108. 58 Shizgal, B., & Arkos, G. (1996). Nonthermal escape of the atmospheres of venus, earth, and mars. Reviews of Geophysics, 34 , 483–505. Sibson, R. (1977). Fault rocks and fault mechanisms. J. Geol. Soc., 133 , 191–213. Skemer, P., Warren, J. M., & Kelemen, P. B. (2010). Microstructural and rheological evolution of a mantle shear zone. J. Petrol., 51 , 43–53. Sleep, N. H. (1988). Gradual entrainment of a chemical layer at the base of the mantle by overlying convection. Geophysical Journal International , 95 , 437–447. Sleep, N. H., & Blanpied, M. L. (1992). Creep, compaction and the weak rheology of major faults. Nature, 359 , 687–692. Sleep, N. H., & Zahnle, K. (2001). Carbon dioxide cycling and implications for climate on ancient Earth. J. Geophys. Res., 106 , 1373–1400. Sleep, N. H., Zahnle, K., & Neuhoff, P. S. (2001). Initiation of clement surface conditions on the earliest Earth. Proc. Natl. Acad. Sci., 98 , 3666–3672. Sleep, N. H., Zahnle, K. J., & Lupu, R. E. (2014). Terrestrial aftermath of the Moon-forming impact. Phil. Trans. R. Soc. A, 372 , 20130172–20130172. Smithies, R. H., Van Kranendonk, M. J., & Champion, D. C. (2005). It started with a plume - early Archaean basaltic proto-continental crust. Earth Planet. Sci. Lett., 238 , 284–297. Smrekar, S. E., Elkins-Tanton, L., Leitner, J. J., Lenardic, A., Mackwell, S., Moresi, L., Sotin, C., & Stofan, E. R. (2013). Tectonic and thermal evolution of venus and the role of volatiles: Implications for understanding the terrestrial planets. (pp. 45–71). American Geophysical Union volume Exploring Venus as a Terrestrial Planet. Smrekar, S. E., Stofan, E. R., Mueller, N., Treiman, A., Elkins-Tanton, L., Helbert, J., Piccioni, G., & Drossart, P. (2010). Recent hotspot volcanism on venus from virtis emissivity data. Science, 328 , 605–608. Solomatov, V. (1995). Scaling of temperature- and stress-dependent viscosity con- vection. Phys. Fluids, 7 , 266–274. Solomatov, V. S. (2004). Initiation of subduction by small-scale convection. J. Geophys. Res., 109 , B01412. 59 Solomatov, V. S., & Moresi, L.-N. (1996). Stagnant lid convection on Venus. J. Geophys. Res., 101 , 4737–4754. Solomatov, V. S., & Moresi, L.-N. (1997). Three regimes of mantle convection with non-Newtonian viscosity and stagnant lid convection on the terrestrial planets. Geophys. Res. Lett., 24 , 1907–1910. Solomon, S. C., Smrekar, S. E., Bindschadler, D. L., Grimm, R. E., Kaula, W. M., McGill, G. E., Phillips, R. J., Saunders, R. S., Schubert, G., & Squyres, S. W. (1992). Venus tectonics - an overview of Magellan observations. J. Geophys. Res., 97 , 13199. Spohn, T., Acuna, M. H., Breuer, D., Golombek, M., Greeley, R., Halliday, A., Hauber, E., Jaumann, R., & Sohl, F. (2001). Geophysical Constraints on the Evolution of Mars. Space Sci. Rev., 96 , 231–262. Stallard, R. F., & Edmond, J. M. (1983). Geochemistry of the Amazon: 2. The influence of geology and weathering environment on the dissolved load. J. Geophys. Res., 88 , 9671–9688. Stamenkovi´c, V., & Breuer, D. (2014). The tectonic mode of rocky planets: Part 1–driving factors, models & parameters. Icarus, 234 , 174–193. Staudigel, H., Hart, S. R., Schmincke, H.-U., & Smith, B. M. (1989). Cretaceous ocean crust at DSDP Sites 417 and 418: Carbon uptake from weathering versus loss by magmatic outgassing. Geochim. Cosmochim. Acta, 53 , 3091–3094. Stegman, D. R., Jellinek, A. M., Zatman, S. A., Baumgardner, J. R., & Richards, M. A. (2003). An early lunar core dynamo driven by thermochemical mantle convection. Nature, 421 , 143–146. Stein, C., Lowman, J. P., & Hansen, U. (2013). The influence of mantle internal heating on lithospheric mobility: Implications for super-Earths. Earth Planet. Sci. Lett., 361 , 448–459. Stein, C., Schmalzl, J., & Hansen, U. (2004). The effect of rheological parameters on plate behaviour in a self-consistent model of mantle convection. Phys. Earth Planet. Int., 142 , 225–255. Stein, M., & Goldstein, S. L. (1996). From plume head to continental lithosphere in the Arabian-Nubian shield. Nature, 382 , 773–778. 60 Stevenson, D., Spohn, T., & Schubert, G. (1983). Magnetism and thermal evolution of the terrestrial planets. Icarus, 54 , 466–489. Strom, R. G., Trask, N. J., & Guest, J. E. (1975). Tectonism and volcanism on Mercury. J. Geophys. Res., 80 , 2478–2507. Sundquist, E. T. (1991). Steady- and non-steady-state carbonate-silicate controls on atmospheric CO 2. Quat. Sci. Rev., 10 , 283–296. Tachinami, C., Ogawa, M., & Kameyama, M. (2014). Thermal convection of com- pressible fluid in the mantle of super-earths. Icarus, 231 , 377–384. Tachinami, C., Senshu, H., & Ida, S. (2011). Thermal evolution and lifetime of intrinsic magnetic fields of super-earths in habitable zones. The Astrophysical Journal , 726 , 70. Tackley, P. (2000a). The quest for self-consistent generation of plate tectonics in In M. A. Richards, R. Gordon, & R. van der Hilst mantle convection models. (Eds.), History and Dynamics of Global Plate Motions, Geophys. Monogr. Ser. (pp. 47–72). Washington, DC: Am. Geophys. Union volume 121. Tackley, P. (2000b). Self-consistent generation of tectonic plates in time- dependent, three-dimensional mantle convection simulations, 1. pseudoplastic yielding. Geochem. Geophys. Geosyst. (G3), 1 , 2000GC000036. Tajika, E., & Matsui, T. (1990). The evolution of the terrestrial environment. In H. E. Newsom, & J. H. Jones (Eds.), Origin of the Earth (pp. 347–370). Tajika, E., & Matsui, T. (1992). Evolution of terrestrial proto-CO2 atmosphere coupled with thermal history of the earth. Earth Planet. Sci. Lett., 113 , 251–266. Takai, K., Nakamura, K., Toki, T., Tsunogai, U., Miyazaki, M., Miyazaki, J., Hi- rayama, H., Nakagawa, S., Nunoura, T., & Horikoshi, K. (2008). Cell prolifer- ation at 122◦C and isotopically heavy CH4 production by a hyperthermophilic methanogen under high-pressure cultivation. Proc. Natl. Acad. Sci., 105 , 10949– 10954. Tarduno, J. A., Cottrell, R. D., Davis, W. J., Nimmo, F., & Bono, R. K. (2015). A hadean to paleoarchean geodynamo recorded by single zircon crystals. Science, 349 , 521–524. 61 Taylor, M., & Lavraud, B. (2008). Observation of three distinct ion populations at the kelvin-helmholtz-unstable magnetopause. In Annales Geophysicae-Atmospheres Hydrospheres and Space Sciences (pp. 1559–1566). [Paris]: Gauthier-Villars, 1988- volume 26. Toth, G., & Gurnis, M. (1998). Dynamics of subduction initiation at preexisting fault zones. J. Geophys. Res., 103 , 18053–18067. Tozer, D. (1985). Heat transfer and planetary evolution. Geophys. Surv., 7 , 213–246. Trail, D., Watson, E. B., & Tailby, N. D. (2011). The oxidation state of Hadean magmas and implications for early Earth's atmosphere. Nature, 480 , 79–82. Trompert, R., & Hansen, U. (1998). Mantle convection simulations with rheologies that generate plate-like behavior. Nature, 395 , 686–689. Turcotte, D., & Schubert, G. (2002). Geodynamics. (2nd ed.). New York: Cambridge Univ. Press. Turcotte, D. L. (1993). An episodic hypothesis for Venusian tectonics. J. Geophys. Res., 98 , 17061–17068. Valencia, D., & O'Connell, R. J. (2009). Convection scaling and subduction on Earth and super-Earths. Earth Planet. Sci. Letts., 286 , 492–502. Valencia, D., O'Connell, R. J., & Sasselov, D. D. (2007). Tectonics on Super-Earths. Astrophys. J., 670 , L45–L48. Inevitability of Plate van der Lee, S., Regenauer-Lieb, K., & Yuen, D. A. (2008). The role of water in connecting past and future episodes of subduction. Earth Planet. Sci. Lett., 273 , 15–27. van Heck, H., & Tackley, P. (2008). Planforms of self-consistently generated plates in 3d spherical geometry. Geophys. Res. Lett., 35 , L19312. van Heck, H. J., & Tackley, P. J. (2011). Plate tectonics on super-Earths: Equally or more likely than on Earth. Earth Planet. Sci. Lett., 310 , 252–261. Van Summeren, J., Gaidos, E., & Conrad, C. P. (2013). Magnetodynamo lifetimes for rocky, earth-mass exoplanets with contrasting mantle convection regimes. Journal of Geophysical Research: Planets, 118 , 938–951. 62 van Thienen, P., van den Berg, A. P., & Vlaar, N. J. (2004a). Production and recycling of oceanic crust in the early Earth. Tectonophysics, 386 , 41–65. van Thienen, P., Vlaar, N. J., & van den Berg, A. P. (2004b). Plate tectonics on the terrestrial planets. Phys. Earth Planet. Inter., 142 , 61–74. Velbel, M. A. (1993). Temperature dependence of silicate weathering in nature: How strong a negative feedback on long-term accumulation of atmospheric CO2 and global greenhouse warming? Geology, 21 , 1059. Viers, J., Oliva, P., Dandurand, J.-L., Dupre, B., & Gaillardet, J. (2014). 7.6 - Chemical Weathering Rates, CO2 Consumption, and Control Parameters Deduced from the Chemical Composition of Rivers. In J. I. Drever, K. K. Turekian, & H. D. Holland (Eds.), Treatise on Geochemistry (Second Edition) (pp. 175 – 194). Oxford: Elsevier volume 7. Vlaar, N. J. (1985). Precambrian geodynamical constraints. In The Deep Proterozoic Crust in the North Atlantic Provinces (pp. 3–20). Springer. Volk, T. (1987). Feedbacks between weathering and atmospheric CO2 over the last 100 million years. Am. J. Sci, 287 , 763–779. Wade, J., & Wood, B. J. (2005). Core formation and the oxidation state of the Earth. Earth Planet. Sci. Lett., 236 , 78–95. Walker, J., Hayes, P., & Kasting, J. (1981). A negative feedback mechanism for the long-term stabilization of Earth's surface temperature. J. Geophys. Res., 86 , 9776–9782. Warren, J. M., & Hirth, G. (2006). Grain size sensitive deformation mechanisms in naturally deformed peridotites. Earth Planet. Sci. Letts., 248 , 438 – 450. Warren, S. G., Brandt, R. E., Grenfell, T. C., & McKay, C. P. (2002). Snowball Earth: Ice thickness on the tropical ocean. J. Geophys. Res., 107 , 3167. Watson, A., Donahue, T., & Walker, J. (1981). The dynamics of a rapidly escaping atmosphere: applications to the evolution of earth and venus. Icarus, 48 , 150–166. Weller, M. B., & Lenardic, A. (2012). Hysteresis in mantle convection: Plate tectonics systems. Geophys. Res. Lett., 39 , 10202. 63 Weller, M. B., Lenardic, A., & O'Neill, C. (2015). The effects of internal heating and large scale climate variations on tectonic bi-stability in terrestrial planets. Earth Planet. Sci. Lett., 420 , 85–94. West, A. J. (2012). Thickness of the chemical weathering zone and implications for erosional and climatic drivers of weathering and for carbon-cycle feedbacks. Geology, 40 , 811–814. West, A. J., Galy, A., & Bickle, M. (2005). Tectonic and climatic controls on silicate weathering. Earth Planet. Sci. Lett., 235 , 211–228. White, A. F., & Blum, A. E. (1995). Effects of climate on chemical weathering in watersheds. Geochim. Cosmochim. Acta, 59 , 1729–1747. White, A. F., & Buss, H. L. (2014). 7.4 - Natural Weathering Rates of Silicate Minerals. In J. I. Drever, K. K. Turekian, & H. D. Holland (Eds.), Treatise on Geochemistry (Second Edition) (pp. 115 – 155). Oxford: Elsevier volume 7. White, R., & McKenzie, D. (1989). Magmatism at rift zones - The generation of volcanic continental margins and flood basalts. J. Geophys. Res., 94 , 7685–7729. White, S., Burrows, S., Carreras, J., Shaw, N., & Humphreys, F. (1980). On my- lonites in ductile shear zones. J. Struct. Geol., 2 , 175–187. Willenbring, J. K., Codilean, A. T., & McElroy, B. (2013). Earth is (mostly) flat: Apportionment of the flux of continental sediment over millennial time scales. Geology, 41 , 343–346. Wise, D. U. (1974). Continental margins, freeboard and the volumes of continents and oceans through time. In C. A. Burk, & C. L. Drake (Eds.), The geology of continental margins (pp. 45–58). Springer. Wolff, R., Goldstein, B., & Yeates, C. (1980). The onset and development of kelvin- helmholtz instability at the venus ionopause. Journal of Geophysical Research, 85 , 7697–7707. Wood, B. J., Walter, M. J., & Wade, J. (2006). Accretion of the Earth and segrega- tion of its core. Nature, 441 , 825–833. Wordsworth, R. D., & Pierrehumbert, R. T. (2013). Water Loss from Terrestrial Planets with CO2-rich Atmospheres. Astrophys. J., 778 , 154. 64 Zahnle, K., Arndt, N., Cockell, C., Halliday, A., Nisbet, E., Selsis, F., & Sleep, N. H. (2007). Emergence of a Habitable Planet. Space Sci. Rev., 129 , 35–78. Zhong, S. (2006). Constraints on thermochemical convection of the mantle from plume heat flux, plume excess temperature, and upper mantle temperature. Jour- nal of Geophysical Research: Solid Earth, 111 . 65
1504.05235
2
1504
2015-04-23T08:43:58
Modelling the brightness increase signature due to asteroid collisions
[ "astro-ph.EP" ]
We have developed a model to predict the post-collision brightness increase of sub-catastrophic collisions between asteroids and to evaluate the likelihood of a survey detecting these events. It is based on the cratering scaling laws of Holsapple and Housen (2007) and models the ejecta expansion following an impact as occurring in discrete shells each with their own velocity. We estimate the magnitude change between a series of target/impactor pairs, assuming it is given by the increase in reflecting surface area within a photometric aperture due to the resulting ejecta. As expected the photometric signal increases with impactor size, but we find also that the photometric signature decreases rapidly as the target asteroid diameter increases, due to gravitational fallback. We have used the model results to make an estimate of the impactor diameter for the (596) Scheila collision of D=49-65m depending on the impactor taxonomy, which is broadly consistent with previous estimates. We varied both the strength regime (highly porous and sand/cohesive soil) and the taxonomic type (S-, C- and D-type) to examine the effect on the magnitude change, finding that it is significant at early stages but has only a small effect on the overall lifetime of the photometric signal. Combining the results of this model with the collision frequency estimates of Bottke et al. (2005), we find that low-cadence surveys of approximately one visit per lunation will be insensitive to impacts on asteroids with D<20km if relying on photometric detections.
astro-ph.EP
astro-ph
Modelling the brightness increase signature due to asteroid collisions Ev McLoughlin, Alan Fitzsimmons, Alan McLoughlin Astrophysics Research Centre, School of Mathematics and Physics, Queen's University Belfast. University Rd, Belfast, Northern Ireland, UK, BT7 1NN Abstract We have developed a model to predict the post-collision brightness increase of sub-catastrophic collisions between asteroids and to evaluate the likelihood of a survey detecting these events. It is based on the cratering scaling laws of Holsapple and Housen (2007) and models the ejecta expansion following an impact as occurring in discrete shells each with their own velocity. We estimate the magnitude change between a series of target/impactor pairs, as- suming it is given by the increase in reflecting surface area within a photometric aperture due to the resulting ejecta. As expected the photometric signal increases with impactor size, but we find also that the photometric signature decreases rapidly as the target aster- oid diameter increases, due to gravitational fallback. We have used the model results to make an estimate of the impactor diameter for the (596) Scheila collision of D = 49− 65m depending on the impactor taxonomy, which is broadly consistent with previous estimates. We varied both the strength regime (highly porous and sand/cohesive soil) and the tax- onomic type (S-, C- and D-type) to examine the effect on the magnitude change, finding that it is significant at early stages but has only a small effect on the overall lifetime of the photometric signal. Combining the results of this model with the collision frequency estimates of Bottke et al. (2005), we find that low-cadence surveys of ∼one visit per luna- tion will be insensitive to impacts on asteroids with D < 20km if relying on photometric detections. Keywords: asteroids, composition, impact processes, collisional physics Email address: [email protected] (Ev McLoughlin) 5 1 0 2 r p A 3 2 . ] P E h p - o r t s a [ 2 v 5 3 2 5 0 . 4 0 5 1 : v i X r a Preprint submitted to Elsevier August 21, 2018 1. Introduction The main asteroid belt is collisionally dominated with large asteroids' shapes, sizes and surface geology controlled by impacts. Studies of collisions help us to understand the evolution of the shape of the asteroid population and in turn the formation of our Solar system. These studies may involve laboratory experiments, computer modelling or observational programmes. The evidence for collisions can be seen indirectly in main-belt asteroid families (Cellino et al., 2002), asteroid satellites and binaries (Merline et al., 2002). It can also be seen directly in recently observed collisions (Snodgrass et al., 2010; Jewitt et al., 2011; Stevenson et al., 2012). There are three possible collisions observed to date. In 2009 the 120 m diameter asteroid P/2010 A2 suffered a collision with a 6-9 m estimated diameter impactor (Snodgrass et al., 2010) (but see section 4.4). In 2010 another asteroid, (596) Scheila (113 km diameter), was hit with a ∼ 35m diameter impactor Jewitt et al. (2011). The most recent potential collision involved the object P/2012 F5 (Gibbs), which like others was originally identified as a potential main-belt comet Stevenson et al. (2012). Events like the (596) Scheila collision should occur approximately every 5 years and collisions with asteroids <10m even more often (Bodewits et al., 2011). Several recent surveys are capable of detecting collisions and cratering events. For example, the Canada-France-Hawaii Telescope Legacy Survey was used to search for Main- Belt comets among 25240 objects in 2003-2009 (Gilbert and Wiegert, 2010), the Thousand Asteroid Lightcurve Survey (924 objects) was conducted with the Canada-France-Hawaii Telescope in September 2006 (Masiero et al., 2009) and the Hawaii Trails project was conducted in 2009 (599 objects) (Hsieh, 2009). While none of the surveys mentioned above were specifically looking for main belt collisions, the methods used in search for main belt comets would have also revealed any collisional events. There are also current surveys fully or partly dedicated to discovering Near Earth Asteroids, such as Pan-STARRS 1 (Kaiser et al., 2002), the Lincoln Near-Earth Asteroid Research (LINEAR, responsible for discovery of P/2010 A2) (Stokes et al., 2000), the Catalina Sky Survey (Spahr et al., 1996) and the VST ATLAS survey (Shanks et al., 2013) that are all capable of detecting main-belt collisions. Much work has been done in modelling the parameters (i.e. shape of debris, brightness, total ejected mass, impactor mass) of known observed collisions (Kleyna et al., 2013), (Ishiguro et al., 2011a), (Holsapple and Housen, 2007), (Housen and Holsapple, 2011) ; 2 and hydrodynamic modelling of generalised collision (Benz and Asphaug, 1999). This work focuses solely on the magnitude change following an impact as it is most likely to be observable by optical telescopes. Rather than looking at a specific object in the main belt, the described model looks at what would be expected with generic asteroids. 2. Model description 2.1. Cratering physics Our model is based on the work by Holsapple and Housen (2007), who provide a summary of scaling laws that allows calculation of crater size using properties of the target and impactor, based on the results of impact experiments. These laws can also be used to calculate the evolution of the ejecta dispersal and consequently estimate the amount of material ejected and increase in brightness following a collision. The decrease in magnitude of the target asteroid is going to depend on the amount of material that was ejected and whether it is optically thin or not. At high impact speeds, transfer of the energy and momentum of the impactor into the target occurs over area on the order of impactor size, while the resulting crater usually exceeds this size by many times. It is therefore a reasonable approximation to assume that impact occurs as a point source. Using theoretical analyses of mechanics of crater forma- tion, Holsapple and Housen showed that the crater and ejected material characteristics depend on the quantity aU µδν, where a is the radius, U is the normal velocity component of the impactor and δ is the density of the impactor; µ and ν are scaling exponents. The scaling exponents depend on the material properties. Theoretical values of µ range from 1/3 to 2/3 (Holsapple and Schmidt, 1987) and are a measure of the energy dissipation by material; a more porous material can dissipate energy more effectively and will have a lower value of this exponent. Experimentally determined values of µ are ∼0.55 for non- porous materials (e.g. rocks and wet soils), 0.41 for moderately porous materials (e.g. sand and cohesive soils) and 0.33 to 0.40 for highly porous materials (Holsapple and Schmidt, 1987). Experimental values for ν were found to be the same for all materials at around 0.4 (Holsapple and Schmidt, 1987). By selecting appropriate material scaling parameters for a given impact and inserting them into a general expression for the relationship between radii of involved objects and crater size, a reasonably accurate estimate of the crater size (as well as crater formation time and transient crater growth) can be made. 3 We now summarise how we use the previous studies in our calculations. Consider a spherical, non-rotating asteroid of radius r following an impact from an object with radius a at sub earth point. The general form of equation for crater size R consists of strength and gravity term: where R = aK1(gravity term + strength term) − µ 2+µ gravity term = Sga U2 ( ρ δgrain 2ν µ ) strength term = ( Y ρU2 ) 2+µ 2 ( δgrain ρ ν(2+µ) µ ) (1) (2) (3) Here K1 is a scaling parameter (1.03, 1.17, 0.725 for sand/cohesive soil, wet soils/rock and highly porous material respectively; (Holsapple and Housen, 2007)), Y is the average strength of the target material, ρ is the grain density of target, U = 5 km s−1 (Bottke et al., 1994) is the normal velocity component, δgrain is the grain density of the impactor, Sg is the surface gravity of the target asteroid with mass M and radius r, calculated as follows: Sg = GM r2 (4) Depending on the asteroid type, different values of bulk density (for calculation of target asteroid mass) and grain density (for calculation of ejected mass) are used. The values and their sources are summarised in Table 1. Bulk densities of C- and S-type asteroids were taken from weighted averages of corresponding subclasses as summarised in Table 3 of Carry (2012). Grain densities of C- and S-types are assumed to be the same as their most likely meteorite analogues (Britt et al., 2002). Density of D-type asteroids is approximated by bulk and grain densities of the Tagish lake meteorite (Zolensky et al., 2002; Izawa et al., 2010). The range of material strengths used is presented in Table 2. The strength value selected in this study was varied for each taxonomic type to explore the relationship between the type, strength and the corresponding magnitude change. The crater radius R calculated in this way has a corresponding mass Mcrater: 4 Table 1: Average bulk and grain density of asteroids depending on taxonomic type. Asteroid type Grain density (kg m−3). 2710 (Britt et al., 2002) 3700 (Britt et al., 2002) 2770 (Izawa et al., 2010) Bulk density (kg m−3) 1840 (Carry, 2012) 2640 (Carry, 2012) 1670 (Zolensky et al., 2002) C S D Table 2: Average material strength used in calculations for a selected scaling. Material Average strength Y (MPa) Sand/cohesive soil Wet soil/rock Highly porous 0.18 (Holsapple, 1993) 1.14 (Holsapple, 1993) 0.001 (Holsapple and Housen, 2007) Mcrater = kcraterρR3 (5) where the scaling factor kcrater is taken to be 0.75 for cohesive soils, 0.8 for wet soils/rocks or 0.4 for highly porous material (Housen and Holsapple, 2011). As we are interested in the ejected mass, since it is only that which contributes to the observed magnitude change of the asteroid, the full crater mass will give an overestimate of brightness. The total crater volume is made up of a volume of ejected mass, a volume of the mass that is uplifted near the crater rim and a volume due to compaction. The fraction kejecta of the total crater mass that corresponds to ejected mass is of order 0.2-0.5 (Housen and Holsapple, 2011). Mejecta = kejectaMcrater (6) Throughout this work we assume kejecta of 0.3 as being most appropriate to asteroids. 2.2. Velocity shell model We consider the ejecta leaving the asteroid surface after the collision event. For sim- plicity, we assume that the debris expands spherically outwards from asteroid, the debris with each velocity vn forming a shell of radius rs (see Figure 1). Effects from rotation of the target or impactor are beyond the scope of the current model. Impact experiments show that there is no significant correlation between velocity and mass of the particles (Holsapple et al., 2002). Therefore, each velocity shell is taken to have the same parti- 5 cle size distribution described below, in Section 2.3. As our aim is to model observable brightening from Earth, we assume that the ejecta cloud is centred on the asteroid, as at early epochs the asteroid itself and the ejecta will be unresolved. We also assume that the brightness of the asteroid plus ejecta is measured through an aperture of fixed radius rap and centred on the asteroid. To obtain the amount of material that is visible in the aperture, and consequently the visible increase in asteroid magnitude, we need to look at the fraction of each debris shell that fits completely or partially in the aperture (ejecta visibility fraction fvis). (a) Velocity shells. (b) Spherical cap. Figure 1: Velocity shell model: (a) Ejecta expands spherically outwards from the asteroid. The nth shell expands with velocity vn. The portion of each shell that fits into the aperture rap is marked in solid line. (b) Spherical cap indicates the portion of each shell that fits the aperture. There are three possibilities for an individual shell at a given time t: 1. the entire shell fits within the aperture and hence fvis = 1 6 hhraprs 2. none of the shell is visible, i.e. the material is not above the surface of the asteroid as it has fallen back, fvis = 0 3. part of the shell fits into the aperture (0 < fvis < 1) (see Figure 1) In the latter case, the amount of material visible can be estimated by calculating the ratio of the surface area corresponding to the arc that fits into the aperture (spherical cap) to the total surface area of the sphere. fvis = 2(2πrh) 4πr2 (7) material, h = r −(cid:113) r2 − r2 Here h is the height of the cap and the extra factor of 2 accounts for near and far side ap and rap is the aperture radius that equals the radius of the cap. In the situation of no gravity acting on the debris, the radii at each velocity can be described straightforwardly by rs = vt, however we need to consider the gravitational case. For simplicity, we consider a single debris shell of radius rs and velocity v over the asteroid of radius r. The acceleration due to the asteroid's gravity that the debris experience is inversely proportional to the square of shell radius ( 1 r2 s derivative of radius with respect to time. ) and also happens to be the second d2rs dt2 = Cr−2 s (8) C is a constant of proportionality which we calculate by considering the boundary conditions. At the asteroid surface rs = r and the acceleration due to gravity is the surface gravity Sg. At the surface the original equation reduces to Cr−2 = Sg, which in turn gives C = Sgr2. Therefore, the radius of the spherical shell of debris at velocity v is governed by a second order differential equation of the form: d2rs dt2 = Sgr2r−2 s (9) with the boundary conditions that at t = 0, rs = r, drs dt = vinitial. Solving this equation will give the shell radii at all times t. The total mass within the aperture Mvis is the integral of all velocity shell fractions within the aperture i.e. the shell mass Mshell multiplied by fvis (cid:90) vmax Mvis = Mshellfvisdv In the code it is calculated by taking a sum of the individual contributions from each vmin 7 velocity shell, thus Mvis = Σvmax v=vmin Mshellfvis As the asteroid fragments will be ejected at a continuous range of velocities, increasing the number of shells considered improves the accuracy of calculations. 2.3. Magnitude change calculation At this stage we implicitly assume that the scattering function for the asteroid and the ejecta particles is the same and that the ejecta is optically thin throughout. Assuming all the ejected particles contribute to reflected light with the same albedo as the target asteroid, the total visible surface area after the collision is Ac = Aa + Ap where Aa is the area of target asteroid, Ap is the area of ejected particles. The latter can be calculated assuming that the particle size distribution follows a power law of the following form (Ishiguro et al., 2011a): N (a)da = (cid:40) N0( a a0 )qda, 0, amin ≤ a ≤ amax a < amin, a > amax (10) (cid:115) (cid:114) Here amin = 0.1 µm is chosen to be the minimum particle radius because scattering is inefficient for particle sizes much smaller than the wavelength of the scattered light, amax is the maximum particle radius, N0 is the reference dust abundance at reference size a0 and q = 3.5 is the assumed power law index (Jewitt, 2009). The maximum particle radius amax is calculated using the expression derived from the power-law function of the ejection velocity of dust particles (Ishiguro et al., 2011a). amax = k 1 V0 2GM r (11) Here V0=80 m s−1 is the reference ejection velocity of 1 µm particles, k = 1/4 is the power index of size dependence of ejection velocity, M and r are mass and radius of the target respectively. After integration we get cross-sectional area per unit mass of σa (m2 kg−1), and the total cross-sectional area of particles around the asteroid Ap are: max − a−0.5 a−0.5 max − a0.5 a0.5 min min σa = − 3 4ρgrain 8 (12) Ap = σaMvis (13) The change in magnitude is going to be related to the ratio of the total area after collision Ac and initial target asteroid area Aa: ∆m = −2.5 log10 Ac Aa (14) Any ejecta in front of the asteroid will obscure a cross-sectional area equal to its own area. Given that we assume the same scattering function for asteroid and ejecta, this means ejecta directly in front of the asteroid does not contribute to the net brightness. Any ejecta behind the asteroid is clearly not visible. Therefore all ejected material along the line of sight of the asteroid should be ignored for brightness calculations. This is implemented by the subtraction of a pair of spherical caps with radius equal to that of the asteroid. 3. Modelling results 3.1. Model setup We have looked at three taxonomic types that are common in the main belt - C, S and D types (DeMeo and Carry, 2014). The differences in the types are accounted for using their corresponding grain and bulk densities during calculation of the ejected mass and magnitude change. In the simulation we varied the taxonomic type (C, S, D) and strength (sand/cohesive soil and highly porous) value to see if any patterns emerge. We have used a 50 by 50 point grid of target and impactor radii, ranging from 1-100 km and 1-100 m respectively and logarithmically spaced, and 500 velocity shells linearly spaced over interval 10−2 to 102 m s−1. The lowest initial velocity has to be above zero for the purposes of calculations and the amount of material being ejected at and beyond 102 m s−1 is a very small fraction of the total ejected mass and therefore considered negligible, so we take 102 m s−1 as a upper limit of the initial velocity range. Given that there is a range of target and impactor pairs that would result in catastrophic disruption (i.e. a dispersive shattering event leaving no remnant larger than 1/2 of the original mass of target asteroid (Greenberg et al., 1978)) rather than in sub catastrophic collision, the results of our model for that region will not be physical. We have estimated the location of the disruption region by calculating the approximate impactor radius necessary for catastrophic disruption of a basalt target assuming average density of 2500 kg m−3 and impact velocity of 5 km s−1 9 (Benz and Asphaug, 1999). This region is marked in black on the figures. The code for this model was written in Python 2.7 (includes NumPy and SciPy packages), plotting of the output was performed in MatLab. The code requires a separate run for each pair of values for the taxonomic type and strength regime (i.e. C-type in sand/cohesive soil and C-type in highly porous regime would be two different code runs). Each run from impact until 7 days after impact with the above parameters takes approximately 2 hours on a single core of an Intel Core i5 2410M processor running at 2.3 GHz. 3.2. Model Results Figure 2 shows the resulting predicted change in magnitude post-collision for a C-type asteroid with both a highly porous and a sand/cohesive soil strength at two epochs after the impact (1 and 7 days). Figure 3 shows the calculated time it takes for the brightness increase to reach −1.0 magnitudes for C-types in the highly porous and the sand/cohesive soil strength regimes. This value has been chosen because the one sub-catastrophic aster- oid collision confirmed so far between the known main-belt asteroid (596) Scheila and an impactor was observed to have a brightness increase of (cid:39) −1 magnitude (Jewitt et al., 2011). Conclusive attribution of a collisional nature to a smaller magnitude decrease may be difficult, due to natural variations in brightness caused by rotational modulation of the light curves and uncertainty in asteroidal absolute magnitudes (Pravec et al., 2012). 10 (a) highly porous regime, 1 day (b) highly porous regime, 7 days (c) sand/cohesive soil regime, 1 day (d) sand/cohesive soil regime, 7 days Figure 2: C-type impactor collision with C-type target: 2(a)-2(b) highly porous strength regime: magni- tude change 1 and 7 days post-collision for a range of target and impactor radii. 2(c)-2(d) sand/cohesive soil strength regime: magnitude change 1 and 7 days post-collision for a range of target and impactor radii. Impactor radii range from 1 to 100 m, target radii: 1-100 km. Colour bar shows the magnitude change. Catastrophic disruption region is marked in black. Region where optical thickness of the debris is potentially significant is above the dashed line. Solid black lines indicate contours where magnitude change is -1, -3, -5, -7 and -9. 11 Target radius (km)Impactor radius (m)Magnitude change 1 day post collision(C target, C impactor, highly porous strength) 100101102102030405060708090100−9−8−7−6−5−4−3−2−1Target radius (km)Impactor radius (m)Magnitude change 7 days post collision(C target, C impactor, highly porous strength) 100101102102030405060708090100−9−8−7−6−5−4−3−2−1Target radius (km)Impactor radius (m)Magnitude change 1 day post collision(C target, C impactor, sand/cohesive soil strength) 100101102102030405060708090100−10−9−8−7−6−5−4−3−2−1Target radius (km)Impactor radius (m)Magnitude change 7 days post collision(C target, C impactor, sand/cohesive soil strength) 100101102102030405060708090100−10−9−8−7−6−5−4−3−2−1 (a) highly porous regime, 1 day (b) highly porous regime, 7 days (c) sand/cohesive soil regime, 1 day (d) sand/cohesive soil regime, 7 days Figure 3: Time taken for the magnitude change to fall to -1 magnitudes after a C-type impactor collision with C-type target. Figures 3(a) and 3(c) show the impactor-target regime where this occurs within 1 day for highly porous and sand/cohesive soil strength regimes respectively. Figures 3(b)-3(d) show the same data but now extended to show where this occurs within 7 days of impact. For each figure the colour coding for this timescale is shown on the right, with the truncated region (timescale greater than 1 day or 7 days) shown in dark blue on the plot. The region where a catastrophic disruption would occur is marked in black. 12 Target radius (km)Impactor radius (m)Time (days) at which magnitude differencefalls to −1, truncated to 1 day 1001011021020304050607080901000.10.20.30.40.50.60.70.80.9Target radius (km)Impactor radius (m)Time (days) at which magnitude differencefalls to −1, truncated to 7 days 100101102102030405060708090100123456Target radius (km)Impactor radius (m)Time (days) at which magnitude differencefalls to −1, truncated to 1 day 1001011021020304050607080901000.10.20.30.40.50.60.70.80.9Target radius (km)Impactor radius (m)Time (days) at which magnitude differencefalls to −1, truncated to 7 days 100101102102030405060708090100123456 Figure 4: Magnitude change with time for three targets of radii 1325, 1600 and 1930 m after collision with 20 m radius impactor (all S-type, sand/cohesive soil regime). Steps are clearly visible due to quantised nature of the modelled velocity shells and occur at a different position for each target. 13 100101102−6.5−6−5.5−5−4.5−4Time since impact (hours)Magnitude changeMagnitude change after collision of a 20 m radiusimpactor with 1325, 1600 and 1930 m radius targets 1325 m1600 m1930 m There are vertical structures in these plots that are not of physical significance, but are rather features of the model limitations. Three processes affect the total magnitude de- crease post collision: the constant expansion of debris moving out of the aperture, material falling back on the surface of the target asteroid and therefore disappearing from observa- tion; and material re-entering the aperture due to the gravitational attraction of the target asteroid. In an ideal simulation of the process, the change in magnitude as a function of time would be a smooth function, being the result of the interplay of all three processes. The function would be smooth because we expect the size and velocity distributions of the debris to be continuous. As velocity shells in this model are quantised, this leads to the creation of artefacts at certain values of target radius. The effect of this can be seen clearly in Figure 4, which shows the predicted brightness increase after an impact of a 20 m radius impactor onto 1325, 1600, 1930 m radius target (all S-type, sand/cohesive soil strength regime). For lower velocity shells that fall back onto the asteroid at early times these steps merge to give a continuous slope. Higher velocity shells never return as they have reached escape velocity, and thus never produce a sharp decline. For a small number of shells with intermediate velocity, the return of the shell to the surface will be sufficiently distinct temporally from the other shells to produce a clearly identifiable step. The velocity shells leaving the aperture do not produce a sharp decline, because the fraction of the shell in the aperture decreases as a continuous function of time after the impact. The location of these steps is independent of the impactor radius and depends solely on the target radius due to the dependance on the surface gravity. These steps are clearly visible in Figures 2 and after as vertical structures that should not be interpreted as being significant. One of the ways to minimise this and have a good resolution in the images is by using a large number of shells, however this increases the computation time. The 500 shells used in this study was selected as a compromise between resolution and run time. Figure 5 shows a comparison between a plot using 500 shells and 5000 shells (20 hours run-time). The latter shows a reduction in vertical structures without substantial difference in output. Difference images show a discrepancy between the 500 shells and 5000 shells results being confined to the vertical structures only. We have chosen to only run the model up to 7 days post-collision, because a large proportion of the material (by surface area) would leave the aperture due to radiation pressure effects, which are not currently included in the model. The time t it takes for 14 (a) 7 days, 500 shells (b) 7 days, 5000 shells Figure 5: C-type impactor collision with C-type target, sand/cohesive soil strength regime: magnitude change 7 days post-collision for a range of target and impactor radii. Impactor radii range from 1 to 100 m, target radii: 1-100 km. Colour bar shows the magnitude change. Catastrophic disruption region is marked in black. Comparison of using (a) 500 and (b) 5000 velocity shells. The plot with 5000 shells shows a clear reduction in the vertical structures in comparison to 500 shells plot. Solid black lines indicate contours where magnitude change is -1, -3, -5, -7 and -9. material to escape the aperture radius rap is approximately equal to (cid:19) 1 2 t ≈ (cid:18) 2rap a (15) where a is acceleration, that can be defined in terms of ratio between radiation pressure acceleration and the local gravity, β, and gravitation acceleration to the Sun at 1 AU g(cid:12): a = βg(cid:12). In turn, g(cid:12) ≈ GM(cid:12) 2 , where Rh is heliocentric distance. Rh This gives the approximate time for radiation pressure to remove material from the aperture as (cid:18) 2rapRh 2 (cid:19) 1 2 βGM(cid:12) t ≈ (16) where Rh is distance from the sun to the asteroid in question. Assuming β is on the order of ≈ 0.1(Fulle, 2004) for the small grains in the ejecta 15 Target radius (km)Impactor radius (m)Magnitude change 7 days post collision(C target, C impactor, sand/cohesive soil strength) 100101102102030405060708090100−10−9−8−7−6−5−4−3−2−1Target radius (km)Impactor radius (m)Magnitude change 7 days post collision(C target, C impactor, sand/cohesive soil strength) 100101102102030405060708090100−10−9−8−7−6−5−4−3−2−1 expected to dominate the scattering, Rh = 3 AU (mid-main belt), the time t is approxi- mately 2 days. Therefore, we do not present any model results that go beyond 7 days post collision, as without the inclusion of the radiation pressure they do not reflect physical reality. 3.3. Optical depth effects Throughout the modelling, all ejecta within the aperture and not behind the asteroid is treated as contributing to the brightness increase via the projected surface area. Effectively, it is assumed that optical depth τ is sufficiently low that the approximation τ (cid:39) 1 − e−τ (For τ < 0.2 this leads to a brightening overestimate of less than 10%). To is valid. estimate the potential optical depth effects, the ejecta cloud was divided into a series of concentric rings and the column density of each was calculated, then converted, using the particle size distribution, into the optical depth τ . Any rings with τ < 0.02 were ignored (this corresponds to an error of less than 1%), and any cases where more than 10% of ejecta mass was contained in rings with τ > 0.02 were considered to have their brightness potentially overestimated. We found that at one day after impact, ignoring opacity may lead to overestimating the brightness increase by up to twenty percent (∼ 0.2) magnitudes. The region where this may be relevant is outlined in shown in figures 2(a), 2(c), 7(a) and 8(a) showing the brightness 1 day post collision. Clearly for all impacts apart from those on the largest asteroids, the correction to the calculated magnitude change will have little effect on detection and can be ignored. By one week after impact, we calculate that optical depth effects are negligible for all target impactor pairs. 4. Discussion A collision between asteroids is a multi parameter problem requiring us to make some initial assumptions. This section is therefore divided into 4 parts. First we describe an application of the model to the only known collision in the parameter space explored - (596) Scheila. In the second part we examine the effect of the different strength regimes in a collision between C-type asteroids (as the most numerous in the outer belt where (596) Scheila is situated). In the third part we keep the strength regime the same (sand/cohesive soil) and vary the taxonomic types to examine the effect it has on the magnitude change. We conclude with application of the model to predicting the type of observable collisions by currently active surveys such as Pan-STARRS 1 and the Catalina Sky Survey. 16 4.1. Modelling the magnitude decrease in the (596) Scheila collision In December 2010 asteroid (596) Scheila was observed with the 0.68 meter Catalina Schmidt telescope to have increased in brightness by approximately 1.3 magnitudes in comparison to the previous month (Larson, 2010). There have been multiple photometric observations of (596) Scheila reported as summarised in Table 3, from which the impact is estimated to have occurred on December 3.5 UT, 2010 (Ishiguro et al., 2011b). All observations measured the total magnitude, i.e. from the asteroid and the surrounding debris. Jewitt et al. (2011) used a 105,600 km aperture for the HST observations of (596) Scheila and this value is used for the calculations throughout this section. We have selected these observations for comparison with our model as giving the true magnitude decrease because they were taken by the same instrument and in the same observational circumstances. Table 3: Summary of observed change in magnitude for (596) Scheila Date of observation (UT) Approximate time Magnitude post-collision (days) change 2010 Dec. 3.4 (Larson, 2010) 2010 Dec. 11.44-11.47 (Larson, 2010) 2010 Dec. 12 (Hsieh et al., 2012) 2010 Dec. 14-15 (Bodewits et al., 2014) 2010 Dec. 27.9 (Jewitt et al., 2011) 2011 Jan. 4.9 (Jewitt et al., 2011) 0 7.9 9 11 24.4 32.4 -1.3 -1.1 -0.86 -0.66 -1.26 -1.00 A collision with a smaller asteroid is the most likely cause of this brightening and there have been a range of impactor diameters suggested in literature: 35 m (Jewitt et al., 2011), 30-50 m (Ishiguro et al., 2011a), 30-90 m (Moreno et al., 2011), <100 m (Bodewits et al., 2011). The impactor diameters above are calculated using the estimated mass of high speed ejecta and the relationship m(v) ≈ mi for v = 10−2vi (Housen and Holsapple, 2011) where m(v) is the mass of ejecta above velocity v, mi is the impactor mass and vi the impactor velocity. Scheila has semimajor axis a = 2.926 AU, eccentricity e = 0.164, and inclination i = 14.7◦, which places it in the outer main belt. The most numerous asteroids in that region and therefore the most likely impactors are C-type (DeMeo and Carry, 2014), although (596) Scheila itself is a D-type (Yang and Hsieh, 2011). The taxonomic type determines the densities used in the model which affect the final output parameter of change in magnitude. 17 Figure 6 shows the post-collision magnitude change during 60 days for a collision between Scheila and C-, D- and S-type impactors of diameters reported in the literature. We also used the observations from Jewitt et al. (2011) to find the best fit size for each impactor type by matching the calculated magnitude change in our model to the observed lightcurve. The results are 59 m, 49 m and 65 m diameter impactors for S-, C- and D-types respectively. As has been mentioned in the previous section, we believe that in its current state the model gives best description of reality if we limit the results to 7 days post-collision. However, in the case of Scheila collision, which is an important real observed collisional event, we lack a sufficient number of data points of good quality in the first 7 days to make a meaningful comparison. However, the large aperture used in this study includes a large portion of the ejecta even after being significantly affected by radiation pressure, allowing comparison with our model. The model results are in approximate agreement with estimated impactor diameters by other authors. The differences most likely come from a variety of initial assumptions made by the different researchers, i.e. density, type of ejecta particle distribution, range of particle sizes and the estimated ejected mass. It is also clear from these results, that to the first order the taxonomic type of the impactor asteroid is not an important factor in diameter estimation. 4.2. Effect of strength regime on post-collision magnitude change for C-type asteroids. We consider a collision where both impactor and the target are generic C-type asteroids. Figures 2(a) and 2(b) show the magnitude decrease post-collision in highly porous regime, while Figures 2(c) and 2(d) show the effect of changing the regime to the sand/cohesive soil. Following the collision, a shock wave travels through the target asteroid and the outcome of the collision depends on how the stress wave propagates and is attenuated through the target. The cratering and amount of ejecta following an impact is determined by the target internal structure, porosity and strength. A porous asteroid would react to the impact differently than a more solid body and it is particularly important to consider impacts on these as most observable main belt asteroids should have significant macroporosities, apart from the very largest. It has been found that highly porous objects such as (253) Mathilde are particularly good at attenuating stress waves due to energy going into the compaction of the pores (Britt et al., 2002). An object with large macroporosity is most likely to be highly fractured or even a rubble pile. Dark and primitive asteroids are more likely to have significant porosity, however it is worth noting that it is not exclusive to C-types. Figures 2(a)/2(b) and 2(c)/2(d) are similar for both strength regimes at each point of time 18 (a) C-type impactors (b) S-type impactors (c) D-type impactors Figure 6: Magnitude change as a function of time following a collision of C-, S- and D-type impactors with a range of diameters with asteroid (596) Scheila. Individual points show reported magnitude change by various authors. Dot-dash line shows our fit to Jewitt et al.'s data. The best fit estimates of impactor diameter with our model are 59, 49 and 65m for C-, S- and D-type impactors respectively. considered, however the sand/cohesive soil regime shows a larger magnitude decrease due to more material ejected, while in the highly porous regime some of the impact energy goes 19 0102030405060−5−4−3−2−10Time (days)Magnitude changeMagnitude change following a collision of (596) Scheilawith C−type impactors of varying diameters. 59m (fit to Jewitt data)Gibbs in Larson et al.,2010Larson, 2010Hsieh et al.,2012Bodewits et al.,2011Jewitt et al.,20110102030405060−5−4−3−2−10Time (days)Magnitude changeMagnitude change following a collision of (596) Scheilawith S−type impactors of varying diameters. 49m (fit to Jewitt data)Gibbs in Larson et al.,2010Larson, 2010Hsieh et al.,2012Bodewits et al.,2011Jewitt et al.,20110102030405060−5−4−3−2−10Time (days)Magnitude changeMagnitude change following a collision of (596) Scheilawith D−type impactors of varying diameters. 65m (fit to Jewitt data)Gibbs in Larson et al.,2010Larson, 2010Hsieh et al.,2012Bodewits et al.,2011Jewitt et al.,2011 into compaction. This tendency is also reflected in plots of amount of time taken for the magnitude change to decrease to 1.0 magnitudes above the pre-impact brightness (Figure 3). The brightness increase caused by the ejecta reaches the assumed observable limit of approximately 1 magnitude faster for the highly porous case than the corresponding impactor-target pair undergoing a collision in the sand/cohesive soil regime. This is due to less ejecta being produced. Recent research by Carry (2012) indicates that the density may be dependant on target size, rather than being a fixed quantity for each taxonomic type. However, the data is sparse in the size range used in this model and unavailable for the D-types. Therefore, our model makes a simplifying assumption that the bulk density stays fixed for all sizes considered and depends only on the taxonomic type. 4.3. Effect of taxonomic type (C, D and S) on post-collision magnitude change in sand/cohesive soil regime It is important to understand the taxonomy effects on the collision lightcurve due to the established composition gradient across the main belt. To investigate we kept the strength regime (sand/cohesive soil) constant and varied taxonomic type of both target and impactor (for S-, C- and D-types). Figures 7 and 8 show the magnitude change 1 and 7 days post-collision for target-impactor pairs. Other combinations of target-impactor types are included in the Supplementary material (available online). S-types impacting each other are more likely to occur in the inner belt, while D-types impacting each other are more likely in the Trojan population. Generally, both sets of plots show qualitatively similar results for magnitude change with time, with a high magnitude change in the day 1, followed by the decline over time, with C-types having the least magnitude increase at 7 days. From our results we observed that collisions between the lowest grain density and highest porosity C-types produce the least amount of magnitude decrease lasting for less time that it does for other taxonomic types. The S-types which have lower porosity and the highest considered grain density show the opposite - a larger magnitude change lasting longer. D-type collisions fall between these regimes. Figure 9 shows a direct comparison between different combinations of target and im- pactor taxonomic types for a 20 km target and 50 m diameter impactor. In a collision with an S-type target, the largest magnitude change is produced by a S-type impactor, then C-type and D-type producing the least magnitude change. The same pattern is followed by a C-type target being impacted by S-, C- and D-type asteroids. However, a D-type 20 target deviates from this pattern, with an S-type impactor giving the largest change in magnitude, followed by D-types and C-types giving the least magnitude change. Overall, S-type impactors produce the brightest collision signature in all types of targets considered, due to their inherent high grain density. (a) 1 day (b) 7 days Figure 7: S-type impactor collision with S-type target, sand/cohesive soil strength regime: magnitude change 1 and 7 days post-collision for a range of target and impactor radii. Catastrophic disruption region is marked in black. Region where optical thickness of the debris is potentially significant is above the dashed line. Solid black lines indicate contours where magnitude change is -1, -3, -5, -7 and -9. 4.4. Detection from current sky surveys The catastrophic collision rate in the main belt has been calculated from observationally constrained dynamical models as ∼ 1 per year at a diameter D ∼ 100 m (Durda et al., 1998; O'Brien and Greenberg, 2003; Bottke et al., 2005). The general hope within the scientific community has been that such collisions would be detected via the on-going wide-field NEO surveys. However, several objects initially identified as potential collisional disruption events are now suspected of being caused by rotational disruption due to YORP spinup e.g. P/2010 A2 (Agarwal et al., 2013), P/2013 P5 (Jewitt et al., 2013). Additionally, a recent study by Denneau et al. (2014) using 1.2 years of Pan-STARRS 1 data found only one plausible candidate of a collision event, and concluded that collisional disruptions of 100-m scale asteroids may be extremely rare. Hence observing much more frequent 21 Target radius (km)Impactor radius (m)Magnitude change 1 day post collision(S target, S impactor, sand/cohesive soil strength) 100101102102030405060708090100−11−10−9−8−7−6−5−4−3−2−1Target radius (km)Impactor radius (m)Magnitude change 7 days post collision(S target, S impactor, sand/cohesive soil strength) 100101102102030405060708090100−10−9−8−7−6−5−4−3−2−1 (a) 1 day (b) 7 days Figure 8: D-type impactor collision with D-type target, Sand/cohesive soil strength regime: magnitude change 1 and 7 days post-collision for a range of target and impactor radii. Catastrophic disruption region is marked in black. Region where optical thickness of the debris is potentially significant is above the dashed line. Solid black lines indicate contours where magnitude change is -1, -3, -5, -7 and -9. sub-catastrophic collisions may be a viable method for constraining the overall collision rate. To look at the likelihood of detecting these events, we first look at collisions with 100 km diameter targets similar to Scheila. According to Bottke et al. (2005), the size of impactor that would disrupt a D = 100 km asteroid is D ≥ 25km and such a disruption would happen every ∼ 107 years in the main belt. Using their CoDDEM model size distribution, a subcatastrophic collision impact by a D ≥ 50 m impactor should occur once per (cid:39) 25 years, in rough agreement with the single such collision observed since the start of modern surveys in the 1990's. The current Pan-STARRS and Catalina surveys are complete down to an apparent magnitude V (cid:39) 20 (equivalent to D ∼ 2 km in the main-belt), but are sensitive to asteroids down to V (cid:39) 22. The Bottke et al. disruption frequency at D ∼ 2km is approximately one per 1000 years, requiring an impactor D (cid:39) 60 m. Scaling by the CoDDEM model size distribution predicts a collision between a D ∼ 2km asteroid and a D ≥ 10m impactor should occur approximately once per year. As ∼ 50% of the asteroid belt is visible at any time, current surveys should be able to detect a sub-catastrophic impact once every couple of years. 22 Target radius (km)Impactor radius (m)Magnitude change 1 day post collision(D target, D impactor, sand/cohesive soil strength) 100101102102030405060708090100−10−9−8−7−6−5−4−3−2−1Target radius (km)Impactor radius (m)Magnitude change 7 days post collision(D target, D impactor, sand/cohesive soil strength) 100101102102030405060708090100−10−9−8−7−6−5−4−3−2−1 Figure 9: The effect of varying taxonomic type in a collision of 20km target and 50m impactor on magnitude change as a function of time. So why are these collisions between smaller bodies not found? In reality, observational surveys are affected by factors such as weather, moonlight and occasional losses due to de- tector gaps in the cameras. Perhaps most importantly, current surveys have a cadence for most individual asteroids measured in weeks rather than days. Using a detection aperture of ∼ 1000 km equivalent to ∼ 1− 3 arcsec as appropriate to surveys such as Pan-STARRS, our model predicts a brightening of ≥ −1 magnitudes for D (cid:39) 10m impacts on D ∼ 2km for only 10 -- 20 days. In reality, effects on the ejecta such as radiation pressure and Ke- plerian shear may shorten this timescale further. We can use the calculations presented by Denneau et al. (2014) (their figure 5) to estimate the likelihood of detection. With a maximum magnitude change of ∼ 2 magnitudes and resulting rate of dimming of ≥ 0.1 magnitudes/day, the probability of the Pan-STARRS 1 survey detecting such a collision during the first 1.2 years of operation was < 0.05. Therefore we conclude that observational losses together with the rapid dispersal of the ejecta can significantly decrease the possi- 23 10−1100101−6−5−4−3−2−10time (days)Magnitude increaseMagnitude decrease with time post−collision for differentcombinations of target and impactor taxonomic types. S−target, S−impactorS−target, C−impactorS−target, D−impactorC−target, S−impactorC−target, C−impactorC−target, D−impactorD−target, S−impactorD−target, C−impactorD−target, D−impactor bility of photometric detection of sub-catastrophic impacts smaller that the (596) Scheila event, and this implies a current low probability of detection using automated photometry software such as MOPS (Denneau et al., 2013). Finally, another additional factor may be inaccurate absolute magnitudes in current reference catalogues, as investigated by Pravec et al. (2012). They found that smaller asteroids were systematically fainter than expected from catalogues. Hence predicted magnitudes would be brighter than in reality, and any small brightness increase due to a collision could be masked by an incorrectly calculated residual magnitude. Of course, during the first few days to weeks the ejecta may be visible as an extended ejecta cloud around the target asteroid, and could be detected via direct manual observation or software algorithms designed for comet comae detection, as in the case of (596) Scheila. On the other hand, if the cadence of a survey is a week or less i.e. significantly less than the decay timescale within the aperture, the chance of photometrically detecting a small impact should be substantially higher. In Figure 10 we plot the predicted total V -band magnitudes for C-on-C and S-on-S collisions 1 day and 7 days after impact, for asteroids at opposition at Rh = 3.5 AU. (Although S-types are predominantly found in the inner main belt, we calculate the predicted apparent magnitude at the likely maximum distance). First, it is clear that for the larger (but less frequent) impacts, the total magnitude will be relatively bright and would produce saturated images in large aperture surveys such as Pan-STARRS. However it may be possible to deal with this situation by either fitting to the wings of asteroid image point-spread function, or by recognising that an expected asteroid in the field was rejected in software processing due to its increased brightness. For fainter impact events, it is clear that Pan-STARRS with a limiting magnitude of V ∼ 22 is able to detect such impacts throughout the asteroid belt, as long as the asteroid is observed soon after the collision. Importantly, the forthcoming ATLAS programme will have a nightly cadence over the visible sky and have a limiting magnitude of V (cid:39) 20 (Tonry, 2011). From comparing both Figure 10 with the brightness increases presented earlier, we find that ATLAS would be able to detect almost all of our studied collisions as well, but here the survey cadence should not be an issue. Therefore we conclude that current and future high-cadence all sky surveys should be able to detect many more asteroid collisions at early epochs. 24 (a) C-type target impactor, 1 day (b) C-type target impactor, 7 days (c) S-type target impactor, 1 day (d) S-type target impactor, 7 days Figure 10: Visual magnitude following a collision. 10(a)-10(b) show magnitude 1 and 7 days post- collision for C-type impactor and target in the sand/cohesive soil strength regime for a range of target and impactor radii. 10(c)-10(d) show magnitude 1 and 7 days post-collision for S-type impactor and target in the sand/cohesive soil strength regime for a range of target and impactor radii. Impactor radii range from 1 to 100 m, target radii: 1-100 km. Colour bar shows the visual magnitude. Catastrophic disruption region is marked in black. Region where optical thickness of the debris is potentially significant is above the dashed line. 25 Target radius (km)Impactor radius (m)Visual magnitude 1 day post collision(C target, C impactor, sand/cohesive soil strength) 100101102102030405060708090100101214161820Target radius (km)Impactor radius (m)Visual magnitude 7 days post collision(C target, C impactor, sand/cohesive soil strength) 1001011021020304050607080901001214161820Target radius (km)Impactor radius (m)Visual magnitude 1 day post collision(S target, S impactor, sand/cohesive soil strength) 1001011021020304050607080901008101214161820Target radius (km)Impactor radius (m)Visual magnitude 7 days post collision(S target, S impactor, sand/cohesive soil strength) 100101102102030405060708090100101214161820 5. Summary The model presented in this paper predicts the brightness increases caused by impacts of small asteroids on larger asteroids with radii ≥ 1 km. We use the scaling laws of Holsapple and Housen (2007) and Housen and Holsapple (2011) to estimate the amount of ejected material. Our model separates the ejecta into discrete velocity shells, and assuming an ejecta size distribution calculates the magnitude change post-collision by calculating the surface cross-sectional area of the asteroid plus ejected material in a small photometric aperture. The scope of the model is limited to non-rotating asteroids whose ejecta particles do not interact with each other, however by extending the model results to the large apertures used for the reported ejects from the (596) Scheila collision, we find an estimate for the impactor size of 40 − 65 m (depending on taxonomic type) similar to previous studies in the literature. The model could be further improved by introducing particle light scattering laws, radiation pressure on the ejecta and the effect of rotation of the target asteroid. However this will lead to a significant increase the complexity of the model and will be developed in future. We believe our results are generic enough to be used to estimate the possibility of detection of such events by current automated surveys such as Pan-STARRS1 and the Catalina Sky Survey. The model estimates that within the parameter space examined (impactor radius 1 − 100 m, target radius 1 − 100 km, sub-catastrophic collisions) , a magnitude change of less than −1 is observable by an automated survey like Pan-STARRS 1 with effective aperture radii of 1000 km for only 10 -- 20 days, which implies low probability of detection given the current low cadence for individual asteroids. However, detection may still be possible by direct manual observation or software algorithms designed for comet coma detection. 6. Acknowledgements EMcL acknowledges support from the Astrophysics Research Centre, QUB. AF ac- knowledges support by STFC grant ST/L000709/1. EMcL and AF also thank Robert Jedicke for helpful comments during the preparation of this paper. 26 7. Supplimentary material (a) 1 day (b) 7 days Figure 11: D-type impactor collision with S-type target, sand/cohesive soil regime: magnitude change 1 and 7 days post-collision for a range of target and impactor radii. Impactor radii range from 1 to 100 m, target radii: 1-100 km. Colour bar shows the magnitude change. Catastrophic disruption region is marked in black. Region where optical thickness of the debris is potentially significant is above the dashed line. Solid black lines indicate contours where magnitude change is -1, -3, -5, -7 and -9. 27 Target radius (km)Impactor radius (m)Magnitude change 1 day post collision(S target, D impactor, sand/cohesive soil strength) 100101102102030405060708090100−11−10−9−8−7−6−5−4−3−2−1Target radius (km)Impactor radius (m)Magnitude change 7 days post collision(S target, D impactor, sand/cohesive soil strength) 100101102102030405060708090100−10−9−8−7−6−5−4−3−2−1 (a) 1 day (b) 7 days Figure 12: C-type impactor collision with S-type target, sand/cohesive soil regime: magnitude change 1 and 7 days post-collision for a range of target and impactor radii. Impactor radii range from 1 to 100 m, target radii: 1-100 km. Colour bar shows the magnitude change. Catastrophic disruption region is marked in black. Region where optical thickness of the debris is potentially significant is above the dashed line. Solid black lines indicate contours where magnitude change is -1, -3, -5, -7 and -9. 28 Target radius (km)Impactor radius (m)Magnitude change 1 day post collision(S target, C impactor, sand/cohesive soil strength) 100101102102030405060708090100−11−10−9−8−7−6−5−4−3−2−1Target radius (km)Impactor radius (m)Magnitude change 7 days post collision(S target, C impactor, sand/cohesive soil strength) 100101102102030405060708090100−10−9−8−7−6−5−4−3−2−1 (a) 1 day (b) 7 days Figure 13: S-type impactor collision with C-type target, sand/cohesive soil regime: magnitude change 1 and 7 days post-collision for a range of target and impactor radii. Impactor radii range from 1 to 100 m, target radii: 1-100 km. Colour bar shows the magnitude change. Catastrophic disruption region is marked in black. Region where optical thickness of the debris is potentially significant is above the dashed line. Solid black lines indicate contours where magnitude change is -1, -3, -5, -7 and -9. 29 Target radius (km)Impactor radius (m)Magnitude change 1 day post collision(C target, S impactor, sand/cohesive soil strength) 100101102102030405060708090100−10−9−8−7−6−5−4−3−2−1Target radius (km)Impactor radius (m)Magnitude change 7 days post collision(C target, S impactor, sand/cohesive soil strength) 100101102102030405060708090100−10−9−8−7−6−5−4−3−2−1 (a) 1 day (b) 7 days Figure 14: D-type impactor collision with C-type target, sand/cohesive soil regime: magnitude change 1 and 7 days post-collision for a range of target and impactor radii. Impactor radii range from 1 to 100 m, target radii: 1-100 km. Colour bar shows the magnitude change. Catastrophic disruption region is marked in black. Region where optical thickness of the debris is potentially significant is above the dashed line. Solid black lines indicate contours where magnitude change is -1, -3, -5, -7 and -9. 30 Target radius (km)Impactor radius (m)Magnitude change 1 day post collision(C target, D impactor, sand/cohesive soil strength) 100101102102030405060708090100−10−9−8−7−6−5−4−3−2−1Target radius (km)Impactor radius (m)Magnitude change 7 days post collision(C target, D impactor, sand/cohesive soil strength) 100101102102030405060708090100−10−9−8−7−6−5−4−3−2−1 (a) 1 day (b) 7 days Figure 15: S-type impactor collision with D-type target, sand/cohesive soil regime: magnitude change 1 and 7 days post-collision for a range of target and impactor radii. Impactor radii range from 1 to 100 m, target radii: 1-100 km. Colour bar shows the magnitude change. Catastrophic disruption region is marked in black. Region where optical thickness of the debris is potentially significant is above the dashed line. Solid black lines indicate contours where magnitude change is -1, -3, -5, -7 and -9. 31 Target radius (km)Impactor radius (m)Magnitude change 1 day post collision(D target, S impactor, sand/cohesive soil strength) 100101102102030405060708090100−10−9−8−7−6−5−4−3−2−1Target radius (km)Impactor radius (m)Magnitude change 7 days post collision(D target, S impactor, sand/cohesive soil strength) 100101102102030405060708090100−10−9−8−7−6−5−4−3−2−1 (a) 1 day (b) 7 days Figure 16: C-type impactor collision with D-type target, sand/cohesive soil regime: magnitude change 1 and 7 days post-collision for a range of target and impactor radii. Impactor radii range from 1 to 100 m, target radii: 1-100 km. Colour bar shows the magnitude change. Catastrophic disruption region is marked in black. Region where optical thickness of the debris is potentially significant is above the dashed line. Solid black lines indicate contours where magnitude change is -1, -3, -5, -7 and -9. 32 Target radius (km)Impactor radius (m)Magnitude change 1 day post collision(D target, C impactor, sand/cohesive soil strength) 100101102102030405060708090100−10−9−8−7−6−5−4−3−2−1Target radius (km)Impactor radius (m)Magnitude change 7 days post collision(D target, C impactor, sand/cohesive soil strength) 100101102102030405060708090100−10−9−8−7−6−5−4−3−2−1 References Agarwal, J., Jewitt, D., Weaver, H., 2013. Dynamics of Large Fragments in the Tail doi:10.1088/0004-637X/769/1/46, of Active Asteroid P/2010 A2. ApJ 769, 46. arXiv:1304.1814. Benz, W., Asphaug, E., 1999. Catastrophic Disruptions Revisited. Icarus 142, 5 -- 20. doi:10.1006/icar.1999.6204, arXiv:astro-ph/9907117. Bodewits, D., Kelley, M.S., Li, J.Y., Landsman, W.B., Besse, S., A'Hearn, M.F., 2011. Col- lisional Excavation of Asteroid (596) Scheila. ApJ Lett 733, L3. doi:10.1088/2041-8205/ 733/1/L3, arXiv:1104.5227. Bodewits, D., Vincent, J.B., Kelley, M.S.P., 2014. Scheila's scar: Direct evidence of impact surface alteration on a primitive asteroid. Icarus 229, 190 -- 195. doi:10.1016/j.icarus. 2013.11.003, arXiv:1310.8515. Bottke, W.F., Durda, D.D., Nesvorn´y, D., Jedicke, R., Morbidelli, A., Vokrouhlick´y, D., Levison, H.F., 2005. Linking the collisional history of the main asteroid belt to its dynamical excitation and depletion. Icarus 179, 63 -- 94. doi:10.1016/j.icarus.2005. 05.017. Bottke, W.F., Nolan, M.C., Greenberg, R., Kolvoord, R.A., 1994. Velocity distributions among colliding asteroids. Icarus 107, 255 -- 268. doi:10.1006/icar.1994.1021. Britt, D.T., Yeomans, D., Housen, K., Consolmagno, G., 2002. Asteroid Density, Porosity, and Structure. Asteroids III , 485 -- 500. Carry, B., 2012. Density of asteroids. Planetary and Space Science 73, 98 -- 118. doi:10. 1016/j.pss.2012.03.009, arXiv:1203.4336. Cellino, A., Bus, S.J., Doressoundiram, A., Lazzaro, D., 2002. Spectroscopic Properties of Asteroid Families. Asteroids III , 633. DeMeo, F.E., Carry, B., 2014. Solar System evolution from compositional mapping of the asteroid belt. Nature 505, 629 -- 634. doi:10.1038/nature12908. Denneau, L., Jedicke, R., Grav, T., Granvik, M., Kubica, J., Milani, A., Veres, P., Wain- scoat, R., Chang, D., Pierfederici, F., Kaiser, N., Chambers, K.C., Heasley, J.N., Mag- nier, E.A., Price, P.A., Myers, J., Kleyna, J., Hsieh, H., Farnocchia, D., Waters, C., 33 Sweeney, W.H., Green, D., Bolin, B., Burgett, W.S., Morgan, J.S., Tonry, J.L., Ho- dapp, K.W., Chastel, S., Chesley, S., Fitzsimmons, A., Holman, M., Spahr, T., Tholen, D., Williams, G.V., Abe, S., Armstrong, J.D., Bressi, T.H., Holmes, R., Lister, T., McMillan, R.S., Micheli, M., Ryan, E.V., Ryan, W.H., Scotti, J.V., 2013. The Pan- STARRS Moving Object Processing System. PASP 125, 357 -- 395. doi:10.1086/670337, arXiv:1302.7281. Denneau, L., Jewitt, D., Weaver, H., 2014. Observational Evidence for Spin-Up as the Dominant Source of Catastrophic Disruption of Small Main Belt Asteroids. Icarus In Press. Durda, D.D., Greenberg, R., Jedicke, R., 1998. Collisional Models and Scaling Laws: A New Interpretation of the Shape of the Main-Belt Asteroid Size Distribution. Icarus 135, 431 -- 440. doi:10.1006/icar.1998.5960. Fulle, M., 2004. Motion of cometary dust. pp. 565 -- 575. Gilbert, A.M., Wiegert, P.A., 2010. Updated results of a search for main-belt comets using the Canada-France-Hawaii Telescope Legacy Survey. Icarus 210, 998 -- 999. doi:10.1016/ j.icarus.2010.07.016. Greenberg, R., Hartmann, W.K., Chapman, C.R., Wacker, J.F., 1978. Planetesimals to planets - Numerical simulation of collisional evolution. Icarus 35, 1 -- 26. doi:10.1016/ 0019-1035(78)90057-X. Holsapple, K., Giblin, I., Housen, K., Nakamura, A., Ryan, E., 2002. Asteroid Impacts: Laboratory Experiments and Scaling Laws. Asteroids III , 443 -- 462. Holsapple, K.A., 1993. The scaling of impact processes in planetary sciences. Annual Review of Earth and Planetary Sciences 21, 333 -- 373. doi:10.1146/annurev.ea.21. 050193.002001. Holsapple, K.A., Housen, K.R., 2007. A crater and its ejecta: An interpretation of Deep Impact. Icarus 187, 345 -- 356. doi:10.1016/j.icarus.2006.08.029. Holsapple, K.A., Schmidt, R.M., 1987. Point source solutions and coupling parameters in cratering mechanics. Journal of Geophysical Research: Solid Earth 92, 6350 -- 6376. URL: http://dx.doi.org/10.1029/JB092iB07p06350, doi:10.1029/JB092iB07p06350. 34 Housen, K.R., Holsapple, K.A., 2011. Ejecta from impact craters. Icarus 211, 856 -- 875. doi:10.1016/j.icarus.2010.09.017. Hsieh, H.H., 2009. The Hawaii trails project: comet-hunting in the main asteroid belt. Astronomy and Astrophysics 505, 1297 -- 1310. doi:10.1051/0004-6361/200912342, arXiv:0907.5505. Hsieh, H.H., Yang, B., Haghighipour, N., 2012. Optical and Dynamical Characterization of Comet-like Main-belt Asteroid (596) Scheila. ApJ 744, 9. doi:10.1088/0004-637X/ 744/1/9, arXiv:1109.3477. Ishiguro, M., Hanayama, H., Hasegawa, S., Sarugaku, Y., Watanabe, J.i., Fujiwara, H., Terada, H., Hsieh, H.H., Vaubaillon, J.J., Kawai, N., Yanagisawa, K., Kuroda, D., Miyaji, T., Fukushima, H., Ohta, K., Hamanowa, H., Kim, J., Pyo, J., Nakamura, A.M., 2011a. Interpretation of (596) Scheila's Triple Dust Tails. ApJ Lett 741, L24. doi:10.1088/2041-8205/741/1/L24, arXiv:1110.1150. Ishiguro, M., Hanayama, H., Hasegawa, S., Sarugaku, Y., Watanabe, J.i., Fujiwara, H., Terada, H., Hsieh, H.H., Vaubaillon, J.J., Kawai, N., Yanagisawa, K., Kuroda, D., Miyaji, T., Fukushima, H., Ohta, K., Hamanowa, H., Kim, J., Pyo, J., Nakamura, A.M., 2011b. Observational Evidence for an Impact on the Main-belt Asteroid (596) Scheila. ApJL 740, L11. doi:10.1088/2041-8205/740/1/L11. Izawa, M.R.M., Flemming, R.L., King, P.L., Peterson, R.C., McCausland, P.J.A., 2010. Mineralogical and spectroscopic investigation of the Tagish Lake carbonaceous chondrite by X-ray diffraction and infrared reflectance spectroscopy. Meteoritics and Planetary Science 45, 675 -- 698. doi:10.1111/j.1945-5100.2010.01043.x. Jewitt, D., 2009. The Active Centaurs. The Astronomical Journal 137, 4296 -- 4312. doi:10. 1088/0004-6256/137/5/4296, arXiv:0902.4687. Jewitt, D., Agarwal, J., Weaver, H., Mutchler, M., Larson, S., 2013. The Extraordinary Multi-tailed Main-belt Comet P/2013 P5. ApJL 778, L21. doi:10.1088/2041-8205/ 778/1/L21, arXiv:1311.1483. Jewitt, D., Weaver, H., Mutchler, M., Larson, S., Agarwal, J., 2011. Hubble Space Tele- scope Observations of Main-belt Comet (596) Scheila. ApJ Lett 733, L4. doi:10.1088/ 2041-8205/733/1/L4, arXiv:1103.5456. 35 Kaiser, N., Aussel, H., Burke, B.E., Boesgaard, H., Chambers, K., Chun, M.R., Heasley, J.N., Hodapp, K.W., Hunt, B., Jedicke, R., Jewitt, D., Kudritzki, R., Luppino, G.A., Maberry, M., Magnier, E., Monet, D.G., Onaka, P.M., Pickles, A.J., Rhoads, P.H.H., Simon, T., Szalay, A., Szapudi, I., Tholen, D.J., Tonry, J.L., Waterson, M., Wick, J., 2002. Pan-STARRS: A Large Synoptic Survey Telescope Array, in: Tyson, J.A., Wolff, S. (Eds.), Survey and Other Telescope Technologies and Discoveries, pp. 154 -- 164. doi:10.1117/12.457365. Kleyna, J., Hainaut, O.R., Meech, K.J., 2013. P/2010 A2 LINEAR. II. Dynamical dust modelling. Astronomy and Astrophysics 549, A13. doi:10.1051/0004-6361/201118428, arXiv:1209.2210. Larson, S.M., 2010. (596) Scheila. iaucirc 9188, 1. Masiero, J., Jedicke, R., Durech, J., Gwyn, S., Denneau, L., Larsen, J., 2009. The Thousand Asteroid Light Curve Survey. Icarus 204, 145 -- 171. doi:10.1016/j.icarus.2009.06. 012, arXiv:0906.3339. Merline, W.J., Weidenschilling, S.J., Durda, D.D., Margot, J.L., Pravec, P., Storrs, A.D., 2002. Asteroids Do Have Satellites. Asteroids III , 289 -- 312. Moreno, F., Licandro, J., Ortiz, J.L., Lara, L.M., Al´ı-Lagoa, V., Vaduvescu, O., Morales, N., Molina, A., Lin, Z.Y., 2011. (596) Scheila in outburst: A probable collision event in the Main Asteroid Belt, in: EPSC-DPS Joint Meeting 2011, p. 31. O'Brien, D.P., Greenberg, R., 2003. Steady-state size distributions for collisional popula- tions:. analytical solution with size-dependent strength. Icarus 164, 334 -- 345. doi:10. 1016/S0019-1035(03)00145-3. Pravec, P., Harris, A.W., Kusnir´ak, P., Gal´ad, A., Hornoch, K., 2012. Absolute magnitudes of asteroids and a revision of asteroid albedo estimates from WISE thermal observations. Icarus 221, 365 -- 387. doi:10.1016/j.icarus.2012.07.026. Shanks, T., Belokurov, V., Chehade, B., Croom, S.M., Findlay, J.R., Gonzalez-Solares, E., Irwin, M.J., Koposov, S., Mann, R.G., Metcalfe, N., Murphy, D.N.A., Norberg, P.R., Read, M.A., Sutorius, E., Worseck, G., 2013. VST ATLAS First Science Results. The Messenger 154, 38 -- 40. 36 Snodgrass, C., Tubiana, C., Vincent, J.B., Sierks, H., Hviid, S., Moissi, R., Boehnhardt, H., Barbieri, C., Koschny, D., Lamy, P., Rickman, H., Rodrigo, R., Carry, B., Lowry, S.C., Laird, R.J.M., Weissman, P.R., Fitzsimmons, A., Marchi, S., OSIRIS Team, 2010. A collision in 2009 as the origin of the debris trail of asteroid P/2010A2. Nature 467, 814 -- 816. doi:10.1038/nature09453, arXiv:1010.2883. Spahr, T.B., Hergenrother, C.W., Larson, S.M., Campins, H., 1996. High Ecliptic Latitude Asteroid and Comet Surveying With the Catalina Schmidt, in: Rettig, T., Hahn, J.M. (Eds.), Completing the Inventory of the Solar System, pp. 115 -- 122. Stevenson, R., Kramer, E.A., Bauer, J.M., Masiero, J.R., Mainzer, A.K., 2012. Character- ization of Active Main Belt Object P/2012 F5 (Gibbs): A Possible Impacted Asteroid. ApJ 759, 142. doi:10.1088/0004-637X/759/2/142, arXiv:1209.5450. Stokes, G.H., Evans, J.B., Viggh, H.E.M., Shelly, F.C., Pearce, E.C., 2000. Lincoln Near- Earth Asteroid Program (LINEAR). Icarus 148, 21 -- 28. doi:10.1006/icar.2000.6493. Tonry, J.L., 2011. An Early Warning System for Asteroid Impact. PASP 123, 58 -- 73. doi:10.1086/657997, arXiv:1011.1028. Yang, B., Hsieh, H., 2011. Near-infrared Observations of Comet-like Asteroid (596) Scheila. ApJL 737, L39. doi:10.1088/2041-8205/737/2/L39, arXiv:1107.3845. Zolensky, M.E., Nakamura, K., Gounelle, M., Mikouchi, T., Kasama, T., Tachikawa, O., Tonui, E., 2002. Mineralogy of Tagish Lake: An ungrouped type 2 carbonaceous chon- drite. Meteoritics and Planetary Science 37, 737 -- 761. doi:10.1111/j.1945-5100.2002. tb00852.x. 37